You are on page 1of 148

Superconductivity

and
Low Temperature Physics
Part I: Superconductivity
Lecture Notes of the Academic Year 2013/14
Rudi Hackl and Dietrich Einzel
Walther-Meissner-Institut
Bayerische Akademie der Wissenschaften
Walther-Meissner-Strasse 8
D-85748 Garching
hackl@wmi.badw.de
c _ preliminary Garching, November 21, 2013
Contents
1 Introduction 1
1.1 A brief history of low-temperature physics . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Condensates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Present Status of Superconductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3.1 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3.2 Areas of Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3.3 Solved and Open Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2 Basic Experiments and Understanding 9
2.1 Key Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1.1 Zero Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1.2 Perfect diamagnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.1.3 Shubnikov phase (mixed state) . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.1.4 Quantization of the ux . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.5 Josephson effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.1.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2.1 Condensation Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2.2 Entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.2.3 Specic heat capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3 Electrodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3.1 The Drude model in the limit . . . . . . . . . . . . . . . . . . . . . . . . 24
2.3.2 Generalized London theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.3.3 The London equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.3.4 Some conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
iii
iv R. HACKL AND D. EINZEL CONTENTS
3 Microscopic Theory 35
3.1 The Cooper Instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.2 Origin of the interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.3 The BCS wave function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3.1 Coherent states in a boson eld . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.3.2 Properties of fermions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.3.3 A coherent state of fermions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.4 Determination of the ground state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.4.1 The BCS Hamilonian in second quantization . . . . . . . . . . . . . . . . . . . 45
3.4.2 Some expectation values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.4.3 Determination of the energy minimum at T = 0 . . . . . . . . . . . . . . . . . . 47
3.5 The general solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.5.1 Approximation of the four-operator term . . . . . . . . . . . . . . . . . . . . . 48
3.5.2 The Bogoliubov-Valatin transformation . . . . . . . . . . . . . . . . . . . . . . 49
3.5.3 Solution of the gap equation for T 0 . . . . . . . . . . . . . . . . . . . . . . . 52
3.6 Connection to experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.6.1 Thermodynamic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
3.6.2 Single particle response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4 Ginzburg-Landau Theory 59
4.1 Phase transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.2 Application to superconductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.2.1 Density of the free energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.2.2 Functional of the free energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.2.3 The Ginzburg-Landau equations . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.3 Two new length scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.3.1 Screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.3.2 Ginzburg-Landau coherence length . . . . . . . . . . . . . . . . . . . . . . . . 64
4.3.3 Energy of the normal-superconductor interface . . . . . . . . . . . . . . . . . . 65
4.4 States with internal ux . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.4.1 The upper critical eld B
c2
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.4.2 The nucleation eld B
c3
on the surface . . . . . . . . . . . . . . . . . . . . . . . 67
4.4.3 The thermodynamic critical eld B
c
. . . . . . . . . . . . . . . . . . . . . . . . 68
4.4.4 The lower critical eld B
c1
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.4.5 The Abrikosov lattice (1957) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
c _ Walther-Meiner-Institut
CONTENTS SUPERCONDUCTIVITY v
5 The Josephson Effect 73
5.1 Weakly coupled superconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.2 The Josephson equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
5.3 The RCSJ model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
5.4 Josephson contact in a microwave eld . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.5 Josephson effect in a magnetic eld . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.5.1 Ring with a single weak link . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.5.2 Ring with two weak links: Quantum interference . . . . . . . . . . . . . . . . . 79
5.5.3 Quantum interference in a long junction . . . . . . . . . . . . . . . . . . . . . . 80
6 Unconventional Materials 83
6.1 Classication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
6.2 The iron-age of superconductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
6.3 Copper-oxygen compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
6.3.1 History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
6.3.2 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
6.3.3 Physical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
6.3.4 Superconductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
6.3.5 Summary and perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
7 An overview of applications 113
7.1 Potential areas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
7.1.1 Economic considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
7.1.2 Areas of application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
7.2 Passive applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
7.2.1 Physical and technical challenges . . . . . . . . . . . . . . . . . . . . . . . . . 115
7.2.2 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
7.3 Active devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
2013
Preface
This is a rst version of lecture notes on superconductivity on an introductory level. It is intended to be
used along with the lectures I deliver at the Technical University Munich. The manuscript is the result of
several courses on the subject during the last years and should not be considered nal. In fact, it will need
a few more iterations. In particular, the referencing is not yet complete. The notes are inspired by several
textbooks and lecture notes on superconductivity or Condensed Matter Physics including the books by J.
F. Annett, N. Ashcroft and D. Mermin, W. Buckel and R. Kleiner, P.-G. de Gennes, R. Gro and A. Marx,
H. Kinder (notes of A. Heinrich), C. Kittel, A. Sudb and C. Fossheim, V. V. Schmidt (edited by P. M uller
and A. Ustinov), and M. Tinkham. We protted from numerous discussions with our colleagues working
in the eld, in particular at the Technical University Munich and at the Walther-Meissner-Institut.
The notes are not for distribution.
Garching, October 16, 2013 Rudi Hackl
vii
Chapter 1
Introduction
Imagine a macroscopic object which obeys the rules of quantum mechanics and can be described by a
single wave function. What would be the consequences? This question is at the heart of the phenomena
to be discussed here. An intuitive answer can be found by considering the usual microscopic quantum
world. The wave function describing the electron in a hydrogen atom, for instance, has an amplitude
and a phase. The amplitude corresponds to the probability of nding the electron in a volume element
around the pointr. The phase determines the type of the stationary state dictated by the uniqueness and
the number nodes of the wave function. In the stationary state there is no dissipation. The superposition
of stationary states leads to interference phenomena and transitions with typically very sharp energies.
Superconductivity and superuidity along with Bose condensation or density wave order (periodic mod-
ulations of the magnetization or charge density) are macroscopic quantum phenomena. In all cases the
establishment to a coherent quantum state facilitates charge, mass or spin transport without energy dis-
sipation. For appropriately prepared experimental settings interference phenomena can be observed. In
the case of superconductivity the characteristic energy is solely determined by constants of nature. Elec-
tromagnetic radiation can be created and extremely sensitive detectors can be constructed which exploit
the rigid phase of the wave function.
The emergence of quantum phenomena on a macroscopic scale, i.e. in a large ensemble which can be
described by statistical methods, can be expected when temperature is not the dominant energy scale any
further. Therefore, the discovery of macroscopic quantum phenomena in the early 20
th
century is closely
related to the development of low temperature techniques although meanwhile a variety of manifestations
have been found to exist also at room temperature and above.
In this chapter we rst sketch briey the hallmarks of physics at low temperature (with the focus
placed on superconductivity), then qualitatively describe examples of condensates and nally provide
an overview over the current status of superconductivity.
1.1 A brief history of low-temperature physics
The liquefaction of air in 1895 [?] can be considered the starting point of low-temperature physics. It took
only 13 years until Heike Kamerlingh
1
Onnes in Leiden succeeded to liquefy Helium, the gas with the
lowest boiling point [?]. Not only does He have the lowest boiling point of all elements but it is also the
only material which does not solidify upon further cooling. Only at an applied pressure of approximately
30 bar the solidication can be induced. This phenomenon is among the rst manifestations of quantum
1
Privy Councillor (Hofrat)
1
2 R. HACKL AND D. EINZEL Introduction
(a) (b)
Figure 1.1: Resistances of metals at low temperature. Note the different temperature scales in (a) and
(b). (a) Proposals for the low-temperature resistivity of metals as of 1900. Temperatures below 15 K
were unaccessible. Curves 1 through 3 were suggested by Dewar, Matthiessen, and Kelvin, respectively.
From W. Buckel, Supraleitung VCH Weinheim 5th Edition (1994), Fig. 1. (b) Resistance of Hg close to
the superconducting transition. From Ref. [1].
effects directly determining the macroscopic properties. If one estimates the amplitude of the zero-point
oscillations on the basis of the uncertainty principle values on the order of 10 to 20% of the atomic
diameter are found which are big enough to prevent the establishment of a crystal lattice.
Only three years later Onnes observed vanishing resistance upon cooling mercury (Hg) below 4.2 K. [1]
At that time the low temperature properties of metals were discussed intensively since the experimental
data ended at some 15 K being close to the solidication point of liquid hydrogen. There were three
main concepts Fig. 1.1 (a): (1) Probably inspired by the Drude model
2
Dewar argued in 1904 that
should asymptotically approach 0 in the limit T 0 since all scattering mechanisms might freeze out.
(2) Matthiessen
3
observed that the increase of the resistivity of a metal as a consequence of a small
concentration of another metal in a solid solution was temperature independent. This implies that the
resistivity of a metal can stem from more than one source, for example a constant one from impurities,

0
, and a temperature dependent one of different origin, (T), which add up.
4
. As a consequence
the resistivity saturates at low temperature at a nite
0
, the residual resistivity. (3) Alternatively, the
resistivity was proposed to diverge due to electron localization (Lord Kelvin 1902).
5
In the experiment on very clean Hg, Onnes and his collaborators unexpectedly found the resistance to
become abruptly unmeasurably small at a nite critical temperature T
c
as shown in Fig. 1.1 (b) rather
than asymptotically in the limit T 0.
He called this qualitatively new phenomenon superconduction. It was clear from the beginning that none
of the three scenarios discussed above was capable of describing the observation. It took until 1957 until
Bardeen, Cooper, and Schrieffer succeeded in presenting a microscopic theory of the superconducting
2
P. Drude 1900
3
A. Matthiessen, Ann. Phys. Chem. (Pogg. Folge) 110, 190 (1860) and A. Matthiessen and G. Vogt, Ann. Phys. Chem.
(Pogg. Folge) 122, 19 (1864)
4
Matthiessens rule is purely empirical and valid only if the different scattering mechanisms are independent and isotropic.
For a discussion see Ashcroft and Mermin Solid State Physics (Holt-Saunders International Edition 1981)
5
Anderson showed in 1958 (Phys. Rev. 109, 1492) that even a small amount of impurities can localize electrons due to
interference of Bloch waves. Even in a completely pure system (with low carrier density) the resistivity can diverge due to the
crystallization of the electrons (Wigner).
c _ Walther-Meiner-Institut
A brief history of low-temperature physics SUPERCONDUCTIVITY 3
state [2]. A discovery of enormous importance and inuence was the demonstration of thermodynamic
nature of the superconducting phase by Ochsenfeld and Meissner in 1932. Soon thereafter in 1935,
Fritz and Heinz London suggested the rst phenomenology [?] and proposed a rst quantum mechanical
approach in 1950. In the same year Ginsburg and Landau applied Landaus theory for phase transitions
to superconductors using a complex order parameter rather than a real one such as for magnetism or
lattice distortions. In the year of publication of the microscopic theory Abrikosov used the Ginsgurg
Landau phenomenology to derived the existence of a mixed state accommodating superconductivity and
a magnetic eld on a mesoscopic scale [?] as observed by Shubnikov and coworkers already in 1936 [?].
This discovery paved the way towards a huge variety of applications since electric currents and very large
magnetic elds can coexist here with superconductivity facilitating the quasi-dissipationless maintenance
of magnetic elds in excess of 20 T. Finally in 1960, Eliashberg made connection to real materials in his
celebrated strong-coupling approach [3] which is used to date for analyzing interaction potentials in
superconductors.
In 1938 a phenomenon analogous to superconductivity was found in an uncharged system. Here, mass
ows without viscosity. This property is called superuidity and was independently discovered by
Kapitsa
6
and by Allen and Misener
7
in Helium below 2.17 K. Rotation corresponds to the magnetic eld
and vortices of electric current around lines of the eld (see section ??) to vortices of the liquid.
If two lumps of a superconductor or two reservoirs of a superuid are weakly connected in a way that
only the supracomponent can pass, the eventually occurring voltage and pressure drops, respectively,
correspond. The spectacular fountain effect (Fig. 1.2) is a result of the pressure difference across a so-
called weak link (realized by ne powder in the lower part of the nozzle which is transparent only for
the superuid) induced by a temperature difference between the reservoirs (by heating the upper part
close to the narrow nozzle). The corresponding phenomenon in a superconductor is the Josephson effect
predicted 1962 [?] and observed one year later [?]. It is one of the most spectacular manifestation of the
quantum nature of the superconducting condensate and became the basis of a highly sensitive detector
of magnetic elds with a broad range of applications [?].
In 1972, Lee, Osheroff, and Richardson discovered superuidity also in the lighter isotope
3
He. The
transition temperature T
c
= 2.2 mK is three orders of magnitude smaller than that of the of the heavier
isotope.
8
At rst glance, superuidity in the two isotopes appears to be only quantitatively different.
However, quantum effects dominate the behavior at these temperatures and the nuclear spin of 1/2 pre-
vents
3
He from Bose condensation occurring in S = 0
4
He. As opposed to a condensate of strongly
interacting
4
He atoms
3
He forms a pair condensate similar to that of a superconductor. In either case
the condensate is neutral. Only one year later A. Leggett proposed a full theoretical description of the
various phases of
3
He [4].
One of the most important technical innovations, proposed in 1962 and rst demonstrated in 1965, was
the realization of the dilution refrigerator using mixtures of
3
He and
4
He. Here, the cooling is achieved
by dissolving
3
He in
4
He. The minimal temperature of this continuously working cryo-system is below
10 mK. The dilution refrigerator is the basis of many ultra-low-temperature experiments and replaced the
cooling by paramagnetic salts almost completely. If a nuclear demagnetization stage is attached a record
base temperature of 8-10 K was realized. Lower temperatures can be reached only in cold gases (see
below). In the range between 10 K and 1 K quite a few new superconductors and unexpected phases of
matter were found as described in the textbooks and to be discussed partially in the second part of the
lecture series.
In these days dilution refrigerators do not need pre-cooling with liquid
4
He any further as in the early
days. With the continuous improvement of closed-cycle refrigerators K can now be generated by
6
P. Kapitsa, Nature 141, 74 (1938)
7
J.F. Allen and A.D. Misener, Nature 141, 75 (1938)
8
Lee, Osheroff, and Richardson, Phys. Rev. Lett. vv, ppp (1972)
2013
4 R. HACKL AND D. EINZEL Introduction
Figure 1.2: Bose-Einstein condensation. (left) Demonstration of the fountain effect in superuid He.
The Photograph was taken by Jack Allen in 1970 who co-discovered superuidity. From R.J. Don-
nelly, Physics Today, July 1995, p. 30. (right) Bose-Einstein condensation of
87
Rb at 500 nK. The
series of pictures shows the momentum distribution of Rubidium atoms in a trap at different tempera-
tures. If the majority of the atoms is in the lowest state the total momentum approaches zero. From
http://www.mpq.mpg.de/cms/mpq/en/departments/quanten/homepage cms/projects/bec/.
combining a pulse-tube unit [?, ?] with an optimized dilution system. The prototype of this innovation
was developed at the Walther-Meissner-Institut [?] and is now commercially available.
Another important discovery at low temperature is the integer and fractional quantum Hall effect which,
in connection with the Josephson effect, completes, for instance, the metrological triangle. Spin and
charge density waves SDW/CDW are ordering phenomena very similar to the superconducting state.
Here, the ordering produces a periodic modulation of the carrier or spin density [5]. The modulation
amplitude can vary between a fraction of a percent and several ten percent of the average charge density.
While in a superconductor two electrons form a Cooper pair having vanishing net momentum density
waves are characterized by a nite ordering vector connecting particle and hole states.
The most spectacular discoveries of the late 20
th
and the early 21
st
century are the Bose-Einstein
condensation of trapped gases [?], the observation of superconductivity above 100 K and 40 K in the
cuprates [?, 6] and, respectively, in iron pnictides [7, 8]. These most recent developments demonstrate
that low-temperature physics in general and superconductivity in specic remain vibrant elds of con-
densed matter physics.
A timetable displaying the history in consecutive order is planned to be provided soon in the appendix.
1.2 Condensates
Schr odinger was the rst to elaborate on a coherent state of Bosons [?] and constructed a coherent
wave function from a superposition of an innite number of harmonic oscillator wave functions [
n
=
1/

n!( a

)
n
[0 with a

the creation operator and [0 denoting the ground state,


[ = e

[[
2
2

n=0

n!
[
n
= e

[[
2
2

n=0
( a

)
n
n!
[0. (1.2.1)
c _ Walther-Meiner-Institut
Present Status of Superconductivity SUPERCONDUCTIVITY 5
Here, = [[e
i
is a complex number in polar representation having the amplitude [[ and the phase
. As shown in problem 1 [ obeys a Poisson distribution which is sharply peaked at the average oc-
cupation number n = N. In addition, one nds an uncertainty relation between the average occupation
number N and the phase to hold implying that for a large number of particle / = 1/

N.
Coherent states have this structure in general. Here the wave function is constructed for Bosons and can
directly applied for the description of the laser eld or for Bose-condensed cold atoms. In principle, it
is also applicable for
4
He but strong coupling effects must be considered in this case and make affairs
much more complicated. Is the wave function also useful for Fermionic systems?
It is the achievement of Schrieffer to construct a coherent state similar to that in Eq. (1.2.1) from the
Fermionic wave function of electrons. The nature of Schrieffers state will be derived in detail in chap-
ter 3. The origins of a superconducting and a Bose condensate are quite different. Superconductivity
originates from pair correlations between the electrons
9
In the case of Helium or, similarly, for vapors
of alkali metals
10
the individual atoms undergo a Bose-Einstein condensation at sufciently low tem-
peratures, which means that a macroscopic number of particles assumes the lowest possible state. The
difference between the alkalis and Helium is just the interaction strength V
e f f
. In the weak-coupling
limit the rst excited state has essentially the same energy as the ground state. With increasing coupling
a gap develops between the ground state and the rst excited state. Then, following the argumentation
of Landau, condensed particles can be accelerated to a nite critical velocity v
c
before they leave the
ground state and dissipate energy. In spite of the fundamental differences between superconductors and
Bose condensed systems all form a coherent ground state which can be described by a single quantum
mechanical wave function.
1.3 Present Status of Superconductivity
1.3.1 Materials
The development was continuous but quite slow in the beginning. While many elements become su-
perconducting as shown in Fig. 1.3 the transition tempertures are moderate, and Nb having T
c
= 9.26 K
holds the record. If pressure is applied even semiconductors may become superconducting. At pres-
sure values in the range of 20-150 GPa (0.2-1.5 Mbar) even insulators or alkali metals occasionally reach
T
c
values between 15 and 25 K (see Table ??). In Fig. 1.3, the elements with pressure-induced super-
conducting transitions are marked with red points. On the other hand, compounds soon exceeded the
transition temperatures of the elements as shown in Fig. 1.4 and Table ??. This was the time period
when materials sciences ourished and many compounds and new properties and effects were discov-
ered which demonstrated the potential of applications while the transition temperatures T
c
increased at a
very low rate of 0.4 K/year (Fig. 1.4). The discovery of superconductivity in the copper oxygen system
La
2x
Ba
x
CuO
4
was a qualitative change in many respects.
11
The slow increase of T
c
which in a sense
was the linear analogue of Moores law was interrupted in the non-Murphyian direction. Most impor-
tantly, the superconducting state turned out to be unconventional meaning that in addition to the gauge
symmetry also the rotational symmetry is broken at the transition. In other words, the order parame-
ter and the energy gap have a lower symmetry than the crystal. Unconventional superconductivity was
not completely new
12
but was considered an exception occurring only in a few exotic systems such as
9
We note here, that Bogoliubov had results equivalent to those of BCS at approximately the same time but the publication
was delayed.
10
Cornell, Ketterle, and Wieman
11
J.G. Bednorz and K.A. M uller, Z. Phys. B (1986)
12
F. Gro et al., Z. Phys. B (1986). It is a coincidence with some charm that the article of Bednorz and M uller on
La
2x
Ba
x
CuO
4
directly follows the rst clear evidence of an unconventional gap with non-zero orbital momentum in UPt
3
and UBe
13
.
2013
6 R. HACKL AND D. EINZEL Introduction
http://www.spring8.or.jp/en/news_publications/press_release/2011/110705_fig/fig1.png[18.10.2013 13:56:26]
Figure 1.3: Transition temperature T
c
of the elements. The latest record is T
c
25 Kin Ca at 210 MPa [?].
Nb having T
c
=9.26 K is the elements with the highest transition under normal conditions. Higher values
were obtained only with applied pressure. From http://www.spring8.or.jp.
150
T
c

(
K
)
HgBa
2
Ca
2
Cu
3
O
8+
150
t
u
r
e

T
g
2 2 3 8+
Bi Sr Ca Cu O
Tl
2
Ba
2
Ca
2
Cu
3
O
10+
100
p
e
r
a
t
YBa
2
Cu
3
O
6+x
Bi
2
Sr
2
Ca
2
Cu
3
O
10+
3 NdFeAsO
1-x
F
x
50

n

t
e
m
La
2-x
Ba
x
CuO
4
MgB
2
3
2 Ba
1-x
K
x
Fe
2
As
2
1 LaFePO
1 x x
n
s
i
t
i
o
n
Hg
NbTi
V
3
Si
Nb
3
Ge
Ba K BiO
Cs
2
RbC
60
2
1900 1925 1950 1975 2000 2025
0
t
r
a
n
f di
Ba
0.6
K
0.4
BiO
3
1
year of discovery
Figure 1.4: Maximal transition temperature T
c
vs. year of discovery. The discovery of superconduc-
tivity in the cuprates is a historical hallmark. Superconductivity in Fe-based compounds was similarly
unexpected.
the heavy fermion compounds (e.g., UPt
3
) or, more recently, in the ruthenates (e.g., Sr
2
RuO
4
)
13
or in
superuid
3
He.
The cuprates, on the other hand, cover a very broad class of materials which have maximal transition tem-
peratures T
max
c
in a range from about 20 K in Bi
2
Sr
2
CuO
6
to 160 K in HgBa
2
Ca
2
Cu
3
O
8
. Surprisingly
enough, apart from T
max
c
the temperature versus doping phase diagrams are similar: all are antiferro-
13
G.M. Luke et al., Nature 394, 558 (1998)
c _ Walther-Meiner-Institut
Present Status of Superconductivity SUPERCONDUCTIVITY 7
magnetic insulators at half lling and superconductivity exists for 0.05 < p < 0.27 with p the number
of holes per CuO
2
formula unit (away from the Cu3d
9
conguration).
14
While the rst applications are
already on the market, the orgin of superconductivity is still an open question at least for the majority of
the people working in the eld. High-T
c
research is indeed among the most vibrant elds in solid state
physics in these days, and we devote a complete chapter to it.
Since the discovery of the cuprates several other quite interesting materials were found to superconduct.
The biggest surprise was certainly MgB
2
with T
c
= 39.5 K.
15
Very recently, also elements were found
to become superconducting at temperatures close to 20 K under sufciently high pressures.
16
In January
2008 superconductivity was reported in Nd(O
1x
Fe
x
)FeAs compounds.
17
By exchanging the rare earth
a maximal T
c
of 55 K was obtained.
18
1.3.2 Areas of Application
It is immediately clear that the state of zero resistance offers enormous opportunities for applications.
The rst realization was a solenoid for elds in the range of 5 T in a volume of a few cubic-centimeters.
In these days, the maximal permanent eld maintained by a superconductor is slightly above 21 T, and the
largest volume is several m
3
in coils for fusion experiments. Applications include now power transmis-
sion, generators, motors, energy storage, fault-current limiters, and lters. The most popular application
which many people have already encountered are superconducting coils for magnetic resonance imaging.
Beyond the state of R = 0 there are various other properties of superconductors which can be exploited.
For instance, the behavior of weakly coupled superconductors opens a wide eld of active devices which
rest on the Josephson effects. For sufciently low currents there is no voltage drop across the weak
link (dc Josephson effect), and the phase difference between the two superconductors remains constant.
For higher currents there is also a voltage drop making the phase difference time dependent. It follows
from the Josephson equations that the current has now oscillating components (ac Josephson effect).
Exploiting the dc effect extremely sensitive detectors for magnetic elds can be made (Superconducting
QUantum Interference Device) which allow one to observe and to locate the currents in the heart and in
the brain. The ac effect can be used to produce and detect electromagnetic waves in the hundred GHz
range. Other very important applications are the voltage standard and, along with the quantized Hall
effect, the determination of Plancks constant h and the elementary charge e.
1.3.3 Solved and Open Problems
The London brothers could clarify the phenomenology and electrodynamics of superconductors rela-
tively soon after the discovery. In 1935 they proposed a model which was capable of describing the
exclusion of magnetic elds from bulk superconductors and discussed the exponential decay of the eld
in a surface sheeth of thickness
L
which is now called the London penetration depth. They understood
that the effect was quantum mechanical in origin and describable in terms of a single wave function be-
ing closely related to coherent states as proposed by Schr odinger in 1926.
19
However, the microscopic
foundation of this wave function was completely unclear in particular since Schr odingers proposal was
for Bosons and not for Fermions. It took another 22 years until Schrieffer could derive a coherent state
wave function also for Fermions which nally completed the BCS theory of superconductivity [2].
14
We note that the AF ordering (N eel) temperature T
N
and T
max
c
do not scale.
15
Akimitsu et al., Nature vvv, ppp (2001).
16
J.J. Schilling, Schrieffers book, ppp (2006).
17
abc et al., J. Am. Chem. Soc. vvv, ppp (2008).
18
abc et al., Nature vvv, ppp (2008).
19
E. Schr odinger, Naturwissenschaften 28, (1926).
2013
8 R. HACKL AND D. EINZEL Introduction
The BCS theory considered only the limit of weak coupling. Eliashberg succeed to establish a con-
nection between the interaction potential V
k,k
/ and spectra density of phonons. This became possible
after Migdals observation that the electron-phonon vertex is of order
_
m/M with the electron and ion
masses m and M. Physically speaking this means that the electron-phonon interaction is retarded or that
the respective energy scales differ by orders of magnitude and can be considered independent. Eliash-
bergs approach can be generalized to other type of interaction as long as the interaction between two
electrons is bosonic. Therefore, the scheme is used today to analyze the superconducting properties even
of cuprates and pnictides. However, while qualitative conclusions are enlightening a quantitative de-
scription including the theoretical derivation of the transition temperature of an existing superconductor
or the prediction of a new superconducting material are still at their infancy. Some progress was made
for conventional systems recently [9]. In spite of enormous progress the understanding of the cuprates is
still limited. One of the major problems is the similarity of all relevant energy scales such as the Fermi
energy E
F
, the exchange coupling J, the energy gap
k
, or the phonon energy h
q
. The solution of these
type of problems is among the most tantalizing questions in condensed matter physics.
c _ Walther-Meiner-Institut
Chapter 2
Basic Experiments and Understanding
Many of the basic observations in superconductors to be described in the rst section of this chapter
could be understood phenomenologically long before the microscopic theory was nally presented in
1957. In particular the London theory highlights the different aspects of the zero-resistance state and the
perfect diamagnetism. If derived from the Schr odinger equation time-dependent phenomena such as the
Josephson effects follow directly highlighting the quantum mechanical nature of the phenomenon in a
simple though instructive way.
2.1 Key Experiments
2.1.1 Zero Resistance
After Onnes succeeded to liquefy Helium (1908) he started immediately to study the resistivities of
metals. Hg was the purest material available at that time. Instead of nding support for one of the
proposals being discussed at his time (see chapter I) he discovered a state of vanishingly small resistance
(R < 10
5
) below T = 4.2 K, the boiling point of liquid He (Fig. ??).
1
The determination of the actual resistivity of a superconductor requires new techniques since the pre-
cision of a direct four-probe measurement is orders of magnitude too low. Already Onnes designed an
experiment which uses the magnetic moment of a persistent current I in a ring to determine the decay rate
of I rather than the resistance itself. In the experiment there are two concentric superconducting rings.
The outer one is xed, the inner one is suspended on a quartz lament. As long as there is no current in
the rings the inner ring turns with the lament. Now we get supercurrents I
i
and I
o
to ow in the rings.
To this end a (for simplicity) homogeneous eld B
0
B
c
(T) is turned on with the axis parallel to the
axes of the concentric rings. If the rings are above their transition temperature T
c
the eld penetrates the
cross sections S
i
and S
o
as well as the material homogeneously. Now the rings are being cooled below T
c
and the eld is expelled from the material due to the Ochsenfeld-Meissner effect (see next paragraph),
i.e. is slightly distorted around the solid superconductors but more or less unchanged in most of the free
space.
What happens if the eld is switched off? This can be derived easily from Faradays law for either of the
rings,
E =
B
t
. (2.1.1)
1
H. Kammerlingh Onnes, Leiden Comm. 120b, 122b, 124c (1911)
9
10 R. HACKL AND D. EINZEL Basic Experiments and Understanding
We integrate over the cross section S of the respective ring and apply Stokes theorem,

_
S
B
t
dA =
_
S
E dS (2.1.2)
=
_

E dl.
Here is the edge of S and runs approximately along the center of the ring. Since the resistance of the
superconductor vanishes for sufciently small elds (see below) there is no voltage drop along the ring
and, consequently,

t
= 0 (2.1.3)
In the last equation BS is the ux penetrating the ring, and the voltage drop for one circulation
around the ring is zero since R = 0. Hence the ux through a closed superconducting loop is constant
and cannot escape after the external eld is switched off. The related eld is maintained by a current I in
the ring. In turn, we would be unable to get the current running when applying the eld at T < T
c
. For
this reason it is misleading to think of inducing a current by changing an external eld. The current is
rather a result of maintaining the ux xed.
If the lament is now twisted there will be a restoring force proportional to
i

o
with the moments
given by = SI. Once the equilibrium position is reached any turn of the inner ring as a consequence of
a variation of I
i,o
can be measured with great precision. In order to determine a limit for the resistivity
from I we consider the ring as a loop with a resistance R and an inductance L for which
RI +L
dI
dt
= 0. (2.1.4)
This differential equation can e integrated by separating the variables, and we nd
R =
L
t
ln
I(t)
I(0)
, (2.1.5)
where I(t) is directly proportional to the torque. Onnes and his collaborators kept the experiment running
for a year - quite heroic at that time. With modern NMR techniques the eld produced by a current can
be measured with much higher precision than with a torque meter, and a lower limit of 10
5
years was
found for I to become signicant.
2.1.2 Perfect diamagnetism
Given the perfect conductivity of a superconductor it is obvious that a magnetic eld is completely
shielded when it is applied below the transition temperature. The induced screening currents do not
decay as in conventional metals with nite resistivity. However, what happens if the superconducting
state is entered by lowering the temperature in a nite external eld?
The question was answered by Meissner and Ochsenfeld in 1933 when they studied the force between
two parallel tin wires. If the currents were applied above and below T
c
the forces were different. In the
latter case the mutual exclusion of the elds due to screening explained the result. However, the same
effect was observed when the wires were cooled through T
c
with the currents on implying that the eld
was also expelled without induction to start the screening currents
2
.
2
W. Meiner and R. Ochsenfeld, Naturwissenschaften 81, 787 (1933)
c _ Walther-Meiner-Institut
Key Experiments SUPERCONDUCTIVITY 11
B
i
=0
B
i
=B
a
( )
Type I SC
T 0
B
0
B
i
B
c
B
i
= B
0
B
i
= 0
Meissner
state
normal
state
(b)
T T
c
B
B
c
(0)
B
c
(T)
zero field cooled (z.f.c.)
field cooled (f.c.)
(a)
Figure 2.1: Field-temperature phase diagram of a superconductor. A given point in the superconduct-
ing can be reached independent of the path. In this way the superconducting state is established as a
thermodynamic phase.
Therefore, the second actually dening property of a superconductor is its perfect diamagnetism: a mag-
netic eld which is sufciently small to not destroy the new state is expelled from a singly connected
specimen irrespective of the route the superconducting state is entered (Fig. 2.1). Hence, superconduc-
tivity is a thermodynamic phase which, in turn, implies the existence of a critical eld B
c
(T) dening the
phase boundary between the normal and superconducting states in the presence of a eld and determin-
ing the condensation energy (see section 2.2). The eld free state is usually called Meissner state. The
magnitude of critical eld can be approximated by a parabola,
B
c
(T) = B
c
(0)
_
1
_
T
T
c
_
2
_
. (2.1.6)
The microscopic theory arrives at a similar phenomenology but there are deviation which depend on the
electron-phonon coupling strength.
The path-independent complete exclusion of magnetic ux from the interior of a bulk superconductor
in the Meissner state raises several questions: (i) How does the eld (induction B =
0
H) change at the
surface? (ii) How are the screening currents set off if not according to Faradays law? (iii) Is it possible
that the ux penetrates partially to reduce the magnetic energy? (iv) What happens in a material which
is not simply connected, i.e. has voids?
The level of complication in nding answers to these questions differs remarkably. (ii)-(iv) can be an-
swered only qualitatively here and will be discussed in detail in the following chapters. (i) follows from
the phenomenological London theory (see section ??). Electrodynamics shows directly that the applied
eld decays exponentially away from the surface into the bulk. The same holds true for the screen-
ing currents. The characteristic length scale is called the London penetration depth
L
and is of order
20. . . 200 nm. The range can be traced back to the carrier concentration. Pippard observed a dependence
of the penetration depth on the purity of a material and concluded that non-local electrodynamics must
be used for properly explaining this effect. This means that the response at a point r depends on the
perturbation in a material dependent volume of order [r
0
[
3
. The new characteristic length scale
0
was explained microscopically 20 years later by Bardeen, Cooper, and Schrieffer (BCS theory) but also
anticipated in the framework of the Ginzburg-Landau phenomenology.
(ii) is also a consequence of the London equation but the microscopic foundation was found only in 1957
by BCS. The reason for the Meissner effect is the existence of a uniform phase of the superconducting
electrons being established at the transition. The current is proportional to this phase modied by the
vector potential A of the applied eld B
0
= A, j ( eA). From a classical point of view the
force on the electrons is transverse as for the Lorentz force but exits only in quantum-mechanical systems
with a coherent phase for all carriers, where a nite vector potential sets off a current irrespective of
2013
12 R. HACKL AND D. EINZEL Basic Experiments and Understanding
the history. The uniform phase was anticipated by Fritz and Heinz London already in 1935 without a
theoretical justication. Also Ginzburg and Landau used a complex order parameter with a single-valued
phase to construct the Free-Energy functional but did not discuss its origin. It needed the Geistesblitz
(ash of genius) of Bob Schrieffer to realize the necessity of a coherent state of fermions and to construct
it in a fashion similar to that proposed by Schr odinger for bosons in 1926. So we have to understand the
full microscopic theory (see chapter 3) to properly explain the Meissner effect. A different approach is
possible via a thermodynamic argumentation which we shall discuss in chapter 4.
The full exclusion of magnetic ux is associated with a rapidly increasing energy E = V/2
0
B
2
with
V the sample volume. If we recall that the critical elds are moderate in elemental superconductors the
new phase will not survive sufciently long to make the materials particularly useful for applications. Is
there a way around as insinuated in (iii)?
There are two qualitatively different answers. The rst one is an effect of the demagnetization as de-
scribed in Appendix 1. Around superconducting samples having shapes different from innite slabs
or cylinders oriented along H
0
the eld changes close to the surface once the Meissner state is estab-
lished. This phenomenon originates in the usual electrodynamic relation B = 0 and the condition
n(H
0
H
i
) = 0 describing the continuity of normal and, respectively, the tangential components of
the elds around a superconductor. Whenever the demagnetization factor n is different from 0 the sur-
face enhancement of the eld around a generally shaped superconductor implies that the critical eld
is reached earlier in locations with enhanced eld than in those without. In the case of a sphere the
enhancement is 1/3 at the equator with the equatorial plane perpendicular to the homogeneous eld with
magnitude B
0
. Consequently the eld starts penetrating for B
0
= 2/3B
c
. In the range 2/3B
c
< B
0
< B
c
,
the sample is in the intermediate state with normal and superconducting regions coexisting on a macro-
scopic scale. If a slab (for instance a thin-lm sample) with area S a
2
and thickness t much smaller
than a is aligned perpendicular to the eld n is close to unity and the eld becomes very large at the edges
rapidly exceeding the critical eld, and ux penetrates even for very small applied elds B
0
B
c
. For
stability reasons the eld in the normal regions must be B
c
and vanishes in the superconducting parts.
For satisfying the conservation of ux ( B = 0) everywhere the density of the superconducting regions
scales with the ratio of the applied to the critical eld as
sc
= 1(B
0
/B
c
).
If the intermediate state is studied with a decoration experiment, with magneto-optics or neutrons one
nds a regular pattern of the normal and superconducting regions even in a perfect crystal. The regions
are aligned along high-symmetry lines of the crystal lattice. It is in fact the electronic structure which
denes preferential orientations as will become plausible in chapter 3. The second part of the answer
refers to a new state which is sufciently important to be discussed in a new paragraph.
2.1.3 Shubnikov phase (mixed state)
The energy needed to keep the magnetic eld completely outside the superconducting volume increases
quadratically with the eld. In turn, a complete penetration quenches superconductivity. Shubnikov
showed 1936 that a small amount of Tl in Pb leads to a new mixed state with some ux starting to
penetrate the material long before superconductivity vanishes [?] The experimental magnetization curves
for a Pb single crystal with 5%Pb replaced by Tl are shown in Fig. 2.2. For the cylindrical shape of the
sample a demagnetization effect can be excluded. Later it turned out that practically all non-elemental
superconductors and Nb share the property of normal regions encircled by superconducting screening
currents around the eld lines [panel (b)]. Each vortex carries only the smallest amount of ux possible
(see next paragraph). There is a lower critical eld B
c1
at which the eld starts to penetrate and an upper
critical eld B
c2
at which the material becomes a normal metal. The creation of the rst ux line at B
c1
does not cost energy. Therefore the magnetization decays at an innite rate as shown in panel (c) of
Fig. 2.2. At B
c2
the magnetization vanishes at a nite rate.
c _ Walther-Meiner-Institut
Key Experiments SUPERCONDUCTIVITY 13
mixed state
Meissner state
Type II SC
-
0
M
B
0
B
c1
B
c2
(a) (b)
(c)
Figure 2.2: Type II superconductor in a magnetic eld. (a) Induction B vs. applied eld of a cylindrical
sample of Pb
95
Tl
5
[?]. There is no jump of the induction at a critical eld but rather a sharp onset at B
c1
and a continuous approach to the normal state induction. At B
c2
the superconducting induction branches
of the normal one with a different slope. (b) Above B
c1
the partially penetrates the material. Instead of
running around the perimeter of the samples the screening currents encircle individual ux lines, each
carrying
0
, and form vortices. (c) Schematic plot of the magnetization
0
M = B
i
B
0
vs. B
0
. Below
B
c1
B
i
= 0. The rate of reduction of
0
M is logarithmically divergent indicating that the energy need
for the rst ux line is vanishingly small. At B
c2
the slope is nite.
The existence of the mixed state separates the superconductors in two classes. Those with complete
Meissner effect are usually called type I while type II materials are characterized by a Meissner state at
low elds and a mixed state between B
c1
and B
c2
. The innite slope of
0
M at B
c1
and the complete
reversibility of the magnetization indicate that the ux lines can move freely in the mixed state. Conse-
quently, even though the eld at which superconductivity vanishes is occasionally very high, also type II
superconductors still appear to be of little practical use since any current leads to a movement of the
ux lines due to the Lorentz force and to dissipation. Only when the pinning of ux lines in disordered
alloys such as NbTi or by the introduction of articial pinning center was achieved applications emerged
rapidly. Materials with strong pinning are called magnetically hard or sometimes type III superconduc-
tors. The applications aspects were among the reasons why Abrikosov earned the 2003 Nobel prize for
his theoretical description of the mixed state on the basis of the Ginzburg-Landau theory (chapter 4). The
study of the ux line lattice remains an important eld of research into superconductors.
As briey mentioned above the vortices carry a ux which, for energetic considerations, is small. From
a topological point of view a superconductor perforated by normal conducting regions is not simply
connected any further. Hence, if a persistent current having a quantum mechanical nature encircles a
vortex quantization effects can be expected. Similar effects can be expected when a macroscopic sample
is not simply connected. This brings us nally to a discussion of question (iv) of paragraph 2.1.2.
2.1.4 Quantization of the ux
When electrons move freely they do not change their quantum state. Therefore, if they form a non-
dissipative current in a ring one can expect the angular momentum to be quantized according to the
Sommerfeld condition. This was the reasoning following Fritz Londons prediction of 1950 when the
2013
14 R. HACKL AND D. EINZEL Basic Experiments and Understanding
B.S. Deaver and W.M. Fairbank, PRL 7, 43 (1961)
R. Doll and M. Nbauer, PRL 7, 51 (1961)
Figure 2.3: Experimental setup for (left) and results of ux quantization measurements [?, ?]. The mag-
netic ux trapped in the superconducting lead cylinder is measured with a resonance method. Once the
cylinder is below T
c
and the static eld in y-direction B
0
is switched off the eld B
M
starts oscillat-
ing at the resonance frequency of the cylinder. The oscillation amplitude as detected with a mirror is
proportional to the magnetic moment of the cylinder, hence the included ux.
experiments on the ux quantization in multiply connected superconductors were discussed in the late
1950ies and early 1960ies. The experiments were performed simultaneously by two groups which did not
have information about each other. Deaver and Fairbanks [?] in Stanford used a vibrating sample (Foner-
type) magnetometer for determining the trapped ux in a small cylinder. In the experiment performed
by Doll and N abauer at the Bavarian Academy of Sciences and Humanities (shown in Fig. 2.3 l.h.s.) the
magnitude of the magnetic eld B
0
stored in a hollow cylinder with cross section S was studied using
a resonance method. For freezing in the ux a homogeneous eld B
0
is applied along the axis of a
lead cylinder. Then the cylinder is cooled below T
c
. The trapped ux and, consequently, the oscillation
amplitude are found to increase in steps. The results demonstrate that the ux in the cylinder = B
0
S
can assume only multiples of h/2e = 2.067833758(46) 10
15
Wb.
Similarly important as the observation of the quantization in itself is the magnitude of the ux quantum.
In contrast to the expectation of London it is only h/2e rather than h/e. Onsager was probably the rst to
consider this possibility soon after the publication of the BCS theory in 1957. The meaning can hardly
be overestimated: the wave function which imposes the quantization condition is that of electron pairs,
exactly as derived by BCS, rather than single electrons.
2.1.5 Josephson effects
Another manifestation of the quantum nature of the superconducting state is the observation of coherence
phenomena between two weakly connected superconductors. The setting is similar to that of a tunneling
contact and the basic condition is that the wave functions of the superconducting condensates rather
than those of the single particles overlap. Then one can observe all phenomena characteristic of two
coupled quantum mechanical oscillators as will be derived in detail in section ??. The main results are
c _ Walther-Meiner-Institut
Thermodynamics SUPERCONDUCTIVITY 15
two equations which describe the statics (dc) and dynamics of carriers across a weak link [?],
I() = I
c
sin(
1

2
) (2.1.7)

t
=
2

0
U (2.1.8)
where
1,2
, , and
0
are, respectively, the phases in the weakly coupled superconductors 1 and 2,
the phase difference and the ux quantum. U is the voltage across the junction. In the case of the dc
Josephson effect there is a dissipationless supercurrent without a voltage drop across, e.g., an insulator.
The second equation describes the dynamics if the current exceeds the critical current I
c
which is the
only material dependent quantity. The Josephson frequency 1/
0
= 2e/h contains only constants and
has the value 483.597 870(11) MHz/V.
2.1.6 Summary
Superconductivity was discovered in 1911 in Hg three years after the rst liquefaction of He. Until
now the resistivity in the superconducting state is zero to within the experimental precision of approx-
imately 10
16
m. The ux quantization and the Josephson effects demonstrate the quantum nature of
superconductivity. In addition, the magnitude of the ux quantum shows directly that the wave function
of the condensate must correspond to pairs of electrons. The perfect diamagnetism as discovered by
Meissner and Ochsenfeld in 1933 demonstrates that superconductivity is a thermodynamic phase allow-
ing us to derive relations for the condensation energy, the entropy, and the specic heat capacity in the
superconducting state.
2.2 Thermodynamics
The thermodynamic properties provide us with a great deal of important information on a material in
particular on the energetics. Clearly, we learn something about the bulk while many spectroscopies
(except for neutrons) suffer from surface sensitivities. However, it is not always trivial to isolate the
desired quantity out of a large variety of contributions. In addition, in the presence of magnetism the
thermodynamic potentials are not uniquely dened (see Appendix 3) and the energy of the eld B itself
needs to be included in a way appropriate for the experimental circumstances. For a superconductor in
the Meissner state, for instance, the eld energy stored in the sample volume equals the condensation
energy. For simplicity we use the magnitudes of B, H, and M rather than the vectorial quantities. This
corresponds to a specialization to cylindrical symmetry and homogeneous media.
2.2.1 Condensation Energy
Since superconductivity is a thermodynamic phase the energy difference between the normal (n) and the
superconducting (s) state must be nite. Using the densities of the free energies f
n,0
and f
s,0
, respectively,
the energy difference for zero eld (index 0) can be expressed as
f
n,0
f
s,0
=
B
2
s
2
0
. (2.2.9)
Here, the condensation energy is expressed in terms of the so far unknown eld B
s
. As long as the
applied eld is smaller than the critical eld B
c
and completely excluded from the sample volume the
condensation energy does not change, hence f
s,0
= f
s,B
. Now we assume that a cylindrical sample of
2013
16 R. HACKL AND D. EINZEL Basic Experiments and Understanding
a superconducting material is put in the homogeneous eld B
0
of a long solenoid. In this idealized
conguration there is no distortion of the eld around the specimen since the demagnetization vanishes
(n = 0). The total volume-integrated free energies F for the interior of the solenoid with the sample in
the normal and the superconducting states for a eld B
0
< B
c
read, respectively,
F
n,B
0
= V f
n,0
+V
B
2
0
2
0
+V
ext
B
2
0
2
0
(2.2.10)
F
s,B
0
= V f
s,0
+V
ext
B
2
0
2
0
, (2.2.11)
where V = A L is the volume of the sample having cross section A and length L and V
ext
is the volume
inside the coil surrounding the sample, V
ext
=V
coil
V. For T < T
c
(B
0
) (and B
0
< B
c
) the difference of
the free energies becomes
F
n
F
s
=V( f
n,0
f
s,0
) +V
B
2
0
2
0
. (2.2.12)
If we increase B
0
towards B
c
from below and use Eq. (2.2.9) we obtain
F
n
F
s
=V
B
2
s
2
0
+V
B
2
c
2
0
. (2.2.13)
Upon crossing the critical eld superconductivity collapses and the eld B
0
= B
c
enters V. Both the
condensation and the eld energy induce an electromotive force E in the solenoid while the generator
maintains the constant current I
c
required for the eld B
c
. If there are N windings of the solenoid over the
length L the eld and the current are related as
0
NI
c
= B
c
L and E = N

with

the time derivative
of the ux through one winding. The total electrical energy W generated at the transition is given by the
time integral over the power,
W =
_
b
a
N

I
c
dt
=
_
a
b
(NI
c
)d
=
B
c
L A

0
_
B
c
0
dB
= V
B
2
c

0
. (2.2.14)
For the ux only the sample cross section is relevant since the eld around the sample does not change.
Since there is no dissipation (S = 0) W is equal to the free energy difference at B
c
given in Eq. (2.2.13)
and consequently B
s
B
c
. Hence the condensation energy is given by the critical eld, and the free
energy changes abruptly by twice the condensation energy upon crossing the B
c
line.
While the (Helmholtz) free energy is useful for studying the electromagnetic energy released at the
transition it is less useful for practical purposes for depending on the volume V and, practically always,
on the magnetization M. In an experiment it is next to impossible to control these variables. Therefore,
a potential which depends on the pressure p and the applied magnetic eld B
0
=
0
H
0
is desirable. The
Gibbs potential (Gibbs free energy) G=UTSpV BH
0
having the differential dG=SdT Vdp
VMdB
0
(upon using the proper U) depends only on controllable parameters and, in addition, is the most
appropriate function for studying phase transition. Since G is the macroscopic version of the chemical
c _ Walther-Meiner-Institut
Thermodynamics SUPERCONDUCTIVITY 17
potential , G = N, G
1
(T, B
c
) and G
2
(T, B
c
) must be equal at the transition between phase 1 and 2.
Therefore, we replace now F by G. For the densities we can write g = f BH
0
and get
G
n,B
0
= V f
n,0
V
B
2
0
2
0
V
ext
B
2
0
2
0
(2.2.15)
G
s,B
0
= V f
s,0
V
ext
B
2
0
2
0
. (2.2.16)
The second term on the r.h.s. of Eq. (2.2.15) is given only by the external eld since M is usually
vanishingly small in a superconducting material, and B = B
0
. The corresponding term in Eq. (2.2.16) is
zero since M =H
0
hence B = 0 in the Meissner state. The difference
G
n
G
s
=V( f
n
f
s
) V
B
2
0
2
0
(2.2.17)
obviously vanishes at B
0
= B
c
. For zero eld the difference of both potentials yields the condensation
energy,
G
n
G
s
=V
B
2
c
2
0
=F
n
F
s
. (2.2.18)
Obviously, the eld energy is irrelevant in this case and F can be replaced by F, which does not include
the vacuum eld. F is preferable for theoretical considerations for containing only the magnetization as
a microscopic quantity. Experimentally, F corresponds to the situation when a superconductor is rst
cooled below T
c,B
and then moved into the eld. Then, if we recall the setting from above the generator
has to supply only the energy of the excluded eld equalling the condensation energy. What happens if
the eld is now cranked up with the sample inside until it reaches B
c
(T)? One may have guessed, it is
exactly the case described above and twice the condensation energy is released. This is no violation of
energy conservation. Rather, the second half of the energy corresponds to the mechanical energy needed
to move the sample into the led from the eld-free region. Using m =VM, the differentials of F and G
read,
dF(T,V, M) = SdT pdV +Bdm (2.2.19)
dG(T, p, B
0
) = SdT +VdpmdB
0
, (2.2.20)
where the index 0 at B
0
is added to unambiguously denote that the eld is controlled from outside. This
is only possible in the case of G. In F, B depends on all other variables and on the location. However,
for a homogeneous material in the Meissner state B = B
0
is a good approximation. For constant T and p
dG
s
can be integrated,
_
B
c
0
dG
s
=
_
B
c
0
mdB
0
(2.2.21)
yielding
G
s
(T, B
c
) G
s
(T, 0) =V
B
2
c
2
0
. (2.2.22)
In the normal state the magnetization is small and
G
n
(T, B
c
) G
n
(T, 0) 0. (2.2.23)
2013
18 R. HACKL AND D. EINZEL Basic Experiments and Understanding
At the the phase boundary G
s
(T, B
c
) = G
n
(T, B
c
), and the difference of Eqs. (2.2.23) and (2.2.22) yields
again the l.h.s. of Eq. (2.2.18). Now, without including of the eld energy, F G+VB
0
M yields F = G
at B
0
= 0, and also the r.h.s. of Eq. (2.2.18) is recovered. At the critical eld one nds
F
s
(T, B
c
) = G
s
(T, B
c
) +VB
c
M
c
= G
s
(T, B
c
) VB
c
H
c
(2.2.24)
F
n
(T, B
c
) = G
n
(T, B
c
). (2.2.25)
Since the Gibbs potentials at B
c
are equal the discontinuity of F is recovered. Finally, we calculate the
variation of G
s
(T, B < B
c
),
G
s
(T, B) G
s
(T, 0) =
V
2
0
B
2
, (2.2.26)
G
s
(T, B) G
n
(T, 0) +
V
2
0
B
2
c
=
V
2
0
B
2
, (2.2.27)
G
s
(T, B) = G
n
(T, 0)
V
2
0
_
B
2
c
B
2
_
(2.2.28)
which is sometimes called the Meissner parabola and shown in Fig. 2.4 (a).
Before deriving other quantities we note that
B
2
c
2
0
is an energy density having the unit of a pressure. This
allows us to get a feeling for the order of magnitude of the condensation energy. For Nb with T
c
= 9.2 K,
B
c
= 0.2 T, and a lattice constant a = 3.3

A we get
B
2
c
2
0
= 16.5
kJ
m
3
(2.2.29)
= 16.5kPa (2.2.30)
corresponding to the pressure at 1.6 m under a water surface and a condensation energy of 2 eV/atom.
2.2.2 Entropy
The entropy measures the degree of order in a system and can be derived from both thermodynamic
potentials (see Eqs. (2.2.19) and (2.2.20)). From an experimental point of view G is more convenient.
For the consideration below, we can ignore the difference and write down the entropy S for either constant
magnetization and volume or constant applied eld and pressure,
S(T) =
F
T

M,V
=
G
T

B
0
,p
. (2.2.31)
In many cases one is interested in the temperature dependence of the entropy and other thermodynamic
quantities. For the entropy change upon entering the superconducting state we obtain for F
S(T) = S
s
S
n
=

T
(F
s
F
n
)
=

T
_
V
2
0
B
2
c
(T)
_
=
V

0
B
c
(T)
B
c
(T)
T
. (2.2.32)
c _ Walther-Meiner-Institut
Thermodynamics SUPERCONDUCTIVITY 19
Strictly speaking we can use only F here since the volume appears in the condensation energy. However,
in a solid the differences between quantities measured for constant volume or constant pressure are on
the order of a few percent and will not be considered below. Apart from this subtlety Eq. (2.2.32) is
extremely useful as soon as the temperature dependence of the critical eld (or of F) is known. In fact
the microscopic theory provides predictions which could be use. However, for getting qualitative insight
we use the phenomenological temperature dependence of B
c
given in Eq. (2.1.6).
From Eq. (2.1.6) we see and from experiments we know that
B
c
(T)
T
has no divergence at any temperature
and the following qualitative conclusions can be derived:
B
c
(T T
c
) = 0 implies that the entropy is continuous at T
c
. In other words, there is no latent heat
in zero eld.
In the limit T 0 the entropy difference vanishes according to the Nernst theorem. Since
B
c
(0) ,= 0 the derivative
B
c
(T0)
T
approaches zero. Obviously, Eq. (2.1.6) has the proper limiting
behavior for T 0. However, systematic studies show that the curvature is not correct.
In the range 0 <T <T
c
B
c
(T) decreases implying that
B
c
(T)
T
<0. As a consequence the entropy in
the superconducting state is smaller than in the normal state. This means that the superconducting
state has a higher degree of order and can transport heat less efciently.
For further considerations the derivatives of Eq. (2.1.6) are useful. We get
B
c
(T)
T
= B
c
(0)
2T
T
2
c
= B
/
[= B
//
T] (2.2.33)

2
B
c
(T)
T
2
= B
c
(0)
2
T
2
c
= B
//
= const. (2.2.34)
allowing us the calculate the latent heat as a function of temperature.
The latent heat is dened as the entropy change times the temperature, L = Q = TS. Using
Eq. (2.2.33) the density of the latent heat reads
L
V
= 4
B
c
(0)B
c
(T)
2
0
_
T
T
c
_
2
(2.2.35)
Except for T =0 and T =T
c
there is a latent heat at the phase transition, and the transition becomes
rst order.
2.2.3 Specic heat capacity
The specic heat capacity c
V,p
at either constant volume or pressure is an extensively used quantity
which reects the bulk properties. In superconductors the low-temperature part is particularly important
for characterizing the energy gap. Close to T
c
the heat capacity has a discontinuity which reects details
of the transition and the coupling strength. There are various experimental complications such as defects
or the superposition of strong other contributions from, e.g., the lattice. At the moment we wish to focus
only on the concept.
2013
20 R. HACKL AND D. EINZEL Basic Experiments and Understanding
Usually the heat capacity is measured at constant pressure. However, c
V
is more desirable since all
energies depend on the Volume. Fortunately the difference [c
p
c
V
[ is small in a solid and can be
ignored here. The specic heat capacity is dened as
c
V,p
=
1
V
Q
T

V,p
_
J
m
3
K
_
(2.2.36)
where Q is the thermal energy supplied in the interval between T and T +T. The molar heat capacity
which is frequently used in earlier publications is then given by c
M
mol

with M
mol
the molar mass and
the mass density.
Using Q = TS and taking the limit we get
c
x
=
T
V
_
S
T
_
x,...
. (2.2.37)
It depends on the thermodynamic potential which of the variables x, . . . is taken constant (see above).
For M and V being constant we obtain
c
V
=
T
V
_

2
F
T
2
_
M,V
. (2.2.38)
In the following the reference to constant variables is dropped for simplicity. The difference of the
superconducting and the normal state heat capacities can be calculated right away from the entropy
difference,
c
s
c
n
=
T
V
(S
s
S
n
)
T
=
T
V

2
T
2
(F
s
F
n
). (2.2.39)
From the condensation energy the famous Rutgers equation is obtained,
c
s
c
n
=
T

0
_
_
B
c
(T)
T
_
2
+B
c
(T)
_

2
B
c
(T)
T
2
_
_
. (2.2.40)
This is a general result that can be used whenever B
c
(T) is known from a microscopic treatment. If the
BCS predictions are used the data of Al, a prototypical weak-coupling superconductor, can be described
well. Similarly well works the Eliashberg theory for strong coupling Pb.
For a qualitative visualization of the thermodynamic functions (Fig. 2.4) and for demonstrating a few
important limiting cases we use the parabolic approximation for B
c
(T) [Eq. (2.1.6)].
In the limit T T
c
B
c
(T) has a nite slope and the heat capacity in the superconducting state is
bigger than in the normal state,
c
s
= c
n
+
8
T
c
B
2
c
(0)
2
0
. (2.2.41)
The discontinuity can be calculated on the basis of the low-temperature limes as will be done
below. If we disregard the parabolic variation of B
c
(T) for a moment we realize that the
discontinuity depends on the rate of variation of the entropy below T
c
with respect to that above
or, sloppily speaking, on the sharpness of the kink in the entropy at T
c
.
c _ Walther-Meiner-Institut
Thermodynamics SUPERCONDUCTIVITY 21
Thermodynamic functions
(
(

|
|
.
|

\
|
=
2
1 ) 0 ( ) (
c
c c
T
T
B T B
0.0 0.2 0.4 0.6 0.8 1.0
-1.0
-0.8
-0.6
-0.4
-0.2
0.0
T
c,B
B
0.95
0.9
0.7
0.5
0.3
0.2
0.0


2

0
[
B
c
(
0
)
]
-
2
[
g
s
(
T
,
B
)
-
g
n
(
T
,
B
)
]
T/T
c,0
B/B
c
(0)
F(T,0.3B
c
)
(a)
0.0 0.2 0.4 0.6 0.8 1.0
-0.6
-0.4
-0.2
0.0
0.2
(b)
T
c,B


[
s
s
-
s
n
]
/

T
c
T/T
c
0.0 0.2 0.4 0.6 0.8 1.0
-0.5
0.0
0.5
1.0
1.5
2.0
(c)
T
c,B


(
c
s
-
c
n
)
/
c
n
T/T
c
Figure 2.4: Thermodynamic function calculated for the parabolic approximation of the critical eld
[Eq. (2.1.6)]. Shown are the respective differences of the superconducting and the normal state functions
for the electronic part of (a) the Gibbs free energy g
s
(T, B)g
n
(T, B), (b) the entropy s
s
(T, B)s
n
(T, B),
and (c) heat capacity c
s
(T, B) c
n
(T, B) normalized to the volume and the condensation energy and
related quantities. In (a) the Gibbs potential is plotted for various elds as indicated. For zero eld the
Gibbs potential and the free energy F coincide (red line) and the slopes close to T
c,0
vanish indicating
a second order transition. For nite eld the transition at T
c,B
is rst order. While [g
s
(0, B) g
n
(0, B)[
becomes smaller with increasing eld for all temperatures and vanishes at B = B
c
[F[ starts to increase
with eld as shown for B = 0.3B
c
(dash-dotted green line). The free energy according to Eq. (2.2.19)
stays constant at T =0 (not shown). The entropy and the heat capacity are shown for zero eld (full line)
and for B/B
c
= 0.5 (blue dashed line). For nite eld the entropy has a discontinuity corresponding to a
latent heat L [Eq. (2.2.35)] and a -like contribution to the heat capacity (thick blue vertical line).
Since the entropy difference vanishes at T =0 and T =T
c
and since the superconducting state has a
higher order than the normal state S is negative for 0 <T <T
c
and has a minimum. Consequently,
c = c
s
c
n
crosses zero or, in other words, c
s
and c
n
have an intersection point. For directly
comparing c
s
and c
n
we need an approximation for c
n
and take the Sommerfeld model which
describes quasi-free electrons in a solid (see, e.g., Ashcroft and Mermin). Here, c
n
= T, and the
constant of proportionality is given by the electronic density of states for both spin projections
at the Fermi energy N(E
F
),
c
n
= T =

2
3
k
2
B
N(E
F
)T. (2.2.42)
In a conventional metal with only the lattice and the electrons contributing to the specic heat,
c(T) = T +bT
3
, can be determined experimentally. If the total heat capacity c/T is plotted
as a function of T
2
the intersection point at T = 0 yields . The discontinuity at T
c
can now be
2013
22 R. HACKL AND D. EINZEL Basic Experiments and Understanding
expressed as
c(T
c
)
c
n
(T
c
)
=
c
s
c
n
c
n
=
1
T
c
8
T
c
B
2
c
(0)
2
0
. (2.2.43)
One could arrive at a complete phenomenology including a prediction for c(T
c
) if could be
derived in some way. There is in fact a possibility opening up by further analyzing c
s
is in the
low-temperature limit [10].
To this end we rewrite Eq. (2.2.40) using the results of Eqs. (2.2.33) and (2.2.34),
c
n
= T = c
s

0
_
_
B
//
T
_
2
+B
c
(0)
_
1
_
T
T
c
_
2
_
B
//
_
. (2.2.44)
Since B
//
is a constant the last term approaches zero linearly for T 0 while T(B
//
T)
2
vanishes
rapidly. Also c
s
varies also faster than T as will be shown immediately and is found experimentally
for basically all known superconductors
3
hence the limit T 0 yields
=
4
T
2
c
B
2
c
(0)
2
0
[ N(E
F
)]. (2.2.45)
With plugged back in Eq. (2.2.44) the superconducting heat capacity at low temperature can be
expressed without a free parameter,
c
s
(T 0) =
12
T
2
c
B
2
c
(0)
2
0
_
T
T
c
_
3
. (2.2.46)
Using quite simple operations we derived several predictions demonstrating the importance of both the
thermodynamic arguments and experiments. Eq. (2.2.45) demonstrates the relation between the density
of electronic states N(E
F
), T
c
, and B
c
(0) hence the condensation energy which is also found in the
microscopic theory. Even though the constant 4 is model dependent, i.e. depends here on the slope of
B
c
(T) near T
c
, Eq. (2.2.45) conveys the important message of an interrelation between the normal and
the superconducting state allowing one to check the plausibility and consistency of experimental results.
According to Eq. (2.2.46) c
s
(T 0) approaches zero as T
3
. Trivially, this justies our assumption
c
s
(T 0) T
1+
with > 0. Less trivially Eq. (2.2.46) predicts a power law for c
s
(T) close to T = 0.
For this reason all early data for the low-temperature specic of superconductors were compared with
T
3
. However the results for tin or vanadium [?] clearly show an exponential variation (see Fig.) which
was immediately recognized as a manifestation of an energy gap in the electronic excitation spectrum.
The discrepancy between the prediction of Eq. (2.2.46) and the observed variation or the BCS result
originates, as mentioned, in the small differences between the parabolic and the proper temperature
dependence of B
c
(T 0). From an experimental point of view affairs are even more complicated.
Exponential variations are in fact rare and are found only in a few weak coupling superconductors such
as Al, Sn or V and some other elements. More recently, exponential temperature dependences were
also used for explaining c
s
(T) in MgB
2
or some iron-based systems. In either case multi-band effects
make the analysis rather complicated and less stringent. In strong coupling conventional materials like
Nb or Pb one nds in fact T
3
while in the copper-oxygen superconductors c
s
T
2
is established (see
chapter 6). Different reasons are at the origin of these deviations: For Pb and Nb strong coupling effects
lead to electronic states inside the gap even at T = 0 making the exponential temperature dependence to
vanish. The T
3
power law can be derived in the framework of the Eliashberg theory [?, 3]. In contrast
3
For a fully gapped superconductor like Al the T dependence is exponential, turning into T
3
in the strong coupling limit. In
the case of a gap with line nodes c
s
T
2
. Only in the case of osculating (kissing) nodes c
s
may vary linearly with T. These
important new developments will be discussed in chapter 6.
c _ Walther-Meiner-Institut
Electrodynamics SUPERCONDUCTIVITY 23
to the conventional metals, the cuprates do not have a constant gap. Rather, the gap vanishes on lines
on the Fermi surface (line nodes). Close to these nodes there are always thermal excitations except for
T = 0. The density of those excitations and, hence, c
s
depends on the k-dependence of the gap close to
the nodes. For the cuprates the variation is linear, i.e. proportional to (
0
), where is the angle on
the Fermi surface away from the position of the node at
0
.
Finally, we plug of Eq. (2.2.45) in Eq. (2.2.43) and obtain a universal number,
c(T
c
)
c
n
(T
c
)
= 2 (2.2.47)
which is larger in comparison to the weak coupling BCS value of 1.43. Obviously, the approximation of
B
c
(T) used for all our considerations here mimics strong coupling effects in general. This was shown
directly in detailed studies of the critical eld [?] the main result of which is reproduced in Fig. In Al
one nds 1.42(2) for the discontinuity in excellent agreement with the weak coupling result while 2.6
and 2.1 are observed for Pb (T
c
= 7.2 K) and Nb (T
c
= 9.26 K), respectively. The BCS value is also
universal and is derived from the entropy of the quasiparticle excitations that reect the gap. Larger
values for the discontinuity are strong coupling effects. The coupling strength in this context refers to
the interaction between the electrons and bosonic excitations such as phonons (considered by BCS) or
charge and spin uctuations (believed to play a role in unconventional superconductors) for instance
mediating the electron pairing below T
c
. In the Landau-Fermi liquid picture the coupling is characterized
by a dimensionless parameter which measures the interaction related reduction of the Fermi velocity
v
F
and the effective electron mass m

at the Fermi level (for an elementary discussion see, e.g., Ashcroft


and Mermin). With increasing interaction v
F
is reduced as (1 +)
1
while m

is enhanced as (1 +)
with varying between 0 (weak coupling) and values of order unity (strong coupling). For Al one nds
0.1. In the strong coupling elements Pb and Nb is close to 2.
We see that the discussion of thermodynamics provides deep insight into the microscopic origin of su-
perconductivity without, however, explaining the origin of the ordered or condensed state. Nevertheless,
the experimental observation of the entropy reduction below T
c
(via the heat capacity around T
c
) and of
the existence of an energy gap in the electronic excitation spectrum (via the heat capacity close to zero
temperature) are hallmarks of research paving the way towards the microscopic explanation. In addition,
the power laws found for the heat capacity of unconventional superconductors highlight the relevance of
thermodynamic studies in contemporary work.
Here, we have to pour some water on the vine. As already mentioned the heat capacity does not reveal the
desired isolated properties of the subsystem of interest which are the electrons here. Except for very low
temperatures the contribution of the phonons dominates the heat capacity, and sophisticated methods
are necessary to isolate the electronic contribution (see Loram). In magnetic systems spin excitations
may contribute enormously and in materials with disorder, even though very little, the contribution of
two-level systems (Schottky anomaly) is very strong. For these reasons, various other thermodynamic
quantities such as thermal conductivity or expansion are derived in order to bypass the problems inherent
to the heat capacity. Concerning the copper and iron-base systems the study of the latter quantities
revealed qualitatively new insights (Zaini, Ando, Taillefer, Meingast, Hardy).
2.3 Electrodynamics
Navely, perfect conductivity occurs if the time between two collisions of an electron or the mean free
path diverge. It is a different story how this can be realized since even in an ideally clean material
perfect conductivity is a quantum mechanical effect and can be expected only at zero temperature. It is
instructive nevertheless to consider the transition to perfect conductivity before discussing the possible
2013
24 R. HACKL AND D. EINZEL Basic Experiments and Understanding
origin in a quantum mechanical context. In 1935, after the discovery of the Meissner-Ochsenfeld effect
Fritz and Heinz London considered electrodynamic consequences of a two uid model consisting of a
mixture of two currents moving with and without dissipation. They did not further specify the origin of
the superuid current and used the limit but were aware that quantum mechanics must be at work.
The Maxwell equations as well as some useful identities and the implications of gauge transformations
required here and in the following will be summarized in Appendix 4.
2.3.1 The Drude model in the limit
The rst description of metallic conduction is due to Paul Drude [11] proposed shortly after the discovery
of the electron by Thomson in 1897. In this model the conduction electron having charge q move freely
between the atoms of a metal with an average drift velocity v
D
. Between two collisions at times t and
t + the electron is accelerated by the electric eld E driving the current j which can be expressed as
j = nqv
D
=
nq
m
p, (2.3.48)
where n, m, and p are, respectively the electronic density, mass, and momentum. In an incremental time
interval dt the electrons momentum changes from p(t) to p(t +dt),
p(t +dt) =
_
p(t) + f (t)dt +O(dt
2
)

_
1
dt

_
, (2.3.49)
driven by the force f (t) = qE. Contributions of higher than linear order in dt will be neglected as
expressed by O(dt
2
). Since the scattering time is nite the probability of an electron to not contribute
to the average momentum change is dt/ reducing the total acceleration by (1dt/). Neglecting terms
of order dt
2
and higher Eq. 2.3.49 can be rewritten as
dp
dt
=
p(t)

+ f (t) (2.3.50)
which, after multiplication by nq/m and substitution of f yields
_
d
dt
+
1

_
j =
nq
2
m
E. (2.3.51)
This result of Drude shows directly that the current increases linearly with time in the presence of a eld
E if becomes large constituting a slightly embarrassing statement. Therefore, before studying the limit
, we solve the differential equation (2.3.51). First we dene the conductivity () as j =E and
then assume E to vary harmonically as E(t) = E
0
e
it
,
_
i +
1

_
E
0
e
it
=
nq
2
m
E
0
e
it
,
yielding directly the nal result for the conductivity,
() =
nq
2
m
1

1
i
, (2.3.52)
or, upon separating real and imaginary part,
() =
nq
2
m

1
+i

2
+
2
=
nq
2
m

1+i
1+()
2
. (2.3.53)
c _ Walther-Meiner-Institut
Electrodynamics SUPERCONDUCTIVITY 25
The real part with q replaced by the electronic charge e (e = 1.602 10
19
Cb) assumes the usual form
() =
ne
2
m

1+()
2
(2.3.54)
which is proportional to the lifetime . In the static limit the usual dc conductivity
0
=
ne
2
m
is recovered.
The frequency dependence is given by a Lorentzian having a width
1
. The integral over all frequencies
obeys the famous f -sum rule,
_

d() =
ne
2
m
(2.3.55)
which states here nothing else but charge conservation. Now we are prepared to derive the limit 0.
Eq. (2.3.52) shows directly that becomes purely imaginary,
lim

() = i
ne
2
m
, (2.3.56)
and Eq. (2.3.55) can be interpreted as the Lorentz representation of the Dirac with spectral weight
ne
2
m
and the width vanishing as
1
. Since the conductivity is an analytic causal function, meaning that the
system responds to a perturbation at time t
0
only at t t
0
, real and imaginary part of are related by
Kramers-Kr onig transformations, e.g. for the real part,
() =
1

d
()

, (2.3.57)
where means taking the principal value. Using the imaginary part of as derived in Eq. (2.3.56)
the function centered at = 0 for the real part is recovered after some calculation and the complex
conductivity in the collisionless limit becomes

s
() =
ne
2
m
_
() +i
_
1

__
. (2.3.58)
The preceding qualitative study of the Drude conductivity in the limit of diverging scattering time in the
context of superconductivity is meaningful only for / h where is the energy gap of the individual
material. is, as we remember, the energy reduction of the electronic system that drives the phase
transition and will be derived microscopically in the next chapter. We will see then that the innite
conductivity at = 0 is a central result for superconductors. However, at nite frequency vanishes
only for h / h and approaches the normal state conductivity asymptotically for >/ h.
The functional form of the imaginary part turns out to be very useful (once again for h < / h) if the
complex conductivity can be determined experimentally. This is either possible by reectivity measure-
ments over wide energy ranges and exploitation of the analyticity of the response, here the reectivity
(Tanner, Basov), or by measuring reection and transmission in thin samples or by ellipsometry experi-
ments which directly return the real and imaginary part of the dielectric function =
1
+i
2
and, after
some simple algebra, the conductivity. Once is derived and is divided by a constant is obtained
which directly reects the plasma frequency

2
pl
=
ne
2

0
m
(2.3.59)
and, as we will derive below, the London penetration depth
L
for a magnetic eld.
2013
26 R. HACKL AND D. EINZEL Basic Experiments and Understanding
The reasoning can be carried on a bit further by taking the curl of Amp` eres law [see Maxwells equations
(problem 3 set 2 and problem 2 set 3)], and one obtains
_

1
c
2

2
t
2
_
B =
0
j (2.3.60)
=
0
()E
=
0
()

t
B. (2.3.61)
Note that we must use vectors here since derivatives with respect to r will be essential. If we solve this
differential equation by using the harmonic ansatz B(r, t) = B
0
(r)e
it
as above for E we obtain

2
B
0
(r) =

2
pl

2
c
2
i
1i
B
0
(r). (2.3.62)
Here the focus is placed on small frequencies <
pl
since superconductivity is destroyed other-
wise and is of order meV while
pl
is in the eV range. Therefore can be neglected, and the screening
equation is directly obtained

2
B
0
(r) =
B
0
(r)
[()]
2
, (2.3.63)
where () is the frequency dependent skin depth of a normal metal,
() =
c

pl
_
1+
i

. (2.3.64)
Formally, the London penetration depth can be obtained in the limit . But wait a minute! This is
only possible as long as the frequency is nite. If we take the hydrodynamic or static limit 0 which
we are interested in for a superconductor in a constant magnetic eld the penetration depth diverges. This
is exactly what happens in a normal metal where determines how fast the eld penetrates. One could
argue that the limit should be taken rst. Then, however, the result is valid for all frequencies
including those well above . Bottom line: A superconductor is not a perfect metal! While the results
for the conductivity are instructive and useful if the limitations are kept in mind they are not capable
to provide a genuine understanding of the superconducting state. To this end a quantum mechanical
treatment is necessary which yields the results also in the proper limits.
2.3.2 Generalized London theory
The quantum-mechanical nature of the condensate was anticipated by the London brothers already in
1935 and was discussed in more detail in 1950 when, for instance, the possibility of ux quantization
was mentioned. This implies that the condensate is described by a single wave function
(r, t) = a(r, t)e
i(r,t)
(2.3.65)
having a real amplitude a(r, t) and a single rigid phase (r, t)
4
rather than O(10
21
) electronic wave
functions in a macroscopic solid. (r, t) will be inserted in the Schr odinger equation allowing us to
derive an expression for a quantum mechanical current.
4
Rigid does not mean constant. In the presence of elds and currents depends on position and time as does the amplitude.
c _ Walther-Meiner-Institut
Electrodynamics SUPERCONDUCTIVITY 27
Lorentz force and canonical momentum
With a magnetic eld B present the kinematic momentum mv has to be replaced by the canonical mo-
mentum p = mv+qA where q is the charge of a point-like particle (q =e, with e = 1.602 10
19
Cb)
and A is the vector potential. The Lorentz force F
q
is directly related to this substitution as can be seen
by replacing E and B by A,
F
q
= m
v
t
= q(E+vB)
= q(
A
t
+(v A) (v )A). (2.3.66)
where all quantities depend on r and t. Using the total derivative of A
dA
dt
=
A
t
+(v )A
and reordering terms one nds a relation between the time derivative of the canonical momentum and a
new scalar potential

=v A at the position of the particle,


d
dt
(mv+qA) =q(v A). (2.3.67)
Thus the Lorentz force is the force on a particle in the co-moving coordinate system and can be derived
from a generalization of Newtons law [Eq. (2.3.67)] with the force q

.
Quantum mechanical derivation of the supercurrent
In quantum mechanics we use the correspondence principle and replace the momentum by the momen-
tum operator,
p
h
i
canonical momentum
mv
h
i
qA. kinematic momentum
Note that the kinematic momentum is not gauge invariant any further. For writing down the Schr odinger
equation we substitute q and m for the charge and the mass to indicate that the formulation (here and
later in the Ginzburg-Landau theory, chapter 4) is more general and applicable also for electron pairs, for
instance,
i h =
1
2m
_
h
i
qA
_
2
+q +. (2.3.68)
The Schr odinger equation is, of course, gauge invariant since a gauge transformation has to be applied
all quantities including and . Eq. (2.3.68) describes a particle having charge q. In a neutral system
(q = 0) only the potential =(r, t) survives which can be the chemical potential or a general potential
energy. For q = 0 and = 0 Eq. (2.3.68) describes a free uncharged particle. However, it is useful to
keep track of possible interactions. We insert now the single-particle wave function Eq. (2.3.65) into
Eq. (2.3.68) (Madelung transformation) and dene the action S(r, t) = h(r, t). Using S and taking the
real part of Eq. (2.3.68) yields (see Problem 3 in set 4)
S
t
+
(SqA)
2
2m
+q+ =
h
2

2
a
2ma
. (2.3.69)
2013
28 R. HACKL AND D. EINZEL Basic Experiments and Understanding
The second term on the l.h.s. is the kinetic energy
1
2
mv
2
and the sum of the second, third and forth terms
is the Hamilton function. The term on the r.h.s. of Eq.(2.3.69) indicates the quantum mechanical origin
of the equation and is proportional to the spatial variation of the amplitude of . The quasi-classical
limit can be obtained by making the transition h
2
0 directly leading to the classical Hamilton-Jacobi
equation,
S
t
=H. (2.3.70)
The imaginary part of Eq. (2.3.68) yields
a
2
t
=
_
a
2
m
(SqA)
_
(2.3.71)
which is a continuity equation for the probability-current density
j
p
=
a
2
m
( h qA) (2.3.72)
since a
2
=

is a probability density the multiplication of Eq. (2.3.72) with q or m yields the respective
charge and mass densities a
2
q and a
2
m and current densities j
p
q and j
p
m. The result is, of course,
equivalent to the usual quantum mechanical current density, which was already derived by Schr odinger
and follows also from the Ginzburg-Landau equations,
j
p
=
a
2
m
_

_
h
i
qA
_
+
_

h
i
qA
_

_
(2.3.73)
and can be obtained by using Eq. (2.3.65) for . The supercurrent in a charged system can be written as
j
s
=
a
2
q
m
( h qA)
n
s
q
2
m
(
h
q
A) (2.3.74)
where n
s
= a
2
is the density of superconducting carriers having mass m and charge q. This equation
can be considered the central result of the London theory. All London equations can be derived directly
from this quantum mechanical expression.
2.3.3 The London equations
As shown experimentally (see paragraph 2.1.4) and derived theoretically (see chapter 3) the charge in
a superconducting condensate is 2e (for traditional reasons the positive sign is used) implying m 2m
and n
s
n/2 for the mass and the density, respectively. These values will be used from now on.
For practical purposes it is more convenient to have a direct relationship between the current and the
magnetic eld. By applying the curl to Eq. (2.3.74) and remembering that = 0 the second
London equation
j(r) =
ne
2
m
B
0
(r). (2.3.75)
follows immediately. It describes the existence of a current in the presence of a static eld B
0
and
explains the Meissner-Ochsenfeld effect phenomenologically. In the static limit and for an isolated piece
of a superconductor the charge density does not change and the continuity equation (2.3.71) yields j =
c _ Walther-Meiner-Institut
Electrodynamics SUPERCONDUCTIVITY 29
= 0. Then the screening equations for the eld B and the current density j can be derived directly
from Eq. (2.3.75) by either substituting j =
1
0
B or taking the curl, respectively,

2
B
0
(r) =
B
0
(r)

2
L
, (2.3.76)
where we substituted m/
0
ne
2
by
2
L
, the square of a length, and

2
j(r) =
1

2
L
j(r). (2.3.77)
The expressions demonstrate that the eld and the supracurrent exist only in a surface sheath of thickness

L
as will be discussed in detail in the following paragraph. Upon substituting B
0
=A in Eq. (2.3.75)
the resulting equation
j(r) =
ne
2
m
A(r) (2.3.78)
is the not gauge invariant version of the second London equation. A closer look reveals that Eqs. (2.3.74)
and (2.3.78) are equivalent modulo a gauge transformation. Using Eq. (2.3.78) one can derive a screening
equation also for A. Finally, by taking the time derivative of Eq. (2.3.78), one gets the acceleration
equation

t
j(r) =
ne
2
m
E(r), (2.3.79)
which was discussed earlier and is sometimes called the rst London equation. Even if we nally arrive
at equations, which are formally equivalent to those derived from the Drude model in the limit
the argumentation here is qualitatively different. Most importantly, the quantum mechanical nature is
properly taken into account allowing us to avoid articial and unphysical arguments such as an innite
lifetime at nite temperature.
5
In addition, the result is valid in the right limit of 0 which
particularly includes the static limit. The upper limit will be derived in chapter 3.
2.3.4 Some conclusions
We discuss rst consequences of the screening and then demonstrate the the main conclusions from the
quantum mechanical nature of the Eq. (2.3.74) before returning to a special case of screening.
The penetration of eld and current into a semi-innite superconductor
Eq. (2.3.76) describes the screening of a magnetic eld by a superconductor. Here we present a solution
for the simplest possible geometry being a semi-innite solid with the yz plane at x = 0 separating the
superconductor (x 0) from the vacuum. The homogeneous eld B
0
= B
z,0
is aligned along the z-axis
making the problem one-dimensional with B
z
(x) only varying along x,
d
2
B
z
(x)
dx
2
=
B
z
(x)

2
L
. (2.3.80)
5
This is a point of view which became clear only after the formulation of the microscopic theory and is not intended to
criticize earlier proposals. However, it is the only sensible approach in these days.
2013
30 R. HACKL AND D. EINZEL Basic Experiments and Understanding
-2 -1 0 1 2
-2
-1
0
1


A
y
(
x
)
/

L
B
z
(
0
)
x/
L
A
y
(x<0)
A
y
(x>0)
B
z,0
B
z,0
B
z
(x)
Figure 2.5: Theoretical prediction for the penetration of the induction B = B
z
(x), the current density
j
s
= j
(s)
y
(x), and the vector potential A = A
y
(x).
The usual ansatz for the differential equation (2.3.80) is an exponential, B
z
(x) = be
x
, yielding the char-
acteristic equation =
1
L
. The boundary conditions B
z
(0) = B
z,0
and B
z
(x ) = 0 determine the
solution completely,
B
z
(x) = B
z,0
e

L
. (2.3.81)
The current density follows from Amp` eres law,
0
j
s
= B e
z
B
z
(x), and is oriented along the
y-axis represented by the unit vector e
y
. The vectorial relation reads
j
s
(x) = e
y
B
z,0

L
e

L
. (2.3.82)
Eq. (2.3.78) gives one possible choice for the vector potential being anti-parallel to the current density,
A(x) =
m
ne
2
j
s
= e
y
m

0
ne
2
B
z,0

L
e

L
. (2.3.83)
B =A leads back to Eq. (2.3.81). The relationship between the various quantities is summarized in
Fig. 2.5. The characteristic length
L
is called the London penetration depth and is given by

L
(0) =
_
m

0
ne
2
= c
_

0
m
ne
2
=
c

pl
. (2.3.84)
The rst equality in Eq. (2.3.76) is found frequently, probably because it follows directly from the screen-
ing equation (2.3.76), but obscures the meaning a little. In fact, Eq. (2.3.84) is a remarkable result since it
relates a genuinely superconducting length scale with the velocity of light and a plasma frequency which
is orders of magnitude above the energies relevant for the condensate. In addition,
pl
is composed only
of quantities characterizing the normal state. Note, however, that one can substitute the values of q, m,
and n
s
without changing
pl
. Typical plasma frequencies for metals are in the visible range and above
(1-5 eV) corresponding to approximately 2(0.2. . . 1)10
15
s
1
and yield a range from 40 to 200 nm for

L
in overall agreement with experiment (see Table 2.1).
Table 2.1: Experimental magnetic penetration depths in nm of selected superconductors.
Al Pb Nb
3
Sn YBa
2
Cu
3
O
7
3015 40 80 9010
c _ Walther-Meiner-Institut
Electrodynamics SUPERCONDUCTIVITY 31

pl
results from the broken gauge symmetry or the unique phase and phase stiffness of the condensate
and is closely related to the Higgs mechanism which makes the photons to acquire a nite mass in the
presence of a gauge eld being equivalent to the vector potential Ahere. [?, 12] Hence, the Meissner
effect is the Higgs mechanism in a superconductor and
pl
reects the rigidity of the phase rather than
the relevant energies. At T
c

L
(T) diverges, the eld penetrates, and the photons lose the mass they
acquired due to xed phase of the condensate.

1
L
is proportional to

n
s
which depends on temperature and vanishes at T
c
. If there is a gap in the
electronic excitation spectrum n
s
saturates exponentially at low temperature while if the gap becomes
zero for specic k-points n
s
can be described by characteristic power laws. For this reason precise
measurements of
L
(T) play an important role for characterizing superconductors.
There are various methods which are particularly sensitive to changes rather than absolute magnitudes.
Prominent examples are microwave, spin rotation, optical (see above) but also magnetometry tech-
niques. For example if a sample is put in a maximum of the B eld of a superconducting resonator the
quality factor Q
2
changes when the penetration depth of the sample changes as a function of tempera-
ture. However, Q
2
can be determined with very high precision using state of the art equipment. The
results of this technique on very pure YBa
2
Cu
3
O
7
samples (Hardy) nally convinced the majority of the
people that superconductivity in the cuprates is unconventional. In a similar fashion by measuring the
changes of the penetration depth as a function of temperature in UPt
3
and UBe
13
with a magnetometer
the proposal of unconventional pairing was put forward for the rst time [13].
Note that all previous equations here are strictly local in that the current in rdepends only on the eld in
r. Upon studying the penetration depth of PbIn alloys having various concentrations of defects and mean
free paths Pippard found the penetration depth to depend systematically on . He concluded that there
must be another length scale characterizing the superconducting state and called it coherence length
0
.
Whenever and
0
are of the same order of magnitude the modication of
L
turned out to be particularly
strong. in fact,
0
measures the volume around rover which one has to integrate to properly calculate the
response to a perturbation in a point r. In other words, the response in a superconductor is not local but
depends on the neighborhood of a point. This purely experimental observation, to which we return when
discussing the London vortex, was reproduced in an excellent fashion by the BCS theory [2].
Vanishing canonical momentum
If the velocity v in the canonical momentum Eq. (2.3.67) is expressed in terms of the supracurrent j
s
=
nev
s
one obtains
p =
_
m
ne
j
s
+eA
_
(2.3.85)
and, upon using Eq. (2.3.78),
p = 0. (2.3.86)
This is another remarkable result that predicts the existence of superconducting charge carriers with
vanishing total momentum and consequently anticipates Cooper pairing of electrons at momenta k and
k.
A simply connected superconductor
For an isolated bulk sample of a superconductor the charge density
s
is constant. This corresponds to
the absence of currents j
s
owing in or out,
0 =

s
t
= j
s
=
e
2
n
s
m
_
h
2e
A
_
. (2.3.87)
2013
32 R. HACKL AND D. EINZEL Basic Experiments and Understanding
Fluxoid Quantisierung
{ } A J Q
M
QN
s
s
Q
V =
S
cS
s
Q L
Q
J A
2
0
+ = V

... , 2 , 1 , 0 ; 2 ) , ( = t = V
}
c
n n t d
S
r r
B
0
{ }
{ }
Q
h
n
n
Q
d
d
S
s
Q L
S
s
Q L
= u
u =
t =
V + =
+ u
}
}
c
0
0
2
0
2
0
'
2

J B S
J A r
Flux(oid) Quant
Figure 2.6: Determination of the ux through a hollow cylinder. The homogeneous applied eld B
0
is parallel to the cylinder axis and perpendicular to the surface S with boundary S being inside the
superconductor.
Except for the term Eqs. (2.3.78) and (2.3.87) look similar. The question is therefore how they can
be reconciled. The problem was already realized by BCS, and they argue that vanishes in the right
gauge. It is elucidating to argue the other way around and use the London gauge

A = 0. Then
2

follows immediately and =const is a direct consequence. Since the starting point was an isolated
superconductor with no currents crossing the surface has to vanish on the surface since the current is
proportional to and [
surface
= 0. With =const 0 follows immediately demonstrating the
rigidity of the phase independent of statistical arguments.
It is interesting to study Eq. (2.3.87) without using the London gauge. To keep things simple we assume
a harmonic spatial variation of all elds having the form f (r) = f
0
e
ikr
yielding the replacement ik
and arriving at
j
s
=
q
2
n
s
m
(

A) =
q
2
n
s
m
(ik

A).
Taking the divergence in the static limit as above one nds ik A+k
2

= 0 allowing one to eliminate

. The resulting current


j
s
=
q
2
n
s
m
_
k(k A)
k
2
A
_
(2.3.88)
is now gauge invariant and equivalent to the second London equation [Eq. (2.3.75)]. k is the wave vector
characterizing position dependent elds on length scales of
L
. k A is the longitudinal projection of the
eld A on the direction of the fastest change of

, i.e.

, and vanishes in the London gauge.


Quantization of the uxoid
The quantum mechanical form of the supercurrent [Eq. (2.3.74)] provides us with the proper tool to
explain the ux quantization experiment described in paragraph 2.1.4 and to get additional insight into
the necessary experimental conditions. The setting, this time from a theoretical point of view, is sketched
in Fig. 2.6. We rearrange [Eq. (2.3.74)] and integrate over the entire boundary S of the surface S (see
Fig. 2.6) on both sides,
h
q
_
S
d =
_
S
_
A+
m
n
s
q
2
j
s
_
d. (2.3.89)
c _ Walther-Meiner-Institut
Electrodynamics SUPERCONDUCTIVITY 33
The line integral over a gradient of a eld is given by the difference of the elds in the starting and
end points. Since is the phase of the wave function Eq. (2.3.65), which has to be dened uniquely in a
position r, is dened only up to integer multiples of 2,
(r) = a(r)e
i(r)
= a(r)e
i((r)+2n)
= (r),
and the integral over the phase gradient yields
_
S
d = 2n. n Z (2.3.90)
Using this result and Stokes theorem Eq. (2.3.89) can be recast,
h
q
n =
_
S
(A) ndS+
0
_
S

2
L
j
s
d, (2.3.91)
where n is a unit vector perpendicular to S. The expression on the r.h.s. is the uxoid since it depends
not only on the ux through the cross section of the hollow cylinder but also on the screening currents
in the cylinder walls. The quantization of the uxoid was demonstrated by Little and Parks [?] in 1963.
Only if the walls are much thicker than
L
one can nd an integration path which is sufciently buried
inside the wall so as to j
s
0 along S and to make the second term on the r.h.s. of Eq. (2.3.91) vanish.
Only in the case j
s
= 0 the ux is quantized,
=
_
S
B ndS =
h
q
n. (2.3.92)
The London vortex
Finally we return to the screening equation in a mathematically more complicated geometry and calculate
the energy in a single vortex line for getting a rst idea of how the mixed state (see section 2.1.3)
develops. First we determine the asymptotic behavior of the eld around the line as a function of the
distance r. The second step is a lengthy calculation (see Fossheim and Sudb), and we write down only
the result. A single vortex can penetrate a simply connected superconducting material for instance in a
type II superconductor very close to B
c1
. Then there is a normal lament with radius
L
(vortex
core) oriented along the eld, e.g. e
z
, where the eld is maximal and supercurrents circulating around the
lament which screen the eld over a length scale of
L
. By replacing
2
B we can write the screening
equation as
B+
2
L
(B) = e
z

L
(r) (2.3.93)
where (r) is the two dimensional Dirac function. The meaning of Eq. (2.3.93) becomes immediately
clear if we integrate over circular surface S perpendicular to e
z
having a radius r
L
,
_
S
BdS+
2
L
_
S
(B)d = e
z

L
. (2.3.94)
For the second term we used again Stokes theorem. With B =
0
j
s
and j
s
= 0 for r
L

L
is
the total ux along the line. If, on the other hand, < r <
L
only a fraction r
2
/
2
L
of the total ux
is screened, and we set
_
S
B ndS zero. The remaining part can be further simplied by exploiting the
cylindrical symmetry,
[B[ =
B
r
=

L
2r
2
L
, (2.3.95)
2013
34 R. HACKL AND D. EINZEL Basic Experiments and Understanding
where 2r results from the line integral, yielding
B(r) =

L
2
2
L
_
c ln
_
r

L
__
. (2.3.96)
Since this approximate expression diverges for r 0 has the role of a cut-off below which B(r)
becomes constant. The full functional form is the solution of a Bessel differential equation resulting
from (B) in cylindrical spherical coordinates and B = e
z
B
z
,
1
r

r
_
r
B
z
r
_

B
z

2
L
= 0 (2.3.97)
that leads to a 0
th
-order Bessel or Hankel function K
0
,
B(r) =

L
2
2
L
K
0
_
r

L
_
(2.3.98)
which, for r <
L
, has the asymptotic behavior described above and decays exponentially for r >
L
. If
one determines the energy of a vortex line by integrating over the eld and the current density
L =
1
2
0
_
B
2
+
2
L
(B)
2
dr (2.3.99)
one nds for the energy per unit length L
L =

2
L
4
2
L

0
_
ln
_

L
__
; 0.1 (2.3.100)
where the term originates in the condensation energy. For a qualitative argument (see Annett) one
can neglect the condensation energy and conne the integration to the region r
L
and substitute
B) by j =
L
/(2r
2
L
) e

)in Eq. (2.3.99). Since we are interested in the energy per unit length the
integration dr = d
2
r = 2rdr yields Eq. (2.3.99) with = 0.
The signicance of Eq. (2.3.100) is in the proportionality of the energy L to the square of the number of
ux quanta, n
2
. This means that the costs of getting more than one ux quantum into a vortex increase
more rapidly than those for creating n vortices, nL
1
< L
n
. Eq. (2.3.100) but also the derivation of the
asymptotic behavior highlight the existence of a second length scale over which superconductivity is
almost completely suppressed while the supracurrents and the eld can coexist between and several

L
. is essentially the same length scale as
0
derived by Pippard upon studying the penetration depth in
disordered superconductors (see above).
0
will turn out to be related to the length scale over which
coherence in the condensed state can be maintained and which reects the diameter of a Cooper pair.
Further details will be discussed in the frameworks of the BCS and the Ginzburg-Landau theory.
c _ Walther-Meiner-Institut
Chapter 3
Microscopic Theory
The microscopic theory was nally presented in 1957. The last break-through was the derivation of a
coherent wave function using a superposition of Fermions by Schrieffer. Cooper, on the other hand,
showed that an innitesimally small attractive interaction between Fermions having E > E
F
leads to
a reduction in their energy by an amount proportional to the cut-off energy of the interaction. The
possibility of an attraction between electrons mediated by phonons was already proposed by Fr ohlich
and studied in detail by Pines and Bardeen. Under the supervision of John Bardeen the three ingredients
were combined. The plan of the chapter is as follows: We rst prove the existence of the Cooper
instability. In the second paragraph we show that the electron-phonon interaction leads indeed to an
attractive interaction. The third paragraph is devoted to the derivation of the coherent wave function.
Finally, we show how the Hamiltonian of the interacting system can be minimized at zero and at nite
temperature. In the last part we calculate a few response functions in order to show how a superconductor
reacts to external perturbations.
3.1 The Cooper Instability
We assume that the Fermi sea is completely lled at zero temperature. We now add two particles with
momenta k
1
, k
2
>k
F
as shown in Fig. 3.1. If we construct a state of two Fermions the total wave function
has to obey anti-commutation relations so as to satisfy the Pauli principle,
(r
1
,
1
, r
2
,
2
) =(r
2
,
2
, r
1
,
1
). (3.1.1)
We can separate the wavefunction into three factors if we introduce relative coordinates, r
1
= R+r/2
and r
2
= Rr/2, with R and r describing the center of mass and the distance of the two electrons. With
spin part represented by (
1
,
2
) and the propagating part by plane waves reads
(r
1
,
1
, r
2
,
2
) =
1

V
e
iKR
e
ikr
(
1
,
2
), (3.1.2)
which can be reformulated as
=
1

V
e
iKR
(r
1
r
2
)(
1
,
2
). (3.1.3)
35
36 R. HACKL AND D. EINZEL Microscopic Theory
Cooper-Paare
+ , k
T =0
E
F
E
F
+e
D
| , k
( )
( ) Triplett
2
1
Singulett
2
1
) , , , ( -
) ( ) , , , (
spin
,
spin
,
1 1 2 2
spin
, 2 1 2 2 1 1
2 1
2 1
2 1

++
+| + |+
||
=
+| |+ =
=
=

o o
o o
o o
|
|
o o
| o o
r r
r r r r
R K i
e
Figure 3.1: Cooper instability of two particles having opposite momenta and spins in a shell with thick-
ness h
0
above the Fermi energy E
F
. At T = 0 the Fermi sea is completely lled hence the additional
particles have an energy above E
F
.
It was shown before that the London theory predicts vanishing canonical momentum of the supercon-
ducting carriers [Eqs. (2.3.78) and (2.3.85)] on purely electrodynamic reasons. This momentum can now
be identied as the center of mass of two electrons implying a state k
2
= k
1
. Consequently, the rst
term in Eq. 3.1.2 is essentially unity in equilibrium. The second term shows that the carriers can have
internal relative momentum. The third term describes the spins. In a system with inversion symme-
try where parity is a good good quantum number the two possible relative orientations of the spins are
parallel and anti-parallel, and one arrives at triplet and singlet wave functions

t
(
1
,
2
) =
_
_
_
[
1

2
([ +[ )
[
(3.1.4)

s
(
1
,
2
) =
1

2
([ [ ). (3.1.5)
If the two spins are exchanged
t
(
1
,
2
) and
s
(
1
,
2
) conserve or change sign, respectively. Since
the center-of-mass part is constant the product of the functions and must be anti-symmetric to
fulll Eq. (3.1.1), has to be symmetric and anti-symmetric for singlet and triplet spin wave functions,
respectively. Possible representations are
s
= cos[k (r
1
r
2
)] and
t
= sin[k (r
1
r
2
)]. Recently,
superconductivity was discovered in CePt
3
Si [?] and other systems without inversion symmetry. As a
consequence the distinction between singlet and triplet states is not possible any further and there are
always mixtures. In addition the Fermi surface splits up into two sheets. However, we wont dwell on
this exotic though exciting and well studied case [?] but use it just as an appetizer for the part on novel
superconductors.
Now, we derive the energy gain for the case of a small attractive interaction between the two extra
electrons outside the Fermi sea. If we leave out all complications this is a simple exercise in perturbation
theory. We assume that the unperturbed system has only kinetic energy

T and count the energy from
the chemical potential . From a statistical point of view we take the grand canonical potential with the
particle number being variable. Subtracting the energy of the Fermi sea means using the Landau-Fermi
quasiparticle concept and excitation energies
k
=
k
rather than band energies
k
. With

V being
an attractive interaction between two electron states k and k having magnitudes larger than the Fermi
momentum, [k[ > k
F
and energy
k
> 0, the eigen energies E of the perturbed Hamiltonian

H =

H
0
+

V
can be found from the eigen states of the unperturbed system,

H
0
[k = 2
k
[k (3.1.6)
c _ Walther-Meiner-Institut
The Cooper Instability SUPERCONDUCTIVITY 37
where the functions [k are a complete set of eigen functions obeying k
/
[k =
kk
/ . Then

H[ = E[ (3.1.7)
[ =

k

k
[k. (3.1.8)
Multiplication with k
/
[ from the left and collecting terms yields
k
/
[(

H
0
+

V)

k
[k = k
/
[E

k
[k

k
/ (2
k
/ E) =

k
k[

V[k. (3.1.9)
The calculation of the matrix element on the right hand side is usually complicated since the poten-
tial between two electrons is unknown and since there are only approximations to the wave functions.
Eliashberg was the rst to present a realistic model [3] in terms of a material-specic electron-phonon
interaction. Only recently, the wave functions could be approximated for a few conventional systems [?].
Bardeen, Cooper and Schrieffer, although having a phenomenology for the attractive interaction, as-
sumed simply that V
kk
/ k[

V[k is attractive for energies smaller than a typical phonon energy h


0
and
zero otherwise,
V
kk
/ =
_
g
2
eff
[
k
[ h
0 otherwise
Now Eq. (3.1.9) can be solved right away by noting that, since [k is a complete set,
k

k
=
k
/
/
k
,

k
/ = g
2
eff

k
2
k
/ E

k
/

k
/ = g
2
eff
k
/

k
2
k
/ E
1 = g
2
eff
k
/
1
2
k
/ E
. (3.1.10)
The k-summation in Eq. (3.1.10) can be transformed into an energy integral as
k
= N
F
_
h
0
0
d where
N
F
is the density of states for both spin projections yielding
E =2 h
0
exp
_

2
N
F
g
2
eff
_
(3.1.11)
for the case N
F
g
2
eff
1. N
F
g
2
eff
is a dimensionless parameter which is often abbreviated by the coupling
constant . In the BCS weak coupling approximation is much smaller than 1. In the strong coupling
case discussed by Eliashberg can be of order 1 or even larger. Here we focus on the weak-coupling
limit and note that even in this case the energy gain E cannot be expanded in powers of . Via the
cutoff energy h
0
, which will be discussed in more detail in the next section, the energy gain depends
on a material property. While there are various other possibilities, in the treatment of BCS h
0
was a
typical phonon energy directly explaining the experimentally observed isotope effect T
c
M

where
is expected to be 0.5 for a harmonic oscillator (Fig. 3.2).
2013
38 R. HACKL AND D. EINZEL Microscopic Theory
BCS-Theorie
Figure 3.2: Isotope effect in tin [10]. The different symbols refer to the results of different authors. The
table shows the exponent for various elements.
3.2 Origin of the interaction
Although suggested by the isotope effect (Fig. 3.2) an interaction depending on the lattice is not the most
obvious way towards an attraction between electrons. Rather, the electro-static force seems to prevail by
far. In order to understand the reasoning at the time of discovery and to get an idea of the relevant physics
we rst have a closer look at the Coulomb interaction in a metal which is distinctly different from that of
point charges given by
V
C
=
e
2
4
0
r
(3.2.12)
where r =[r
2
r
1
[ is the distance between two electrons at r
1
and r
2
. The Fourier transform is given by
V
C
(q) =
_
drV
C
e
iqr
=
e
2

0
q
2
. (3.2.13)
In a metal, in contrast, the Coulomb interaction is screened. For instance, a charged impurity is hardly
visible just a few lattice constants away. The exact distance depends on the charge density and is de-
scribed ba the Thomas-Fermi theory. The screened potential was proposed by Yukawa in the context of
nuclear matter, reads
V
C,TF
(r) =
e
2
4
0
r
e

r
r
TF
, (3.2.14)
and has the Fourier transform
V
C,TF
(q) =
e
2

0
(q
2
+k
2
TF
)
. (3.2.15)
If there is one electron per lattice site r
TF
a and k
TF


a
close to the Fermi momentum. The potential
does not decay strictly exponentially but oscillates around the zero (Friedel oscillations). In any case
c _ Walther-Meiner-Institut
Origin of the interaction SUPERCONDUCTIVITY 39
E
i
t
E
i
k
k
-k
-k
-k-q
-k-q
k+q
k+q
-q q
(a) (b)
Figure 3.3: Feynman diagrams for the interaction between two electrons via the emission and reabsorp-
tion of a bosonic particle, here a phonon.
the perturbation of the charge density is screened over distances of a few lattice constants, and other
interactions can gain inuence. In conventional metals the most important additional interaction is the
electron lattice interaction. Forces from charge and spin modulations are important in systems like
cuprates or iron-based superconductors but will not be studied here at the moment. In most of the cases
the energy scales of the electronic degrees of freedom are much higher than those of the secondary
interactions implying the existence of distinctly different time scales. In the case of the phonons this
difference is quite obvious since the masses of the involved particles are roughly 4 orders of magnitude
different. While the electrons follow a perturbation practically immediately the ions react slowly but
remember the perturbation for longer. This is called the retardation effect and can be understood as
follows: An electron with momentum k moves in the lattice and attracts the ions close to its trajectory.
This distortion of the lattice survives on time scales much longer than the lifetime of an electron between
collisions and provides an attractive potential for the other electrons especially (for phase space reasons)
for those having opposite momenta k. Pines and Bardeen have studied the resulting potentials between
the electrons in the framework of the jellium model (electrons on a homogeneous background of positive
charge) and found the result used for the BCS theory [?, ?]. The derivation can be found in the textbooks
[14] and will not be repeated here. The result for the full potential is expressed in terms of the dielectric
constant one obtains
V
kk
/ =
e
2

0
q
2
(q, )
, (3.2.16)
and if the Thomas-Fermi screening and the electron-ion interaction are included in the expression for
(q, ) the full result can be written down as
V
kk
/ =
e
2

0
(q
2
+k
2
TF
)
_
1+

2
q

2
q
_
. (3.2.17)
One can get an idea (not a rigorous derivation) of the quantum mechanical processes by considering
the energy change due to emission and reabsorption of phonons by electrons as suggested by de Gennes
and shown in Fig. 3.3. To this end we have to calculate in second order perturbation theory the matrix
element k[

V
epe
[k
/
from two transitions of the formk[

H
ep
[i =W

q
(i) and sum over all intermediate
2013
40 R. HACKL AND D. EINZEL Microscopic Theory
states i. Since there are two time-reversed processes we get
k[

V
epe
[k
/
=
1
2

i
k[

H
ep
[i
_
1
E
k
/ E
i
+
1
E
k
E
i
_
i[

H
ep
[k
/
. (3.2.18)
If we ignore the inuence of the small differences between the states k and k
/
on the electron phonon
matrix elements the second element on the r.h.s. is just W
q
(i). The initial and nal energies are E
k
= 2
k
and E
k
/ = 2
k
/ , respectively, since
k
is a symmetric function of k as is the phonon energy h
q
w.r.t. q.
In the intermediate state the crystal momentum and the energy are conserved, yielding equal energies for
diagram (a) diagram (b)
momentum energy momentum energy
electron 1 k
/
= k+q
k
/ k
k
electron 2 k
k
k
/
=(k+q)
k
/
phonon q h
q
+q h
q
both processes in the intermediate states, E
(a)
i
=E
(b)
i
=
k
+
k
/ + h
q
. After replacing (
k

k
/ ) by h
the energy denominators in Eq. (3.2.18) read h h
q
. Since there is no reference to the intermediate
state any further the summation is only over the matrix elements and the total expression for the energy
change due to electron-phonon coupling becomes
k[

V
epe
[k
/
=
[W
q
[
2
h
1
2
_
1

q

1
+
q
_
. (3.2.19)
For the full interaction including Coulomb part we use the Thomas-Fermi result [Eq. (3.2.15)] in addition
to Eq. (3.2.19) and get
k[

V[k
/
=V
C,TF
(q) +
[W
q
[
2
h

2
q
. (3.2.20)
Although this result is anything else but quantitative it provides a feeling for the underlying physics:
the energy dependence of the effective electron-electron interaction results from a phonon-mediated
perturbational correction to the electronic energies (Fig. 3.4). It is limited to a range of the order of the
Debye energy h
D
. At energies well above h
D
the usual screened but repulsive Coulomb interaction
prevails. For small energies the effect of the Coulomb interaction is overcompensated by the lattice,
sometimes called overscreening effect, and can lead to a appreciable attractive potential that provides the
interaction needed for Cooper pairing.
3.3 The BCS wave function
As outlined in chapter 1 [Eq. (1.2.1)] the construction of a coherent state of bosons is relatively straight-
forward. With bosons the macroscopic occupation of the ground state can be realized directly since
all particles can be in the same state of vanishing momentum at sufciently low temperatures. With
Fermions the Pauli principle precludes a direct solution. Before we present Schrieffers solution to the
problem we have a brief look at the coherent state proposed by Schr odinger [?] since it has already all
relevant properties needed later and is easier tractable. We only summarize the results and leave the
detailed calculations as an exercise (problem 1 of set 1).
c _ Walther-Meiner-Institut
The BCS wave function SUPERCONDUCTIVITY 41
-2 -1 0 1 2
-4
-2
0
2
4
6
e-p interaction
BCS approx. V
0


(
V
C
,
T
F
)
-
1
<
k
|
V
|
k

>
/
q
Figure 3.4: Effective electron-phonon interaction for a mode h
q
. For the plot Eq. (3.2.17) is used with a
small damping /
q
= 0.05. Both the energy and the potential are normalized. The red curve visualizes
the BCS approximation for the attractive potential using an arbitrary magnitude of V
0
=1.5.
3.3.1 Coherent states in a boson eld
Bosonic and fermionic properties can be formulated in second quantization using creation (a

k
) and an-
nihilation (a
k
) operators which usually depend on momentum k. The operators obey commutation or
anti-commutation relations [a, b]

= ab ba reecting the symmetry properties of the wave functions.


In the case of bosons they read,
_
a
k
, a

k
/
_
= a
k
a

k
/
a

k
/
a
k
(3.3.21)
=
k,k
/ (3.3.22)
[a
k
, a
k
/ ] = 0 (3.3.23)
_
a

k
, a

k
/
_
= 0 (3.3.24)
where
k,k
/ is the Kronecker . The simplest case of a coherent state is a superposition of an innite
number of harmonic oscillator wave functions without momentum dependence, [
n
= 1/

n!( a

)
n
[0
with [0 denoting the ground state,
[ = e

[[
2
2

n=0
( a

)
n
n!
[0 = e

[[
2
2
e
( a

)
[0. (3.3.25)
After discussing the Madelung transformation of the Schr odinger equation and deriving the supracurrent
[Eq. (2.3.74)] it becomes clear that writing the complex number as = [[e
i
means establishing a
new wave function. Since is an eigenvalue of the annihilator a, a[ = [ the expectation values
of the number operator N n = [ a

a[ and all its moments can be calculated right away without


referring to Eq. (3.3.25) yielding the following main results:
N = [[
2
, (3.3.26)
N
N
=
_
n
2
n
2
N
=
1

N
, and (3.3.27)
N
1
2
. (3.3.28)
2013
42 R. HACKL AND D. EINZEL Microscopic Theory
Kohrente Zustnde
-1 0 1 2 3
0,0
0,1
0,2
0,3
0,4 <n> = 10
<n> = 25
<n> = 75


<
n
>
1
/
2
p
(
<
n
>
,
n
)
n/<n>
n
<n>
0.4
0.3
0.2
0.1
0.0
Figure 3.5: Poisson distribution P
n
(N) for different expectation values N = n as a function of the
occupation number n. The n-axis is normalized to N. For small N the asymmetry is still visible. The
ordinate axis is normalized to

N
1
since the maximum of P
n
(N) scales as

2N
1
in the limit N .
In the sum O(

N) distributions P
n
(N) contribute for large N. In other words, in absolute units the width
increases while the relative width decreases.
Obviously, [[
2
corresponds to the average occupation number of the state [. The distribution around
N becomes increasingly sharp with increasing N and the amplitude and the phase are conjugate variables.
The phase becomes innitely rigid in the large N limit. Although all properties are describe in this way
it is instructive to write down the expectation value for the number operator explicitly using Eq. (3.3.25),
N = n =

n=0
n
[[
2n
n!
e
[[
2
(3.3.29)
= [[
2
e
[[
2

n=1
n[[
2(n1)
n(n1)!
= [[
2
.
The r.h.s. of Eq. (3.3.29) is a Poisson distribution for n, n =

n=0
nP
n
(N) [see also Eq. (3.3.26)].
If P
n
(N) is plotted for different values of N as a function of n Eq. (3.3.27) can be visualized directly
(Fig. 3.5). If the k dependence is restored wave functions of atoms in a trap or laser elds can be
described,
[ = exp
_

k
a

1
2
[a
k
[
2
_
. (3.3.30)
for instance, being, however, beyond our interest here. Rather we focus now on fermions.
3.3.2 Properties of fermions
Fermions are governed by the Pauli principle stating that each quantum state can only be occupied once.
For this reason there is only the ground state [0 and a singly occupied state [1, and the wave functions
are anti-symmetric as formulated in Eq. (3.1.1). Using creators (c

k
) and annihilators (c
k
), the Pauli
c _ Walther-Meiner-Institut
The BCS wave function SUPERCONDUCTIVITY 43
principle corresponds to the following properties,
c
k
[0 = 0 (3.3.31)
c

k
[0 = [1 (3.3.32)
c
k
[1 = [0 (3.3.33)
c

k
[1 = 0. (3.3.34)
For the anti-commutation relations we use braces
+
in order to make affairs as clear as possible. In
addition, we have to now take care of the spin ,
c
k
, c

k
/

+
= c
k
c

k
/

/
+c

k
/

/
c
k
=
kk
/

/ (3.3.35)
c
k
, c
k
/

/
+
= 0 (3.3.36)
c

k
, c

k
/

+
= 0. (3.3.37)
We dene now n
k
= c

k
c
k
and nd 0[ n
k
[0 = 0, 1[ n
k
[1 = 1, etc. These properties are sufcient
to derive a coherent state of fermions.
3.3.3 A coherent state of fermions
The target is to derive a fermion state similar to that of Eq: (3.3.30). To this end we rst dene pair
annihilation and creation operators
P
k
= c
k
c
k
and (3.3.38)
P

k
= c

k
c

k
, (3.3.39)
respectively, and calculate their properties. P

k
[0 corresponds to the creation of a state with an electron
at k having up-spin and simultaneously one at k having down-spin. Since the two involved fermions
have opposite spin it is sensible to look at commutators rather than anti-commutators, yielding (as shown
in problem 3 of set 5)
[P
k
, P
k
/ ]

= 0, (3.3.40)
_
P

k
, P

k
/
_

= 0, and (3.3.41)
_
P
k
, P

k
/
_

=
kk
/ (1 n
k
n
k
). (3.3.42)
The latter relation is neither a commutator nor an anti-commutator in the usual sense for being different
from 1 or 0 in the general case. We nally need the powers of P

k
for calculating the exponential series,
P

k
P

k
=
_
P

k
_
2
= c

k
c

k
c

k
c

k
= c

k
c

k
c

k
c

k
= 0. (3.3.43)
2013
44 R. HACKL AND D. EINZEL Microscopic Theory
Note that we exchanged c

k
c

k
in the second line producing a minus sign and a series of two equal
creators acting on the ground state thus making the whole operator to vanish. Using this and the anti-
commutator in Eq. (3.3.41) two properties, Schrieffers ground state can now be written down as
[BCS = a exp
_

k
P

k
_
[0
= c

k
exp
_

k
P

k
_
[0
= c

k
_
1+
k
P

k
_
[0. (3.3.44)
For the last transformation we used that all powers of P
k
beyond linear vanish. We set BCS[BCS = 1
for determining the constant c and nd
1 BCS[BCS
= c 0[

k
(1+

k
P
k
)(1+
k
P

k
)[0.
(3.3.45)
The normalization condition can be satised if all factors are unity,
1 = c 0[(1+

k
P
k
)(1+
k
P

k
)[0
= c
2
(1+[
k
[
2
),
yielding the constant c. Using this normalization the BCS ground state can be written down as
[BCS =

k
(u

k
+v

k
P

k
)[0 with (3.3.46)
u

k
=
1
_
1+[
k
[
2
and
v

k
=

k
_
1+[
k
[
2
where
[u
k
[
2
+[v
k
[
2
= 1. (3.3.47)
[BCS is a wave function of a coherent state with the complex number
k
having an amplitude and a
phase. The same holds true for the coherence factors u

k
and v

k
which will be derived in detail later.
Using the conjugate complex follows the convention in Annetts book and is the most popular but not the
only one in the literature. The BCS wave function describes the coherent superposition of the vacuum
and 2, 4, 6. . . electrons. Below we derive the average number of particles in the condensate and its
uctuations and nd full agreement with the coherent state Eq. (3.3.25).
3.4 Determination of the ground state
Using [BCS we have to nd the ground state of the Hamiltonian

H
BCS
of the system by minimizing
E = BCS[

H
BCS
[BCS. There are various ways. BCS used real numbers, u
k
= sin
k
and v
k
= cos
k
,
which obviously satisfy the normalization [Eq. (3.3.47)]. The most direct approach is the Bogoliubov
transformation. However, using Lagrange multiplicators helps visualizing several relevant details (An-
nett). Therefore, we make this detour before deriving the Bogoliubov transformations. The problem we
wish to solve is minimizing the ground state energy, E = BCS[

H
BCS
[BCS. To this end we switch to
momentum space which is equivalent to Fourier transform the Hamiltonian. The technique is named
second quantization.
c _ Walther-Meiner-Institut
Determination of the ground state SUPERCONDUCTIVITY 45
3.4.1 The BCS Hamilonian in second quantization
As already demonstrated in Problem 2 of set 5 the position dependent wave functions

(r) can be
expanded in plan waves. The transformations and the corresponding inverse transformations read

(r) =
1

k
c
k
e
ikr
c
k
=
1

V
_

(r)e
ikr
dr (3.4.48)

(r) =
1

k
c

k
e
ikr
c

k
=
1

V
_

(r)e
ikr
dr. (3.4.49)
The Hamiltonian in conguration space is a sum over all N electrons and has a kinetic energy term, one
from the external elds such as the potentials (r) or A(r), and the interaction between the electrons
which is supposed to lead to superconductivity,

H =

i
_
h
2
2m

2
i
V
ext
(r)
_
+
1
2

i, j
V(r
i
r
j
) (3.4.50)
The factor 1/2 in front of the interaction tern takes care of double counting. If the expansions Eqs. (3.4.48)
and (3.4.49) are inserted the sum over the particles transforms into a k sum, and integrating over all space
completes the Fourier transformation,

H =
_

(r)
_
h
2
2m

2
V
ext
(r)
_

(r)dr+
1
2
_
V(rr
/
)

(r)

(r)

(r
/
)

(r
/
)drdr
/
. (3.4.51)
While the evaluation of the kinetic energy term is straightforward that of the other contributions is more
time consuming and will be left as an exercise. The kinetic energy

T works as follows:

T =
_

(r)
_
h
2
2m

2
_

(r)dr
=
1
V

k,k
/
_
_
c

k
/

e
ik
/
r
h
2
k
2
2m
c
k
e
ikr
_
dr
=

k
h
2
k
2
2m
c

k
c
k
. (3.4.52)
After Fourier transforming the interaction term we can nally write

H =

k,

k
c

k
c
k
+

k
1
,k
2
,q,
1
,
2
V
q
c

k
1
+q,
1
c

k
2
q,
2
c
k
2

2
c
k
1

1
. (3.4.53)
The interaction term can be simplied by using the assumptions for a Cooper pair, k
2
= k
1
and, for a
singlet state,
2
=
1
and the resulting BCS Hamiltonian reads

H
BCS
=

k,

k
c

k
c
k
+

k,k
/
,
V
k,k
/ c

k
c

k
c
k
c
k
. (3.4.54)
The BCS Hamiltonian is a rather general and allows, for instance, for anisotropies in the interaction
potential. It does neither cover triplet pairing nor is it sufcient to deal with exotic cases such as non-
centrosymmetric systems with mixtures of singlet and triplet pairing and spin-orbit splitting of the Fermi
surface. However, for all our purposes below, including the discussion of the cuprates, the iron-based,
and heavy fermion systems with nodes and sign changes of the energy gap
k
. Before deriving the
ground state energy we determine a few expectation values for the BCS wave function.
2013
46 R. HACKL AND D. EINZEL Microscopic Theory
3.4.2 Some expectation values
We now calculate the expectation value for the momentum distribution function for one spin projection,
n
k
= BCS[ c

k
c
k
[BCS
= 0[(u
k
+v
k
c
k
c
k
) c

k
c
k
(u

k
+v

k
c

k
c

k
)[0
= [u
k
[
2
0[ c

k
c
k
[0 +u
k
v

k
0[ c

k
c
k
c

k
c

k
[0 +
+v
k
u

k
0[ c
k
c
k
c

k
c
k
[0 +[v
k
[
2
0[ c
k
c
k
c

k
c
k
c

k
c

k
[0
= [u
k
[
2
0+u
k
v

k
0+v
k
u

k
0+[v
k
[
2
(10[ c

k
c
k
[0)
= [v
k
[
2
. (3.4.55)
The term with the prefactor [v
k
[
2
can be broken down as 0[ c
k
c
k
c

k
c
k
c

k
c

k
[0 = 0[ c
k
(1
c

k
c
k
)(1 c

k
c
k
) c

k
[0 = 0[ c
k
c

k
[0 0[ c
k
c

k
c
k
c

k
[0 0[ c
k
c

k
c
k
c

k
[0 +
0[ c
k
c

k
c
k
c

k
c
k
c

k
[0. The rst term is 1 either since the operators in this order yield 1[1
or can be rearranged as above. In all other cases the last two operators anti-commute, and . . . c[0 always
vanishes.
The total number of particles is just the sum over all k-points,
N =

k,
[v
k
[
2
= 2

k
[v
k
[
2
=

k
(1[u
k
[
2
+[v
k
[
2
). (3.4.56)
With the expectation value of the four-particle interaction operator given by c

k
c

k
c
k
/

c
k
/

=
v
k
v

k
/
u
k
/ u

k
the energy expectation value of the BCS Hamilton operator can be written as
E E = 2

k
[v
k
[
2
+

k,k
/
V
k,k
/ v
k
v

k
/ u
k
/ u

k
. (3.4.57)
Before starting to determine the minimum of E we write down two other expectation values which
will be useful below and are christened Gorkov amplitudes after one of the pioneers of the theory of
superconductivity,
c
k
c
k
= u
k
v

k
g
k
(3.4.58)
c

k
c

k
= u

k
v
k
g

k
, (3.4.59)
which can easily be recognized as the expectation values of the pair annihilator and creator, respectively.
Finally, we look at the uctuations N of N. To this end we need quantities
such as N =
_
N
2
N
2
=
_
2
k
( n
2
k
n
k

2
) =
_
2
k
[v
k
[
2
2
k
[v
k
[
4
=
_
2
k
[v
k
[
2
2
k
(1[u
k
[
2
)[v
k
[
2
=
_
2
k
[u
k
[
2
[v
k
[
2
. This expression is proportional to

N. So
we nd immediately that N/N =

N
1
and realize that the relative uctuations of the particle number
vanish in the limit N (whereas the uctuations N diverge as

N). In other words the particle
number has the statistical behavior of a coherent state [Eqs. (3.3.25)(3.3.27)] meaning that order

N
particles uctuate in an out of the unpaired part of the electrons. Where are they going to?
To nd that out we determine the uctuations of the condensate which have to be determined from
c

k
c

k
c
k
c
k
. The amplitude or particle number uctuations are described by the + sign, and
N is proportional to

N as above and the relative uctuations vanish as

N
1
. Correspondingly, the
absolute uctuations of the phase to be derived from c

k
c

k
c
k
c
k
vanish in limit N as
expected for a coherent state. (I have to prove that mathematically but I believe it is true.)
c _ Walther-Meiner-Institut
Determination of the ground state SUPERCONDUCTIVITY 47
3.4.3 Determination of the energy minimum at T = 0
We show here how the energy minimum for the BCS ground state at T = 0 can be found. The solution
will be included in that of the following section but the derivation here is almost free of formalities
and instructive. The minimum will be calculated with the constraints of (1) the conservation of the
total number of particles and (2) of [u
k
[
2
+[v
k
[
2
= 1. The method used for this procedure is that of
the Lagrangian multipliers. Since there are two constraints we have two factors
1
and
2
the meaning
of which will become transparent during the derivation. Although returning a result identical to that
Bardeen, Cooper, and Schrieffer obtained by using sine and cosine trigonometric functions for u
k
and
v
k
, xing the condition Eq. (3.3.47), the method used in the textbook by Annett has the advantage of
clarifying the physical meaning of the quantities and E
k
and highlighting the quasiparticle concept.
The two constraints are

1
= N

k
(1[u
k
[
2
+[v
k
[
2
) = 0

2
= u
k
u

k
+v
k
v

k
+1 = 0,
and upon setting zero the partial derivatives of E +
1

1
+
2

2
w.r.t. u

k
and v

k
one obtains the eigen-
value equations
(
k
+
1
)u
k
+
k
v
k
=
2
u
k
(3.4.60)

k
u
k
(
k
+
1
)v
k
=
2
v
k
,
where

k
=

k
/
V
k,k
/ u
k
/ v

k
/

k
/
V
k,k
/ g
k
/ (3.4.61)
was used along with the denition in Eq. (3.4.58).
k
is the gap parameter which has the full momentum
dependence here and corresponds to the binding energy of the Cooper pairs and the energy gain due to the
reduction of the electronic energy. The rst multiplier shifts the energy, and we identify
1
= where
is the chemical potential. The second multiplier
2
=E
k
corresponds to the energy of the quasiparticles
excited out of the condensate across the gap (Bogolubov quasiparticles, see next paragraph), and upon
solving the eigenvalue equation we nd
E
k
=
_
(
k
)
2
+[
k
[
2
=
_

2
k
+[
k
[
2
(3.4.62)
[u
k
[
2
=
1
2
_
1+

k
E
k
_
(3.4.63)
[v
k
[
2
=
1
2
_
1

k
E
k
_
(3.4.64)
u
k
v

k
=

k
2E
k
, (3.4.65)
where we have also dened the quasiparticle energy
k
=
k
in the normal state. Eqs. (3.4.63) and
(3.4.64) are the coherence factors and describe the occupation probability of unpaired holes and electrons,
2013
48 R. HACKL AND D. EINZEL Microscopic Theory
respectively. Eq. (3.4.65) describes the occupation of the condensate. Upon inserting Eq. (3.4.65) into
Eq. (3.4.61) the fully momentum dependent gap equation,

k
=

k
/
V
k,k
/

k
/
2
_

2
k
/
+[
k
/ [
2
, (3.4.66)
is obtained. This equation cannot be solved analytically in the general case. If the gap and the interaction
potential are assumed to be k-independent one obtains an equation similar to (3.1.10). After transforming
the sum into an integral the solution for = N
F
[g
eff
[
2
(attractive interaction V =[g
eff
[
2
) and the cut-off
of the attractive range h
0
reads
1 = ln
h
0
+
_
( h
0
)
2
+[[
2
[[
ln
h
0
_
2+
1
4
_
[[
h
0
_
2
_
[[
ln
2 h
0
[[
, yielding
= 2 h
0
exp
_

_
. (3.4.67)
3.5 The general solution
In the preceding section it was shown how the ground state of a superconductor can be constructed
from the coherent wave functions of the condensate [Eq. (3.3.46)]. For deriving the properties at nite
temperature we have to deal with the 4-operator term c

k
c

k
c
k
c
k
in the interaction energy.
3.5.1 Approximation of the four-operator term
Wicks theorem states that four-operator terms can be approximated by a sum of pairs of operators times
the expectation value of the corresponding pair, c

k
c

k
c
k
c
k
c

k
c

k
c
k
c
k
+. . . . This can be
made plausible by using the identities
c
k
c
k
= g
k
+ c
k
c
k
g
k
= g
k
+g
k
c

k
c

k
= g

k
+ c

k
c

k
g

k
= g

k
+g

k
with the Gorkov amplitudes dened in Eqs. (3.4.58) and (3.4.59). The interaction energy can now be
expressed as
E
int
=

k,k
/
V
k,k
/ (g

k
+g

k
)(g
k
/ +g
k
/ )


k,k
/
V
k,k
/ (g

k
g
k
/ +g
k
/ g

k
+g

k
g
k
/ )
=

k,k
/
V
k,k
/
_
g

k
c
k
/

c
k
/

+g
k
/ c

k
c

k
g

k
g
k
/
_
=
_

k
/

k
/
c
k
/

c
k
/

k
_

k
c

k
c

k
g

k
_
_
. (3.5.68)
c _ Walther-Meiner-Institut
The general solution SUPERCONDUCTIVITY 49
Here all terms higher than linear in g
()
k
were neglected, and Eq. (3.4.61) and its conjugate complex
were used. On noting that the quasiparticle energies are symmetric in momentum,
k
=
k
, the kinetic
energy

T can be rewritten as

T =

k,

k
c

k,
c
k,
=

k

k
_
c

k
c
k
+ c

k
c
k
_
=

k
_

k
c

k
c
k

k
c
k
c

k
+
k
_
, (3.5.69)
where Eq. (3.3.35) was applied in the last step. If the preceding two equations are added the mean eld
Hamiltonian

H
MF
N can be written in matrix form as proposed rst by Nambu,

H
MF
N =

k
_

k
+g

k
+
_
c

k
, c
k
_
. .

k
_

k

k

k

k
_
. .

k
_
c
k
c

k
_
. .

C
k
_

_
. (3.5.70)
The operators

C

k
and

C
k
are vectors of two creators/annihilators and are called spinors.
k
is the energy
matrix with the quasiparticle energies in the diagonal and the gap as off-diagonal elements. For nd-
ing the minimum of

H
MF
N the energy matrix has to be diagonalized by a unitary transformation.
The problem is similar to that of a two-level system with nite coupling between quantum mechanical
oscillators.
3.5.2 The Bogoliubov-Valatin transformation
For diagonalization we rst insert the transformation matrices B
k
and B

k
into Eq. (3.5.70) and then
determine their coefcients,

H
MF
N =

k
_

k
+g

k
+

C

k
(B
k
. .
B

k
)
k
(B
k
. .
B

k
)

C
k
. .
_
. (3.5.71)
For a unitary transformation the product of B
k
and B

k
has to yield the unity matrix, B
k
B

k
= 1. Conse-
quently one can write the transformation in terms of u
k
and v
k
etc. with [u
k
[
2
+[v
k
[
2
= 1 as above,
B
k
=
_
u
k
v

k
v
k
u

k
_
(3.5.72)
B

k
(B

k
)
/
=
_
u

k
v

k
v
k
u
k
_
. (3.5.73)
This is the Bogoliubov-Valatin transformation which yields new operators

B

k
=

C

k
B
k
and

B
k
= B

k

C
k
reading

k
= u
k
c

k
v
k
c
k
(3.5.74)

k
= v

k
c

k
+u

k
c
k
(3.5.75)
2013
50 R. HACKL AND D. EINZEL Microscopic Theory
and, respectively,

k
= u

k
c
k
v

k
c

k
(3.5.76)

k
= v
k
c
k
+u
k
c

k
, (3.5.77)
and the new energy matrix
B

k
B
k
=
_
E
k
D
k
D
k
E

k
_
. (3.5.78)
The coefcients u
k
and v
k
are found by the condition D
k
= 0 and the normalization and lead to the same
result as the minimization of the energy as displayed in Eqs. (3.4.62)(3.4.65). The result is equivalent
to the solution of the eigenvalue equation Eq. (3.4.60) and corresponds to nding the ground state. Here
we made an important additional step by deriving the new

operators. The two equations (3.5.74)
and (3.5.76) dene these new operators as a coherent superposition of creators and annihilators, hence
electrons and holes, at k and k. It can be shown directly by insertion that the

operators obey fermion
anti-commutation relations,

k
,

k
/

+
=
kk
/

/ (3.5.79)

k
,

k
/

/
+
= 0 (3.5.80)

k
,

k
/

+
= 0. (3.5.81)
Consequently, in the superconductors there are no electrons or hole any further but, rather, composite
fermions carrying the name Boguliubov quasiparticles and the condensate of Cooper pairs after their
inventors. The probability of nding one of those objects in energy and momentum space is given
by the coherence factors u
k
and v
k
which can be visualized particularly easily by assuming a linear
electronic dispersion around the Fermi energy and momentum,
k
hv
F
(k k
F
) with v
F
and k
F
the Fermi velocity and the Fermi momentum. In the state k, k with particle-hole mixing every band
crossing the chemical potential is reected about . Since the particles in the reected band are those
of the original one the two bands mix in the usual quantum mechanical fashion and the levels repel each
other by virtue of the interaction energy 2
k
. At the intersection point k
F
[u
k
[
2
and [v
k
[
2
are both 1/2
and the character of the Boguliubov quasiparticles is exactly 50% electronic and 50% hole-like. Away
from k
F
the character changes between 0 and 100%. Fig. 3.6 shows a theoretical result for particle-
hole mixing in a strong coupling material, here the cuprate Bi
2
Sr
2
CaCu
2
O
8
. (From [15].) The band is
reected about and the color-coded occupation probability is given by the coherence factors shown in
Fig. 3.7.
If the new

operators [Eqs. (3.5.74) and (3.5.76)] and the coherence factors [Eqs. (3.4.63) and (3.4.64)]
are plugged back into Eq. (3.5.71)

H
MF
N can be separated into a ground state part and a part of
thermally excited quasiparticles,

H
MF
N =

k
_

k
+g

k
+
_

k
,

k
_
_
E
k
0
0 E
k
_
_

k
__
=

k
_

k
+g

k
_
+

k
_
E
k

k
E
k

k
_
=

k
_

k
+g

k
_
+

k
_
E
k

k
+E
k

k
E
k
_
=

k
_

k
+g

k
_
+

k,
_
E
k

k,

k,
E
k
_

H
MF
N =

k
_

k
E
k
+g

k
_
+

k,
E
k

k,

k,
. (3.5.82)
c _ Walther-Meiner-Institut
The general solution SUPERCONDUCTIVITY 51
0.10
0.05
0.00
-0.05
-0.10
E
n
e
r
g
y

k
(
e
V
)
=0 meV =20 meV
0
0

a

a
Figure 3.6: Dispersion
k
along the (0, 0) (, 0) (X) direction, typical for a cuprate superconduc-
tor, above (left) and below T
c
. The chemical potential (yellow horizontal line) is at zero energy in the
quasiparticle picture, the Fermi momentum is indicated in red. (The blue vertical line is an artifact.)
Shown here is the result of a strong coupling calculation in false color representation with warm colors
indicating the maxima of the spectral functions. Above T
c
the band crosses = 0 at k = k
F
. Below T
c
the band is reected about . The two bands intersecting at k
F
interact similarly as states in a two-level
system. The spectral weight changes according to the coherence factors plotted in Fig. 3.7. For every
xed k the integral over all energies is unity corresponding to one quasiparticle. From [15].
-6 -4 -2 0 2 4 6
0.0
0.2
0.4
0.6
0.8
1.0
|
u
k
|
2
,

|
v
k
|
2
,

u
*
v
,

f
(

k
)
Quasiparticle energy
k
()
|u*u|
|v*v|
|u*v|
f
0.0 0.5 1.0
0.0
0.5
1.0
(
0
)
-1
/k
F
=/t =0.2
|
u
k
|
2
,

|
v
k
|
2
,

u
*
v
k (/a)
Figure 3.7: Coherence factors [u
k
[
2
, [v
k
[
2
, and [u
k
v

k
[ for electron, hole, and pair occupation, respectively.
(a) shows the coherence factors as a function of momentum. The underlying band structure corresponds
to the 2D tight-binding representation
k
=t[cos(k
x
a) +cos(k
y
a)] (a is the lattice constant) along
k
y
= 0 at half lling, t. For better visualization the gap is as large as 10% of . For a conventional
metallic superconductor / = O(0.001) in the cuprates the maximal gap may reach 510% of . The
half width (FWHM) k of [u
k
v

k
[ denes the BCS coherence length,
BCS

0
= 2/k. (b) shows the
coherence factors as a function of energy. The Cooper pairs live around the Fermi energy but extend
substantially into the band. As a result all electrons including the unpaired ones are enslaved and have
the same phase. The Fermi function at T = 1.76T
c
, f (
k
, 1.76T
c
), has an energy dependence similar to
that of [v
k
[
2
.
The

operators obey fermionic anti-commutation relations [Eqs. (3.5.79)(3.5.81)]. Hence the expec-
tation values of the corresponding number operators are described by Fermi statistics,

k
= (E
k
)
1
exp
_
E
k
k
B
T
_
+1
(3.5.83)

k
= 1(E
k
). (3.5.84)
Here we have written the Fermi distribution function as (E
k
) rather than f (E
k
) as usual in order to
2013
52 R. HACKL AND D. EINZEL Microscopic Theory
highlight that the energy of a Bogoliubov quasiparticle is in the argument. These equations show that the
second term on the r.h.s. of Eq. (3.5.82) describes the kinetic energy of free Fermions having dispersion
E
k
. It vanishes in the limit T 0 and in the limit 0. (Note that for correctly determining the
transition into the normal state the limit is lim
0
E
k
= [
k
[ at any temperature.) Since the second term
vanishes in the limit T 0 the rst term is the ground state energyU
BCS
. Using these limits appropriately
Eq. (3.5.82) is the basis for determining the energy gain in the superconducting state (see Problem 2 in
set 7) yielding
E
T=0
=

H
MF
N

H
n

=
1
4
N
F

2
(3.5.85)
for a constant gap and N
F
the density of states at the Fermi energy E
F
for both spin projections (which
is twice the number of k points).
From Eq. (3.5.83) the expectation value of the Gorkov amplitude at nite temperature can be derived
(see problem 1, set 7),
c
k
c
k
= 12(E
k
)u
k
v

k
= tanh
_
E
k
2k
B
T
_
u
k
v

k
(3.5.86)
with u
k
v

k
given by Eq. (3.4.65). This result enables us to write down the gap equation at nite tempera-
ture.
3.5.3 Solution of the gap equation for T 0
Together with Eq. (3.5.86) the gap equation (3.4.61) with the denition Eq. (3.4.58) can be written down
as a major result of this section,

k
=

k,k
/
V
k,k
/ c
k
/

c
k
/

=

k,k
/
V
k,k
/ u
k
/ v

k
/ tanh
_
E
k
2k
B
T
_

k
=

k,k
/
V
k,k
/

k
/
2E
k
/
tanh
_
E
k
2k
B
T
_
. (3.5.87)
This equation reproduces the result of Eq. (3.4.67) of section 3.4.3 if the limit T = 0 is taken. At nite
temperature one can calculate the limit 0 and derive the transition temperature T
c
(where the gap
vanishes). To simplify we use the BCS approximation, as above, and assume that
k
= is constant, the
dispersion is parabolic and isotropic, and that the interaction is attractive and also isotropic yielding
1
N
F
[g
eff
[
2

1

=
_
h
0
0
d
E
tanh
_
E
2k
B
T
_
. (3.5.88)
The limit E [[ and the substitution x = /2k
B
T
c
yield
1

=
_
h
0
/2k
B
T
c
0
dx
tanhx
x
= ln
_
2e

h
0
2k
B
T
c
_
, (3.5.89)
c _ Walther-Meiner-Institut
The general solution SUPERCONDUCTIVITY 53
where = 0.5772 is the Euler constant. Combining this result and that of Eq. (3.4.67),
k
B
T
c
=
e

2 h
0
e

(3.5.90)
= 2 h
0
e

returns the famous universal BCS relationship between the gap and the transition temperature,

k
B
T
c
=

e

= 1.764. (3.5.91)
In the majority of cases the ratio of twice the gap to T
c
is given, 2/k
B
T
c
= 3.528. Note that any
dependence on materials properties, in particular the cutoff energy h
0
, corresponding to the Debye
frequency h
D
in the case of electron-phonon coupling, and the electron-phonon coupling parameter
drop out entirely. While this is typical for a mean eld theory as here it is not realistic for most of the
materials. Only a few of the elements such as Al, Sn or V are close to the BCS prediction. Elements like
Pb or Nb and most of the alloys have gap ratios well above 3.53 and reach 5.6 in Nb
3
Sn. This means
that the inuence of the cutoff or, more physically, of details of the interaction cannot be ignored any
further. Eliashberg solved that problem in 1960 [3]. In essence the solution takes into account the full
energy dependence of the interaction potential leading to energy dependent gap functions as well. The
treatment requires the knowledge of Greens function techniques, and the discussion is far beyond the
scope of these lecture notes.
If the interaction potential V
k,k
/ is either repulsive or strongly momentum dependent the gap function

k
may become momentum dependent as well and may even change sign on the Fermi surface [16, 17].
What means a sign change of the gap? Formally, this is a simple question since we can write down the
gap in terms of k-dependent functions such as

k
=

0
2
[cos(k
x
a) cos(k
y
a)] , (3.5.92)
where a is the lattice constant of a square lattice. Eq. (3.5.92) reproduces the d
x
2
y
2 gap in the copper-
oxygen compounds, for instance. Fossheim and Sodb describe in some detail how this gap follows
from a specic choice of V
k,k
/ . The problem with visualizing such a gap lies in the way it is measured
experimentally. In the typical spectroscopic experiments such as tunneling, photoemission, optical or
Raman spectroscopy or thermodynamic experiments such as heat capacity, heat transport or penetration
depth only the magnitude of the gap [
k
[ enters. In addition, it is hard to imagine howan energy reduction
below the chemical potential can be both positive and negative. The solution is that the gap has a phase
deriving from the wave function of the coherent state. This phase enters whenever the relative phases of
two coupled condensates are probed. Also remember that the expression for the current contains a phase
gradient. Consequently, whenever experiments with coupled superconductors are performed the phase
and, hence, the sign of
k
becomes intuitively clearer. The seminal phase-sensitive experiments in the
cuprates [18, 19] will be discussed in more detail in chapter 6.
A momentum-dependent gap raises additional questions as to the possible outcome of thermodynamic
experiments or the condensation energy. In the latter case Eq. (3.5.87) has to be solved for the respec-
tive momentum-dependent gap. In the case of d
x
2
y
2 gap in Eq. (3.5.92) the mean eld condensation
energy leads to a maximal gap larger than the canonical BCS value, and 2
0
= 4.28. In the cuprates
one nds 2 8 indicating very strong coupling. If there is a multiband system the smallest gap can
be well below 3.53 while the largest one always exceeds 3.53 since otherwise the condensation energy
is not reached. Concerning thermodynamics the answer is similarly complicated. For a system with a
nite gap everywhere on one of the Fermi surfaces one typically nds activated behavior, i.e. a response
being characterized by a gap. Then the smallest gap dominates (but does not exclusively determines) the
2013
54 R. HACKL AND D. EINZEL Microscopic Theory
activation energy since, with increasing temperature, the branches with the smaller Bogolyubov energy
E
k
are populated rst. If the gap has nodes, (k
0
) = 0, there is no activated behavior any further because
unpaired particles exist whenever T ,= 0, hence always. Then the specic heat, the penetration depth
or the thermal conductivity are described by temperature power laws rather than exponentials. Promi-
nent examples are UPt
3
, UBe
13
[13] and YBa
2
Cu
3
O
7
[20]. For predictions the explicit variation with
temperature of the gap and of the condensate density is required.
Neither the temperature dependence of the gap nor that of other thermodynamic quantities can be written
down explicitly. Rather the implicit Eq. (3.5.88) needs to be solved. M uhlschlegel derived numerical
values which are tabulated [21]. In addition, interpolation schemes can be derived [22] which provide
an analytic expression for 0 < T T
c
which agrees with the exact numerical solution to better than a
fraction of a percent and is sufcient for most of the practical purposes,

0
(T)

0
(0)
= tanh
_
k
B
T
c

0
(0)

2
3
c
s
c
n
c
n

T
c
[
0
(0)]
2
[
k
(0)]
2

FS
_
T
c
T
1+0.191
T
T
c
_
_
. (3.5.93)
Note that the T
c
/T term is not dened at T = 0 while the limit T 0 exits. The result of Eq. (3.5.93),
as plotted in Fig. 3.8, has the right asymptotic behavior and can be used for estimates of other prop-
erties. It is derived for the weak-coupling BCS limit but it is also valid approximately for the strong
coupling case. In particular, it is also valid for superconductors with anisotropic gaps. Then the gap ratio
0.0 0.2 0.4 0.6 0.8 1.0
0.0
0.2
0.4
0.6
0.8
1.0

(
T
)
/

(
0
)
T/T
c
Figure 3.8: Temperature dependence of the superconducting energy gap according to Eq. (3.5.93). The
deviations from the numerical solution [21] is of the order of a percent.

0
(0)/k
B
T
c
(see above), the discontinuity of the heat capacity
c
s
c
n
c
n
at T
c
, and the Fermi surface average
of [
k
/
0
(0)]
2
have to be used accordingly. In the case of an isotropic gap one nds 1.764, 1.426, and
1, respectively, for the so-called BCS-M uhlschlegel parameters. For the case of d
x
2
y
2 gap [Eq. (3.5.92)]
in a tetragonal system, which, cum grano salis, applies for the cuprates, the weak-coupling values are,
respectively, 2.14, 0.951 and 0.5 [22]. The inuence of replacing the BCS-M uhlschlegel parameters on

0
(T) is mild. Other quantities may change substantially.
3.6 Connection to experiments
In this section we derive a few quantities which follow from the BCS theory and can be observed experi-
mentally including thermodynamic properties, the tunneling density of states, and two-particle properties
for which the coherence factors play an important role.
c _ Walther-Meiner-Institut
Connection to experiments SUPERCONDUCTIVITY 55
3.6.1 Thermodynamic properties
The solution of the gap equation for all temperatures enables us to determine the energy o the Bogoliubov
quasiparticles E
k
(T) = [
2
k
+[
k
(T)[
2
]
1/2
and their thermal occupation (E
k
(T)) leading directly to the
thermodynamic functions. For transforming the k sums into integrals the momentum dependence will
be dropped occasionally. Since the Bogoliubov quasiparticles are a system of free fermions and since
the condensate does not contribute for being a fully ordered state the density of electronic entropy in the
superconducting state comes only from the quasiparticle and can be written in terms of (E
k
(T)),
s
s
=2k
B
k
[
k
ln
k
+(1
k
)ln(1
k
)] . (3.6.94)
From Eq. (3.6.94) the heat capacity c
s
follows directly as
c
s
= T
ds
s
dT
. (3.6.95)
Some writing can be saved if one takes the derivative w.r.t. = (k
B
T)
1
,
c
s
=
ds
s
d
= 2k
B
k

k
E
k
_
E
2
k
+
1
2

d
d
[
k
(T)[
2
_
. (3.6.96)
Here the derivative of the occupation number w.r.t. the quasiparticle energy leads directly to the Yosida
function Y(T) which enters many thermodynamic properties,
N
F
Y(T) =

k
E
k
=
N
F
4k
B
T
c
_

d
_
cosh
_
E
k
B
T
__
2
(3.6.97)
which is exponentially small at low temperature for an isotropic gap and shows directly that the heat
capacity show activated behavior then. At T
c
the heat capacity of the normal state is also needed and can
be derived from Eq. (3.6.96) by setting the gap zero reproducing the result (2.2.42),
c
n
= T =

2
3
k
2
B
N(E
F
)T.
From that and Eq. (3.6.96) one nds for an isotropic gap,
c(T
c
) = N(E
F
)
_
d
dT
[(T)[
2
_
and for the ratio to the normal-state value another universal number,
c
c
n
= 1.43. (3.6.98)
Using the result in Eq. (3.6.96) the internal energy and nally the free energy density in the normal and
the superconducting state and, thereof, the critical eld as a function of temperature can be found. Now,
by using the BCS expressions, the variation with T is slightly different from the parabola as shown in
Fig. 3.9. As mentioned earlier weak coupling superconductors are described well by the BCS prediction
while those with strong coupling rst get closer to the parabola and then even go beyond. Fig. 3.10 shows
the temperature dependence of the thermodynamic functions as derived above.
2013
56 R. HACKL AND D. EINZEL Microscopic Theory
Figure 3.9: Temperature dependence of the critical eld. The lower curve follows from the BCS predic-
tion the upper one represents the parabola from the two uid model proposed rst by Gorter and Casimir.
From Ref. [2].
Figure 3.10: Temperature dependence of the thermodynamic function according to the BCS model. From
Tinkham.
3.6.2 Single particle response
Since the condensate is a coterie only pairs of electrons can be added or removed (Andreev). Therefore,
if single particle properties are being studied, only the quasiparticle contribute and the response from
the condensate can be disregarded. This holds for (single particle as opposed to Josephson) tunneling
and photoemission spectroscopy for instance whenever the nal state is outside the superconductor. One
of the most important quantities in this is the single particle density of states N(E) which is a momen-
tum integrated function. Since the number of particles above and below T
c
is conserved on can simply
equate N()d = N(E)dE and, using the approximation N() N(0) in the normal state, nds for the
c _ Walther-Meiner-Institut
Connection to experiments SUPERCONDUCTIVITY 57
superconducting density of states,
N(E)
N(0)
=
d
dE
=
_
E

E
2
+
2
E >
0 E <,
(3.6.99)
exhibiting the typical square root singularity at the gap edge. This result is well reproduced in weak-
coupling isotropic superconductors.
2013
Chapter 4
Ginzburg-Landau Theory
The model of Ginzburg and Landau (GL) proposed 1950 is an extension of the London theory and allows
the treatment of spatially inhomogeneous mixtures of normal and superconducting regions in a material.
In this way all congurations of currents and elds in a superconductor can be treated. Therefore the
GL theory has a wide range of applications in practice. Gorkov showed in the beginning of the 1960ies
that the BCS theory and the GL model are equivalent in the temperature range around T
c
where the GL
theory is valid.
4.1 Phase transitions
Before discussing the results for superconductors we briey summarize the theory for second order phase
transition originally proposed for magnets by L. D. Landau in 1937. Landau expanded the density of the
free energy in powers of the magnetization M being a natural order parameter different from zero in
the ordered state (T < T
c
) and identical zero above. Hence the description of magnetism is a clear-cut
problem. The purpose of the theory, formulated a long time before numerical methods were successfully
introduced to the problem of magnetism, was essentially to derive thermodynamical properties in the
ordered state and later also above in the uctuation regime from the simplest possible assumptions. Then
the density of the free energy can be written as
f
M
= f
n
+(p, T)M(r)
2
+
(p, T)
2
M(r)
4
. (4.1.1)
Without external elds only the magnitude of M(r) =[M(r)[ enters. In the simplest case (no position de-
pendence) the free energy becomes minimal if the rst derivative of f vanishes and the second derivative
is positive,
0 =
f
M
M
= 2(p, T)M+2(p, T)M
3
(4.1.2)
0 <

2
f
M
M
2
= 2(p, T) +6(p, T)M
2
. (4.1.3)
A non-trivial solution of Eq. (4.1.2) is
M
2
=

(4.1.4)
which has a real solution only if and have opposite sign and ,= 0. This is fullled at T < T
c
for
(T) =
0
(T T
c
) (4.1.5)
59
60 R. HACKL AND D. EINZEL Ginzburg-Landau Theory
with
0
and positive constants. If M
2
from Eq. (4.1.4) is inserted into Eq. (4.1.3) 4 is obtained
being positive in the ordered phase, and the extremum of f
M
is a minimum if and only if < 0 and
> 0. The evolution of the free energy landscape for various temperatures is shown in Fig. 4.1. The
-1.0 -0.5 0.0 0.5 1.0
0.00
0.05
0.10


f
(
p
,
T
,
M
)
M/M
0
= 0.1
= 0.03
= 0
=-0.03
=-0.1
Figure 4.1: Free energy density in a magnetic system as a function of the temperature. The temperature
is expressed in terms of . For negative the minimum of f
M
is found at nite magnetization M.
difference
f
M
f
n
=

2
2
(4.1.6)
corresponds to a reduction of the free energy in the ordered state and opens the way towards the derivation
of all thermodynamic properties as in section 2.2. With given in Eq. (4.1.5) the transition is of second
order, implying that there is no volume and entropy change and, consequently, no latent heat at T
c
. Then
the Ehrenfest relations [?] between the discontinuities in the heat capacity c
p
, the thermal expansion,

T
, and the compressibility,
T
can be derived (Landau-Lifshitz, Statistical Physics),

T
c
=
dp
dT

T
c

T
c
(4.1.7)
c
p
T
c
=
dp
dT

T
c

T
c
. (4.1.8)
The pressure dependence of the transition temperature T
c
, dT
c
/dp is closely (but certainly not trivially)
related to microscopic mechanisms at the origin of a phase transition, as can be seen for instance from
the Eq. (3.5.90) in the case of superconductivity, and is therefore a highly desirable quantity. Since c
p
and
T
are experimentally accessible Eq. (4.1.8) can be used to indirectly determine dT
c
/dp. This is
sometimes simpler than a direct measurement of T
c
in a pressure cell. In any case, one obtains an internal
thermodynamic consistency check and can get the compressibility from Eq. (4.1.7). Obviously, this type
of approach proves very useful already in its simplest version. Ginzburg and Landau showed that it can
be applied to superconductivity and even well beyond [12].
4.2 Application to superconductivity
A major problem is the proper denition of an order parameter. One could argue that the temperature
dependent condensate density of the London theory would be the right choice. However, it becomes im-
mediately clear that one would not even get a current where the phase gradient plays a crucial role. This
c _ Walther-Meiner-Institut
Application to superconductivity SUPERCONDUCTIVITY 61
was the motivation for Ginzburg to suggest an order parameter in the spirit of the Madelung substitution
for the wave function, Eq. (2.3.65),
(r, t) = a(r, t)e
i(r,t)
,
where a(r, t) and (r, t) are real functions of position and time. In the GL theory the function depends
on also temperature T is normalized as
2[(r, t, T)[
2
= 2a(r, t, T)
2
= n
s
(T) = nf
s
(T). (4.2.9)
where n is the constant density of electrons in a material, and n
s
(T) is the number of condensed electrons
that is twice the number of Cooper pairs. f
s
(T) is a dimensionless function that vanishes at T
c
and reaches
unity if all electrons are condensed in the limit T 0. Note that the GL model is valid only for T T
c
.
4.2.1 Density of the free energy
The density of the free energy of a superconductor can be derived in the spirit of Eq. (4.2.16) as long as
there are no external elds or currents,
f
s
= f
n
+(p, T)[(r, t)[
2
+
(p, T)
2
[(r, t)[
4
(4.2.10)
f
n
+(p, T)[a(r, t)[
2
+
(p, T)
2
[a(r, t)[
4
, (4.2.11)
which has solutions equivalent to those of Eq. (4.2.16). Here, only the amplitude is important for nding
the minimum, and the phase can still be chosen arbitrarily as shown in Fig. 4.2. In the gure this
f
s
()
Im
Re
Figure 4.2: Free energy density in a superconductor with out eld and currents. The energy minimum
depends only on the amplitude of but is independent of the phase .
corresponds to moving the order parameter around the bottom of the valley. Moving out of the valley is
an amplitude uctuation for which energy must be invested.
As soon as there are currents and elds the kinetic energy of the electrons and the eld energy becomes
important. In addition there is an interrelation. The kinetic and eld energy are given by
1
2
Mv
2

1
2M
[
h
i
(r, t) QA(r, t)[
2
(4.2.12)
and, respectively, by
W =
1
2
0
_
B
2
M
dV =
1
2
0
_
[B(r) B
0
]
2
dV (4.2.13)
2013
62 R. HACKL AND D. EINZEL Ginzburg-Landau Theory
neglecting demagnetization effects. The magnitude is taken for arriving always at real numbers for the
energies. Nowadays one would call that an educated guess since the result is correct but there is no
justication otherwise. Since all functions depend now on position only appropriate integrals over the
sample and over all space lead to the proper energy minimum corresponding to deriving and minimizing
the functional of the free energy.
4.2.2 Functional of the free energy
To this end the energies related to the order parameter are integrated over the sample volume while the
magnetic energies have to be taken in the entire space yielding
F
s
= F
s
+
_
sample

+

2
(

)
2
+
h
2
2M
..

+
Q
i h
A

2
dV +
1
2
0
_

[B(r) B
0
]
2
dV. (4.2.14)
Originally the parameter was used in order to indicate that the third term in the rst integral is another
expansion coefcient. It will turn out, as expected, that M and Q are the mass and the charge of a
Cooper pair having being unknown at the time the GL theory was proposed. For nally deriving the GL
equations the variation of Eq. (4.2.14) has to be determined. Since the absolute value is inconvenient in
this context we rewrite the term,
h
2
2M
_
sample

+
Q
i h
A

2
dV =
h
2
2M
_
sample

_
+
Q
i h
A
_
dV +
i h
2M
_
S

(i h+QA) ndS,
(4.2.15)
where S is the sample surface and n the surface normal in point r. In addition we will need the variation
of the magnetization reading

1
2
0
_
B
2
M
dV =
1

0
_
B
M
B
M
dV
=
1

0
_
B
M
(A
M
)dV
=
1

0
_
(B
M
) A
M
(B
M
A
M
)dV
=
1

0
_
(B
M
) A
M
dV (4.2.16)
4.2.3 The Ginzburg-Landau equations
For the deriving the GL differential equations the variational derivative of F has to be evaluated. This
means nding the change F if or A are replaced by + and A+A. If F is minimal
F

= 0;
F
A
A = 0
where for traditional reasons the variation of

is considered, and the subscript M referring to the


magnetization has been dropped for simplicity. The variation of

is obviously calculated only inside


c _ Walther-Meiner-Institut
Two new length scales SUPERCONDUCTIVITY 63
the rst integral which vanishes only if the integrand vanishes for all possible

yielding,
1
2M
_
h
i
QA
_
2
+( +

) = 0 (4.2.17)
_
h
i

n
QA
n
_
= 0 (4.2.18)
_
h
i

n
QA
n
_
= b. (4.2.19)
Eq. (4.2.17) is the rst Ginzburg-Landau equation (GL1) from which most of the properties can be
derived. Eq. (4.2.18) results from the second integral of Eq. (4.2.15) over the sample surface and says
that the gradient of the wave function perpendicular to the surface vanishes in zero eld meaning that
no current ows in or out and that there are no variations of the pair density close to the surface as
we have seen already before. If one attaches a metal to the surface of a superconductor one arrives
at Eq. (4.2.19). Note that deriving such an equation implies that Eq. (4.2.15) is valid also outside the
superconductor, at least in a small distance. This observation is called proximity effect and is consistent
with the microscopic theory in that the coherent wave function of the pairs leaks out of the material
over a length scale given by the coherence length. Along with the proximity effect, which means that
superconductivity can be induced in a normal metal attached to a superconductor, the Josephson effect is
another important consequence.
The variation of the vector potential A leads to the second Ginzburg-Landau equation (GL2) which
describes the current density in the presence of superconductivity,
j
s
=
h
i
Q
2M
(

)
Q
2
M
A

(4.2.20)

Q
2M
_

_
h
i
QA
_
+c.c.
_
(4.2.21)
where c.c. means complex conjugate and indicates that the current density is a real quantity. If we insert
again the Madelung wave function we obtain
j
s
=
a
2
Q
2
M
_
h
Q
A
_
(4.2.22)
and with the usual replacement Q 2e (traditionally no minus sign here), M 2m, and 2a
2
n,
j
s
=
ne
2
m
_
h
2e
A
_
(4.2.23)
reproducing the London result Eq. (2.3.74).
4.3 Two new length scales
The equations for the current (4.3.25) and (4.2.22) and the rst GL equation (4.2.17) without eld lead
to two new length scales tha turn out to be similar to the scales
L
and
0
. However, though being related
some care has to be taken to keep the denitions clear and to understand the meaning properly. We shall
outline the relation without derivation later.
2013
64 R. HACKL AND D. EINZEL Ginzburg-Landau Theory
4.3.1 Screening
For a moment we go back to Eq. (4.2.22) and take the curl,
j
s
=
a
2
Q
2
M
(A) =
a
2
Q
2
M
B (4.3.24)
and nd upon using Amp` eres law
(B) =
0
j
s

2
B =
0
a
2
Q
2
M
B. (4.3.25)
Eq. (4.3.25) is formally equal to Eq. (2.3.76), and if a
2
=

, Q = 2e, M = 2m are substituted the GL


penetration depth is obtained,

2
GL
=
m
2
0
[[e
2
. (4.3.26)
It is tempting to use 2a
2
= n as above to recover the London penetration depth but then we get the zero
temperature limit while the GL theory is valid only close to T
c
. However, we can insert the temperature
dependence of and nd
GL
to diverge as

T
c
T
1
at T
c
and, in the clean limit,

GL
(T, ) =

L
(0)
_
2
T
c
T
T
c
. (4.3.27)
4.3.2 Ginzburg-Landau coherence length
If A = 0 in the rst GL equation (4.2.17) and a
2
0
= / is substituted for a real equilibrium order
parameter GL1 can be made dimensionless by dividing the entire equation by a
0
,
h
2
4m[[

2
f
x
2
+ f + f
3
= 0. (4.3.28)
Here the problem is conned to one dimension and f = a/a
0
. For dimensional reasons the prefactor of
the second derivative w.r.t. to x must by a length squared and denes already the GL coherence length

GL
,

2
GL
(T) =
h
2
4m[
0
(T T
c
)[
(4.3.29)
the meaning of which becomes transparent by looking at small deviations from the equilibrium, f =
f 1. By neglecting all terms of order ( f )
2
and higher one nds a differential equation for f ,

2
x
2
( f ) =
2

2
GL
( f ), (4.3.30)
which identies
GL
as the healing length of a perturbation to the equilibrium order parameter. With
this length scale we have a phenomenological argument in favor of Eq. (4.2.19) and observe that the
order parameter does not change abruptly but rather varies continuously over a characteristic distance

GL
. However, similarly as in the case of the penetration depth,
GL
is different from the BCS
0
in
both magnitude and meaning. While the magnitudes of
GL
and
0
differ only by a factor of order 1
c _ Walther-Meiner-Institut
Two new length scales SUPERCONDUCTIVITY 65
the interpretation of
GL
is purely phenomenological. It has neither a connection to Cooper pairs nor to
non-locality: the GL theory is strictly local. However, in contrast to the London limit with the coherence
length exceeding the penetration depth
GL
is the smallest length scale in those cases where the GL
theory develops its full power. Hence, effects of non-locality can be ignored safely.
The ratio of the penetration depth and the coherence length is an important quantity and we dene the
dimension less GL parameter,
=

GL

GL
=


2
0
2m
he
=


2
0
1

B
. (4.3.31)
where
B
is the Bohr magneton. depends only weakly on temperature through , so in a rst approx-
imation it can be considered constant close to T
c
. However, Eq. (4.3.31) and, similarly, the equations
(4.3.29) and (4.3.27) conceal that
GL
and
GL
are not solely given by constants of nature but, rather,
depend in subtle ways on the BCS coherence length
0
and the mean free path of the electrons. The
functional dependence is obtained by comparing the results of BCS and GL close to T
c
. While the de-
pendences of
GL
(
0
, ) and
GL
(
0
, ) are very important whenever real materials are to be analyzed
quantitatively they will not particularly further the basic understanding here. We study now the conse-
quences of varying . In practice this means looking at different (clean) materials covering a range of
approximately 0.1 < 100.
4.3.3 Energy of the normal-superconductor interface
We consider now an innite superconductor. At x < 0 the magnetic eld is homogeneous, points along
the z-axis, and just reaches the critical eld B
c
. For x 0 the eld decays, hence the surface x = 0 sepa-
rates the normal from the superconducting part. Then superconductivity is almost completely suppressed
at x = 0 implying that the order parameter, i.e. the density of pairs a
2
, at the surface is vanishingly small
and negligible in comparison to the equilibrium value a
2
0
deep inside the superconductor for x . On
the other hand, the magnetic eld is screened over a typical length scale of a few
GL
.
What is energetically more favorable, small or large ? Small means that the eld is pushed out over a
short length scale while condensation energy is lost in the range
GL

GL
needed for a/a
0
to recover
to unity. In the opposite case the eld can penetrate over a much larger range than that needed for a to
approach a
0
. Obviously, condensation energy is lost here in a smaller volume than in the previous case,
and less energy has to be invested to keep the eld out. This situation is energetically more favorable and
can be written down by considering the free energy density for x 0,
_
f
s
dx =
_
_
f
n

B
2
c
2
0
_
dx +(
GL

GL
)
B
2
c
2
0
. (4.3.32)
The last term is an estimate for the trade-off between condensation and eld energy. The energy is readily
written down since the condensation energy has exactly the same magnitude as the the energy stored in
the eld. The signs of
GL
and
GL
are determined in a way that the free energy increases with
GL
and decreases with
GL
. In this way a surface energy can be dened. The sign of determines as to
whether or not the eld penetrates,
=f
B
2
c
2
0
(
GL

GL
) (4.3.33)
where a negative sign corresponds to an energy gain. For calculating numbers the functional form of the
eld penetration and the recovery of a/a
0
cannot be neglected but will not be considered here since it
2013
66 R. HACKL AND D. EINZEL Ginzburg-Landau Theory
will be derived in a very simple way below. The result yields a limiting value for for which changes
sign,
<
1

2
typeI (4.3.34)
>
1

2
typeII (4.3.35)
dividing the world into type I superconductors which exclude the eld always completely except for the
thin surface layer carrying the screening current and type II superconductors into which the eld can
penetrate. We have seen earlier that vortices are formed which carry one ux quantum each. Since
reaches values in the range of 100 for copper-oxygen superconductors the upper critical eld up to which
superconductivity survives can be orders of magnitude higher than the eld B
c1
at which the rst ux
line penetrates and the thermodynamical critical eld B
th
c
(to be dened later).
4.4 States with internal ux
For > 1 it becomes increasingly favorable for the superconductor to let the eld penetrate as observed
rst by Shubnikov and coworkers (see Fig. 2.2). The consequences are far-reaching and the basis of
many applications such as coils for generating high magnetic elds. The way the eld penetrates rst
was discussed in section (2.3.4), and the eld penetration depth
L
was found to be the characteristic
dimension of the vortex while a mysterious cutoff
L
was postulated to keep the problem tractable
by avoiding a divergence in the center of the vortex. It is intuitively clear now that the scale
0
introduced there is related to (but not equal to) the GL coherence length
GL
. We show now that
GL
is
intimately related to the upper critical eld B
c2
at which superconductivity collapses.
4.4.1 The upper critical eld B
c2
To this end the rst GL equation (4.2.17) has to be analyzed. Assuming that the applied eld B
0
is
only slightly below B
c2
Eq. (4.2.17) can be linearized since the order parameter

is suppressed
substantially below its equilibrium value [
0
[
2
= / and

[[. The linearized equation


reads
h
2
2M
(i+QA)
2
=[[. (4.4.36)
Eq. (4.4.36) is formally equal to a Schr odinger equation for a particle with mass M and charge Q in a
eld B =A, and the problem can be mapped on a known one. For doing so, the eld is assumed to
point along the z-axis, B = B
0
e
z
, and a gauge is chosen for which A = e
y
B
0
x. With the abbreviations

c
=
QB
0
M

eB
0
m
(e > 0); (4.4.37)
x
0
=
hk
y
M
c
_
=
k
y

0
2B
0
_
and the ansatz (4.4.38)
= exp(i[k
y
y +k
z
z]) f (x) (4.4.39)
where
c
is the cyclotron frequency, the equation of a one-dimensional harmonic oscillator,

h
2
2M

2
f (x)
x
2
+
M
c
2
(x x
0
)
2
f (x) =
_
[[
h
2
k
2
z
2M
_
f (x) E f (x), (4.4.40)
c _ Walther-Meiner-Institut
States with internal ux SUPERCONDUCTIVITY 67
is found which is entirely equivalent to nding the Landau levels of an electron in a normal metal. x
0
is
the equilibrium position, and
c
is the eigenfrequency. The harmonic oscillator has the spectrum
E =
_
n+
1
2
_
h
c
(4.4.41)
where n is an integer. The nal step is to nd the largest eld B
0
for which Eq. (4.4.41) has a solution.
With the usual substitutions for M and Q, and E given in Eq. (4.4.40) one obtains
_
n+
1
2
_
h
eB
0
m
+
h
2
k
2
z
4m
=
0
(T
c
T). (4.4.42)
The eld becomes maximal for n = 0 and k
z
= 0, hence
B
c2
=
4m
0
(T
c
T)
h
2
h
2e
(4.4.43)
=

0
2
2
GL
(T)
. (4.4.44)
The second equation shows immediately that the vortex lines are essentially at a distance of
GL
(T)
before superconductivity collapses. As a consequence, for 1, the eld and the Cooper pair density
a
2
oscillate weakly on similar length scales. For large the eld penetrates the superconductor almost
homogeneously and a
2
oscillates between zero in the vortex core and a value much smaller than a
2
0
. The
temperature dependence of B
c2
is linear close to T
c
, as can be seen directly from Eq. (4.4.43), hence
is consistent with Eq. (2.1.6). To make it consistent with the BCS prediction the prefactors become
important. This problem has been addressed by Gorkov and will be discussed in a later version of the
lecture notes.
In a nite eld B < B
c2
there is no phase transition at T
c
but only at a lower temperature T
c
(B) which can
be obtained by inverting Eq. (4.4.43),
T
c
(B) = T
c
(0)
_
12
(0)
GL
B

0
_
, (4.4.45)
where
(0)
GL
= h
2
/4m
0
T
c
(0) was used. Not unexpectedly, the transition temperature into the supercon-
ducting state decreases linearly with increasing eld. It should be noted, however, that some care should
be taken again upon extrapolating the result to low temperatures and identify
0
/2
(0)
GL
with the critical
eld at zero temperature. In other words, mind the range of validity!
4.4.2 The nucleation eld B
c3
on the surface
So far only an innite superconductor was considered. If there is a surface, for instance at x = 0, the
boundary conditions (4.2.18) or (4.2.19) have to be taken into account depending on whether the su-
perconductor is either in vacuum or covered by an insulator or, respectively, by a metal. In the case of
vacuum, which is usually the case for a typical experimental situation, (or in the case of an insulator) the
gauge invariant current across the surface must vanish,
_
h
i

x
2eA(x)
_
(x)

x=0
= 0. (4.4.46)
2013
68 R. HACKL AND D. EINZEL Ginzburg-Landau Theory
In the above paragraph we found that the maximal eld corresponds to vanishing momentum along the
z-axis, k
z
= 0, and the lowest level of the harmonic oscillator, n = 0. The corresponding eigenfunction
reads
f (x) =C exp
_

_
x x
0

2
GL
_
2
_
, (4.4.47)
which, however, satises the boundary condition only for x
0
=0 and x
0
= giving the eigenvalue above.
The question is as to whether or not eigenfunctions and -values can be found in the intervall 0 < x
0
<
for which the eld is different from B
c2
. Saint-James and de Gennes [?] found a solution in 1963 and
showed that for x
0
= 0.59
GL
0.59
heB
0
2m
=
0
(T
c
T) (4.4.48)
replaces Eq. (4.4.42) (in the case of k
z
=0 and n =0) leading to the surface critical eld B
c3
=1.695B
c2
.
The exact solution is complicated and requires numerics. In particular, the eigenfunction must have a
vanishing derivative at x = 0 which is not the case for Eq. (4.4.47) at x
0
,= 0. Kittel and de Gennes
suggested variational approaches using analytical functions which are described briey in the textbooks
by Tinkham and de Gennes. Fossheim and Sudb give a quite detailed derivation.
Physically speaking the enhanced surface eld corresponds to an enhanced Cooper pair density at the
surface being enforced by the boundary condition Eq.(4.4.46). Without the boundary condition the order
parameter would have a negative slope at x = 0 whenever x
0
> 0 while it is pushed up otherwise. The
boundary condition can be satised with any function which is symmetric about x = 0. The simplest
ground-state wave function in agreement with this requirement is a Gaussian centered at x = 0. For
x
0
> 0 symmetry can be imposed by a superposition of two Gaussians at x
0
or, simpler, by a trial
wave function f (x) = (1 +ax
2
)exp(bx
2
). If the potential energy in Eq. (4.4.40) is also symmetrized
about x = 0 one arrives at a double well potential. For appropriate values of x
0
the maximum at x = 0
is low and the potential is wide enough to facilitate a lower eigenvalue. The modulus of corresponding
eigenfunction has then a small depression at x =0 but maxima at x x
0
just as the trial wave function for
a < 0. This symmetry considerations make the reasoning behind the surface effects more plausible and
show how boundary conditions can be imposed. Needless to say that all parts for x < 0 have no physical
meaning and are constructed in a similar fashion as mirror charges for instance.
The existence of the surface critical eld B
c3
shows directly that the transition at B
c2
cannot be sharp
under realistic conditions but, rather, must be a crossover. In particular, if the surface is rough on the
order of the coherence length the maximal value of B
c3
is not reached and, in addition, vortices are
pinned to the surface making affairs even more complicated. In other words, for determining the range
of superconductivity in the presence of magnetic elds and, similarly important, currents, which we
have disregarded here for the sake of simplicity, become very important. Hence, for characterizing
superconductors and determining their intrinsic properties the conditions of the experiments have to
be controlled. What we discuss here are the intrinsic properties of clean materials in the limit of full
reversibility in order to understand the thermodynamic and microscopic properties at the origin of the
condensed state. Deviations from the described behavior are equally important since they are the basis
of applications and will be discussed there.
4.4.3 The thermodynamic critical eld B
c
While the thermodynamic critical eld is intuitively clear in a type I superconductor the denition is less
obvious when the ux can penetrate. However, the GL approach provides a direct interpretation since
c _ Walther-Meiner-Institut
States with internal ux SUPERCONDUCTIVITY 69
the difference between the free energy densities in the superconducting and the normal state is given by
f
s
f
n
=

2
2
=
B
2
c
2
0
. (4.4.49)
Recalling the expressions for
GL
and
GL
in Eqs. (4.3.29) and (4.3.27) we nd that the ratio [[/ can
be derived from
GL
while
GL
depends on [[ alone and obtain
[[ =
h
2
4m
2
GL
(4.4.50)
[[

=
m
2
0

2
GL
e
2
(4.4.51)
B
2
c
=
h
2
2(2e)
2

2
GL

2
GL
=

2
0
2(2)
2

2
GL

2
GL
B
c
=

0

22
GL

GL
. (4.4.52)
Along with the denition of in Eq. (4.3.31) and the expression for the upper critical eld in Eq. (4.4.44)
we nd,
B
c2
=

2B
c
, (4.4.53)
immediately proving the dichotomy spelled out in Eqs. (4.3.34) and (4.3.35): Whenever

2 < 1 the
nucleation eld B
c2
is smaller than the condensation eld making it energetically unfavorable for the ux
to penetrate into the material. Rather the eld is completely expelled at B
c
. We may ask as to whether
B
c2
has still a physical meaning. In fact, neglecting surface effects, B
c2
is now the supercooling eld,
and (in very clean samples) an applied eld decreasing from B
0
> B
c
cannot be expelled before reaching
B
c2
< B
c
. Once the eld is expelled B
c
can be approached from below. In optimal conditions, including
nearly atomically at surfaces and the suppression of edge effects, the eld can be cranked up further
and penetrates only at B
c1
[?]. Consequently, as expected for a rst order phase transition, there is a
hysteresis. It is also clear from the discussion of thermodynamics that the integral over the magnetization
corresponds to the condensation energy.
4.4.4 The lower critical eld B
c1
In the opposite case the nucleation takes place at B
c2
> B
c
but what happens at B
c
? Unfortunately
nothing, making the determination of the condensation energy particularly complicated in all type II
superconductors with high B
c2
. For this reason there is still a debate in the case of the cuprates after
more than two decades. However, since the thermodynamical critical eld is still determined by the
integral over the magnetization there must be a compensation of the contributions to the integral at high
elds in the low-eld range in the spirit of a Maxwell construction. In other words, the eld is completely
excluded up to a critical eld B
c1
and then starts to penetrate. The rst ux line penetrates without energy
since the Gibbs potentials with and without ux should be the same in this point of the phase diagram.
Then, as a rst guess, one would expect B
2
c
B
c1
B
c2
and B
c

2B
c1
. In reality and for 1, the
core enhances the energy stored in a single vortex line by a factor of order ln + [see Eq. (2.3.100)],
and the better estimate for B
c1
is given by (see textbooks)
B
c1
=
B
c
4
2
GL
(ln +0.08) 1 (4.4.54)
yielding B
c

B
c1
B
c2
O(ln). The relation between the various bulk elds is sketched in Fig. 4.3.
2013
70 R. HACKL AND D. EINZEL Ginzburg-Landau Theory
B B
c2
B
c1
B
c
-
0
M
Figure 4.3: Magnetization of type I and type II superconductors as a function of the applied eld B for
various GL parameters (sketch). For < 1/

2 the eld is excluded completely (black). The red and


the blue curves correspond to 0.8 and 2, respectively. As indicated for 2 the shaded areas above
and below the thermodynamical eld B
c
are equal (Maxwell construction). The innite slope at B
c1
is
only observed in ideally clean samples and indicates that the rst ux line can penetrate free of energy.
4.4.5 The Abrikosov lattice (1957)
For 1 and parallel ux lines along, e.g. e
z
, it can be shown relatively easily (Tinkham) that the force
density f on one ux line from all other ux lines and a potential transport current is given by
f = j
s
e
z

0
(4.4.55)
where j
s
is the sumof all contributions to the supercurrent density at the location of the ux line, and
0
is
the ux quantum. f is repulsive for two neighboring ux lines in equilibrium. If there is a transport current
j
tr
perpendicular to the ux lines the screening currents around the vortices have components parallel
and anti-parallel to j
tr
leading to a Lorentz force on the vortices and making them move perpendicular
to j
tr
and e
z
. This movement results in a voltage drop parallel to j
tr
hence a nite resistance and is
potentially detrimental if large currents have to transported in a magnetic eld such as in a solenoid.
Only if the movement can be suppressed efciently type II superconductors will be useful for this kind
of applications. As will be discussed in chapter 6 the ux lines can be pinned in various ways, and the
type II materials with pinning (sometimes called type III superconductors) can in fact be used for high
current and high eld applications.
Here we focus on the problem of having forces on a ux line whenever the line is not in a fully symmetric
environment where all screening currents from neighboring lines cancel out by symmetry. Abrikosov
found a solution to this problem by deriving a general solution
L
for the linearized rst GL equation
which is strictly valid only in the limit B
0
B
c2
. With the eld in e
z
direction we found B
c2
by evaluating
the lowest eigenvalue of Eq. (4.4.40) at n = 0 and k
z
= 0. Hence, only k
y
has to be considered in
Eq. (4.4.39) and will be abbreviated by k in the following. Eq. (4.4.38) shows that for each k there
exists a slab in the yz-plane centered at x
0
carrying one ux quantum. Since we need a symmetric, i.e. a
periodic solution, we set
k = nq (4.4.56)
with a xed q yielding a real space periodicity y = 2/q and
x
n
=
nq
0
2B
0
(4.4.57)
showing directly that the spacing in x-direction is
x =
q
0
2B
0
(4.4.58)
c _ Walther-Meiner-Institut
States with internal ux SUPERCONDUCTIVITY 71
and that B
0
xy =
0
as to be expected. Since the functions

n
= exp(inqy)exp
_

_
x x
n

2
GL
_
2
_
(4.4.59)
are orthogonal for different n because of the factor exp(inqy) the general solution is a superposition of
all functions
n
,

L
=

n
C
n
exp(inqy)exp
_

_
x x
n

2
GL
_
2
_
(4.4.60)
with coefcients C
n
to be determined. While periodicity in y-direction is built in via Eq. (4.4.56) the
periodicity in x-direction is recovered only if the coefcients C
n
are periodic, C
n+m
=C
n
. From the dis-
cussion above a square lattice follows immediately for index-independent coefcients C
n
or, equivalently,
m = 1. For C
1
= iC
0
and m = 2 a triangular lattice is found. For deciding as to which of the possible
lattices is being realized Abrikosov found that the linearized GL equation is insufcient and observed
that the normalization independent parameter

A
=

4
L

2
L

2
(4.4.61)
should be as close to 1 as possible.
A
= 1 obviously holds for constant
L
corresponding to the eld
free case. All other variations of
L
(dictated by ux penetrating the material) lead to larger values of
A
.
For the square lattice
A
= 1.18 for the triangular lattice
A
= 1.16. Hence the triangular lattice wins by
a very small margin. Since numerical calculations are necessary it is not that surprising that Abrikosov,
in his original paper in 1957, found the square lattice to prevail. Kleiner, Roth, and Autler showed in
1964 that the triangular lattice out of all periodic solutions has in fact the optimal value of
A
. Tinkham
notices that, for the same density of ux lines, the distance a
i
between the ux lines is slightly larger in
the triangular lattice (closest packing) than in the square lattice,
a
triangle
=
_
4
3
_1
4
1.075a
square
(4.4.62)
favoring the rst one since it reduces the positive repulsive energy between the lines. The ux line lattice
has been visualized rst by Essmann and Tr auble in 1967 by decoration and, more recently, by scanning
techniques as shown in Fig. 4.4.
The difference between the possible periodic solutions is in fact sufciently small to allow solutions other
than triangular in many cases were either the underlying lattice or the symmetry of the order parameter
introduce additional anisotropies. For instance in Nb one nds various transitions between different ux
line congurations as a function of temperature and eld [?]. In YBa
2
Cu
3
O
7
the d
x
2
y
2 symmetry of the
order parameter is sufcient to make the square lattice more favorable [?].
2013
72 R. HACKL AND D. EINZEL Ginzburg-Landau Theory
Abricosov vortex lattice
(
(

|
|
.
|

\
| u
=

2
2
0
2
) ( 2
1
exp q
B
n
x
T
e C
c
n
inqy
n

(a) (b) (c)


Figure 4.4: Flux line lattice. (a) and (b) show the different lling for the square (a) and the triangular (b)
ux line lattice. From Tinkham [?]. (c) Scanning tunneling map of NbSe
2
in the Shubnikov phase [23].
The dark areas correspond to the vortex cores where the superconducting gap and, hence, the order
parameter vanishes.
c _ Walther-Meiner-Institut
Chapter 5
The Josephson Effect
Elementary optics and quantum mechanics show that waves can tunnel through regions of space even
in the case of a potential barrier. For instance, light can travel across a sufciently narrow gap between
two dielectrica even if the incoming wave is totally reected from the inner surface of one of the ma-
terials. Another example are electrons tunneling through an insulator sandwiched between two metals.
In all cases the wavelength and the height of the barrier govern the transmitted intensity. Typically,
for distances larger than a few wavelengths the transmitted intensity vanishes. What happens in these
circumstances with the Cooper pairs of a condensate, and is it worthwhile to bother?
If we could couple two condensates weakly it appears very intriguing to utilize the rigid phases on either
side of the junction so as to generate beat frequencies in the MHz or GHz range having unprecedented
stability and watch quantum mechanics at work on a macroscopic scale. When Josephson considered this
possibility in 1962 [?] the experts were skeptical. John Bardeen for instance argued that the tunneling
probability would be orders of magnitude too small for being the square of the tunneling efciency of
a single electron. It turned out that the rules for condensates are different and that the single particle
probability still determines the order of magnitude of the matrix element.
5.1 Weakly coupled superconductors
The Ginzburg-Landau model gives us an idea of how we should think of pair tunneling at least in the case
of a metallic interface between two superconductors. Eq. (4.2.19) explicitly describes how far the wave
function extends into the normal metal via the proximity effect, and Eq. (4.3.29) shows
GL
(T) to be
the approximate length scale. A schematic view of the experimental setup for observing the Josephson
effects is shown in Fig. 5.1. The two superconductors which may be made of different materials are
separated by a thin insulating or metallic layer (grey). Other realizations of weak links are either narrow
superconducting bridges with lateral dimensions close to
GL

0
or boundaries between differently
oriented single crystals or intrinsic contacts in natural very anisotropic materials [?] or hetero structures.
Most importantly, the critical supercurrent density across the weak link must be orders of magnitude
smaller than in the bulk and coherence between the two condensates must be maintained. The junctions
can be either voltage or current biased. In most of the cases the current I is determined from out side and
the voltage U is the dependent quantity. Nevertheless, the current-voltage characteristics will be plotted
the other way around partially for traditional reasons. However, since the energy is given eU this is also
more instructive as we will see below.
73
74 R. HACKL AND D. EINZEL The Josephson Effect
superconductor superconductor
insulator/
metal
~
0
U
I
Figure 5.1: Schematics of a weak link. The grey slab separating the superconductors on the left and
the right can be either an insulator or a semi conductor or a metal. If the cross section is reduced to
dimensions of the coherence length
0
even a superconductor can be used a weak link. The thickness of
the layer should be on the order of
0
or less. Usually the contact is current biased and the voltage drop
U is the dependent quantity.
5.2 The Josephson equations
The dynamics of the Cooper pairs can be described by time-dependent Schr odinger equations for the
condensate wave functions
1,2
(r, t) for superconductors 1 and 2 having stationary energy eigenvalues
E
1,2
. Then the most physical approach to a weak link is to introduce an interaction W
12
E
1,2
between
the the two superconductors leading to an equation system of two coupled harmonic oscillators [?, ?],
i h
1
(r, t) = E
1

1
(r, t) +W
12

2
(r, t)
i h
2
(r, t) = E
2

1
(r, t) +W
21

1
(r, t),
where is the partial derivative of w.r.t. time. For solving these equations we use again the Madelung
representation of the wave function [Eq. (2.3.65)]. In contrast to the chapter on GL theory we substitute
here a
2
1,2
(r, t) =Qn
pair
(r, t)
Q
(r, t) for the amplitude. After separating imaginary and real part we get,
respectively,
a
2
1
=
2W
12
a
1
a
2
h
sin(
2

1
) = a
2
2
and (5.2.1)
ha
2
1

1
= E
1
a
2
1
+W
12
a
1
a
2
cos(
2

1
) (5.2.2)
ha
2
2

2
= E
2
a
2
2
W
21
a
1
a
2
cos(
1

2
). (5.2.3)
(5.2.4)
From the rst equation we obtain that a
2
1
= a
2
2
meaning that the rate of change of the charge density on
the left side is equal but opposite to that on the right side of the junction. In addition, the time derivative
of the density is related to a current density by virtue of the continuity equation. However, it is clear that
an imbalance of charges is not possible. Hence, so long as no current is supplied from outside there wont
be a current across the junction, and the phase difference vanishes. In turn, if an external current from a
source is supplied Cooper pairs may move from one side to the other side, and
2

1
will be different
from zero. The maximal current is proportional to the matrix element W
12
W
21
and is expected to be
small. These considerations allow us to write down the rst Josephson equation,
j = j
c
sin(
2

1
) j
c
sin(), (5.2.5)
c _ Walther-Meiner-Institut
The Josephson equations SUPERCONDUCTIVITY 75
where j
c
is the critical Josephson current density, which states that there is a nite supercurrent across,
e.g., an insulator without a voltage drop. The current is proportional to the sine of the relative phase of
the condensates in superconductor 1 and 2 or directly to for small currents.
An estimate of the critical current can be obtained from GL theory or from microscopic considerations.
The latter was achieved by Ambegaokar and Baratoff [?] who found the full temperature dependence of
the product of the critical current I
c
and the normal resistance R
n
of the junction which can be approxi-
mated as R
n
(T
c
),
I
c
R
n
=

2e
tanh
_

2k
B
T
_
. (5.2.6)
This important result is universal, i.e. independent of the type of junction. In the limit T 0
it reduces to I
c
R
n
= (0)/2e while close to T
c
the temperature dependence becomes linear and
I
c
R
n
= 2.34k
B
/e(T
c
T) with the constant being 635V/K.
The phase difference is not gauge invariant implying that the current would depend on the selected
gauge of the vector potential A in an applied magnetic eld. Therefore, before proceeding, we introduce
the gauge-invariant phase difference ,
=
2

0
_
2
1
A ds, (5.2.7)
where the integration limits correspond to the two sides of the weak link. Applying the gauge transfor-
mations A A+ and +
2e
h
immediately proves Eq. (5.2.7). is generally valid with and
without eld and has to used later when we derive properties of junctions in a eld and study quantum
interference effects. The gauge invariant rst Josephson equation the reads
j = j
c
sin. (5.2.8)
For the analysis of Eqs. (5.2.2) and (5.2.3) we assume the same material on each side and get a
1
= a
2
and W
12
=W
21
. Since the cosine is an even function, the sum of the two expressions yields the second
Josephson equation,
d
dt
=
2e
h
U, (5.2.9)
where U =E
1
E
2
and Q=2e. Obviously, upon exceeding j
c
there is a nite voltage across the junction,
and the phase difference becomes time dependent. The differential equation can be integrated right away
yielding
(t) = (0) +
_
t
0
2

0
Udt
/
= (0) +2
U

0
t. (5.2.10)
If we insert (t) into the rst Josephson equation we obtain a current
j(t) = j
c
sin
_
(0) +2
U

0
t
_
. (5.2.11)
that oscillates at the frequency U/
0
where
0
is the ux quantum. The inverse of the ux quantum is
the celebrated Josephson frequency
1

0
= 483.597870(11)
MHz
V
, (5.2.12)
meaning that for a voltage drop of 1V a frequency of 483.597870(11) MHz can be observed that de-
pends only on constants of nature. Along with the resistance normal given by the von-Klitzing constant
h/e
2
=R
K
=25812.8074434(84)the Josephson frequency plays an important role in metrology since
the voltage can be linked to elementary constant.
2013
76 R. HACKL AND D. EINZEL The Josephson Effect
5.3 The RCSJ model
While the fundamental properties are easily derived the details important for the design and understand-
ing of sensors and other devices require the study of individual junctions. For that we note that a weak
link is not only a highly non-linear resistance but has an Ohmic and a capacitive shunt which are of cru-
cial importance for the dynamics. Therefore we analyze the properties of the device sketched in Fig. 5.2.
The bias current I
b
is distributed between the three branches across the resistor, the weak link, and the ca-
Figure 5.2: Equivalent circuit of a resistively and capacitively shunted junction (RCSJ). (a) The weak
link is symbolized by a cross. The total bias current I
b
is distributed between the three branches as
I
R
= U/R, I
J
= I
c
sin, and I
C
= C

U. (b) and (c) The dynamics of the RSCJ model can be visualized
by a washboard potential where U() = E
J
cos I
b
/I
c
where is the coordinate. (b) If the
I
b
< T
c
the phase is time independent and a supercurrent ows across the junction. (c) Above the critical
current becomes time dependent but does not vary linearly as expected for a harmonic oscillator.
From Clarke.
pacity as I
R
=U/R, I
J
=I
c
sin, and I
C
=C

U. We use the second Josephson equation (5.2.9) to eliminate
U and

U and get
I
b
I
c
= sin +

0
2I
c
R
+

0
C
2I
c
. (5.3.13)
Eq. (5.3.13) is equivalent to the second order differential equation of a driven pendulum. The term
I
c
sin replaces the linear term of the harmonic oscillator equation with small amplitude and allows for
full rotations. The mechanical analogue reads
D = mgl sin + + (5.3.14)
where D is the driving torque, mgl sin is the gravitational energy if the pendulum is rotated out of
equilibrium, is the damping, and is the moment of inertia. In the case of the RCSJ model it is
customary (and convenient) to introduce the parameters

c
=

0
2I
c
R
, (5.3.15)

c
=
2I
c
R
2
C

0
Stewart McCumber parameter (5.3.16)
=
RC


RC

c
Q
2
quality factor,
=
t

c
, and dt =
c
d, (5.3.17)
c _ Walther-Meiner-Institut
The RCSJ model SUPERCONDUCTIVITY 77
and obtain the dimensionless sine-Gordon equation
1
i = sin +

+
c

2
. (5.3.18)
The Stewart-McCumber parameter
c
corresponds to a quality factor and indicates how freely the system
can oscillate. However, independent of
c
the time dependences of i() and () will never be harmonic.
Rather, as we will see below,
c
indicates as to whether or not the I V characteristics is hysteretic.
For
c
= 0 we nd
i = sin +

(5.3.19)
which can be integrated for i > 0 over one period allowing us to derive an average voltage from (5.2.9).
Separation of the variables and integration between 0 and 2 and, respectively, 0 and the period duration

0
yields
2

i
2
1
=
0
(5.3.20)
and after restoring units and inserting the average oscillation frequency 2/T =2(
c

0
)
1
in Eq. (5.2.9)
we nd
U = R
_
I
2
b
I
2
c
. (5.3.21)
This result shows that the voltage is zero for I
b
= I
c
and then asymptotically approaches Ohmic behavior
for I
b
I
c
. For I
b
< I
c
integral (5.3.20) diverges implying that the phase does not depend on time and
U vanishes.
It is not possible to solve Eq. (5.3.18) analytically for
c
> 0, and the I V characteristics can only be
obtained numerically. For
c
> 0.5 the I V curves become increasingly hysteretic as shown in Fig. 5.3.
This means that the voltage assumes a nite value immediately at I
c
jumping directly to the Ohmic line
for
c
1. On reducing I
b
the voltage follows Ohms law down to I
<
where U nally jumps back to
zero and the current ows free of losses. If we integrate over one period we nd that
R. Kleiner
Dissertation TUM
Figure 5.3: I V curves for differently damped Josephson contacts. The numerical simulation is accord-
ing to the model of a resistively and capacitively shunted junction (RCSJ) for
c
as indicated. From R.
Kleiner (PhD thesis, TUM 1092).
1
The equation is the non-linear analogue of the one-dimensional Klein-Gordon equation, being the relativistic form of the
Schr odinger equation. Goldstein, Poole, and Safko (Addison Wesley 2002) speculate that the name is a frivolous pun.
2013
78 R. HACKL AND D. EINZEL The Josephson Effect
_
T
0
Udt =
0
(5.3.22)
meaning that the phase changes exactly by one ux quantum per cycle. The energy is radiated off the
junction with a frequency on the order of the Josephson frequeny times the voltage.
5.4 Josephson contact in a microwave eld
We go now the other way around and irradiate the junction with a microwave eld U
HF
= U
0
+
U
1
cos(
1
t) with U
1
U
0
. Then the phase varies as
(t) =
0
+
2e
h
U
0
t +
2e
h
U
1

1
sin(
1
t) (5.4.23)
which has to be inserted into Eq. (5.2.8). The general solution is given by Bessel functions of order n, J
n
as
j(t) = j
c

n=0
J
n
_
2eU
1
h
1
_
sin
_
2eU
0
h
t n
1
t +
0
_
. (5.4.24)
We search now for the dc contributions to j(t) and observe that the time integral over the sum is different
from zero only if

1
=
2e
h
U
0
(5.4.25)
and integer multiples thereof. Hence the distance on the voltage axis U
n
between different steps is an
integer multiple of U
0
,
U
n
=
n h
2e
= nU
0
(5.4.26)
The way the calculation works can be visualized relatively easily for the rst step (see Problem 2 in set
10).
5.5 Josephson effect in a magnetic eld
In a magnetic eld the gauge invariant phase difference (5.2.7) across the junction becomes relevant as
one may anticipate already from the supercurrent (2.3.74).
5.5.1 Ring with a single weak link
We consider now the setup sketched in Fig. (5.4). There is a ring having a central hole. The total diameter
of the ring is at least an order of magnitude larger than the penetration depth
eff
of the magnetic eld
B =A implying that a contour can be found along which the screening currents vanish. The eld
can penetrate the weak link between points 1 and 2 and the superconductor above and below up to
eff
.
The distance of 1 and 2 from the weak link is larger than
eff
. Then Eq. (2.3.74) yields
_
1
2
ds =
2e
h
_
1
2
A ds
1

2
=
2e
h
_
1
2
A ds (5.5.27)
c _ Walther-Meiner-Institut
Josephson effect in a magnetic eld SUPERCONDUCTIVITY 79
J
s
= 0
1
2
(ds)
Figure 5.4: Ring with a weak link in a magnetic eld. The eld B =A points out of the plane. The
ring walls are much thicker than the penetration depth
eff
of the eld. Therefore the current J
s
along
the contour (integration variable ds) vanishes. The eld penetrates the weak link and the adjacent
superconductor as long as the distance from the weak link is smaller than
eff
. Points 1 and 2 are in the
eld free part.
and

1
= +
2e
h
_
2
1
A ds. (5.5.28)
Upon adding Eq. (5.5.27) and (5.5.28) one obtains
2n = +
2e
h
_

A ds
= +
2e
h
_
B dS (5.5.29)
where S is the total are inclosed by . Since the phase is unique only modulo 2n the factor 2n, where
n is an integer, has to be added. The last integral is the magnetic ux through the ring. If we insert the
ux quantum
0
we nd that the gauge invariant phase
= 2
_
n

0
_
(5.5.30)
changes proportional to the ux through the ring. The rate of change is given by
0
which is, as we
know, a very small number. Hence, a ring as shown in Fig. (5.4) can in principle be a very sensitive
sensor if the can be measured.
5.5.2 Ring with two weak links: Quantum interference
To this end we consider a ring having two weak links as shown in Fig. 5.5 (a). With this setup the total
ux penetrating the ring is proportional to the phase differences
1
and
2
of links 1 and 2,

2
= 2

0
. (5.5.31)
2013
80 R. HACKL AND D. EINZEL The Josephson Effect
Figure 5.5: Superconducting quantum interference device (SQUID). (a) Schematic view. (b) I V curve
as a function of the applied eld. (c) Modulation of the maximal critical current as a function of the eld.
Note that the Josephson junctions have to have a relatively small
c
so as to avoid hysteresis.
If the two contacts have exactly the same properties the total current across the two links is simply the
sum of the individual currents, I
tot
= I
c
(sin
1
+sin
2
). With the substitution

1
=
0
+

2
=
0

0
where
0
takes care of all problems with screening currents (which are a complicated issue in practice)
we nd immediately
I
max
(B) = 2I
c
sin
0

cos

. (5.5.32)
Eq. (5.5.32) states that the current is maximal for vanishing eld. This is not surprising since the two
currents add symmetrically. In all other cases there is a superposition of the bias and the screening current
which lead to a pattern of I
max
(B) which is equivalent to that of an optical interference experiment with
a double slit. Therefore the ring with two weak links is named superconducting quantum interference
device (SQUID)
2
. If
c
of the two contacts is in the right range I
max
(B) oscillates around a mean value
but does not reach zero. In addition all hysteresis effects vanish. Then the voltage in the resistive state is
a unique function of B as long as 2 <
0
. If the eld through the SQUID is compensated in the most
sensitive part of the I V curve [Fig. 5.5 (b)] one has a very sensitive eld detector (10
6

0
typically).
If not the voltage oscillates [Fig. 5.5 (c)].
5.5.3 Quantum interference in a long junction
So far we assumed that the current is constant across the entire junction. However, the width of the
superconductor perpendicular to the eld B and the current I is usually as large as several penetration
depths as shown in Fig. 5.6. Hence, the gauge invariant phase may vary substantially in the direction
perpendicular to the current and the eld. The width of the contact a is much larger than the penetration
depth , and the eld points into the positive z-direction. Note that d can be larger than in type I
superconductors. Then the ux through the area enclosed by the curve (blue) is given by
(x) = Bx(d +2) (5.5.33)
2
John Clarke who invented the SQUID in 1966 is a gourmet.
c _ Walther-Meiner-Institut
Josephson effect in a magnetic eld SUPERCONDUCTIVITY 81
yielding the gauge invariant phase,
(x) = (0) +2
(x)

0
. (5.5.34)
For getting the maximal supercurrent I
max
in y direction Eq. (5.5.34) will be inserted in the rst Josephson
0
a/2
-a/2
x
y
B
x
d

Figure 5.6: Long (wide) weak link. The width a in x-direction is much larger than the penetration depth
. The current I is along y, and the eld B points along the z-direction. For getting the ux (x) one has
to integrate in the x y plane inside the curve (blue). The screening currents j
screen
() are indicated
schematically (red). Here they are larger than the critical current across the weak link. in y direction the
sample is larger than shown.
equation (5.2.8) that is then integrated in x and z direction in the limits [a/2] and [0, c] where c is the
thickness of the sample:
I
max
= I
c
sin((0))

sin
_
(x)

0
_
(x)

. (5.5.35)
Eq. (5.5.35) is equivalent to the Fraunhofer pattern in optics describing the intensity distribution of the
diffracted light after a single slit while the SQUID corresponds to the double slit experiments. In practice
the ux dependence of the maximal current across a SQUIDis a superposition of a sin and cos modulation
since the legs of the SQUID have also a nite width. If Eq. (5.5.34) is inserted in the Eq. (5.2.8) an x-
dependent current density as shown in Fig. 5.7 is found which explains the zeros in the maximal current
physically.
If the eld is reduced it can be expelled from the junction. However, since the critical current across the
junction is very small in comparison to that in the bulk the Josephson penetration depth
J
,

J
=
_

0
2
0
j
c
d
_1
2
, (5.5.36)
is usually much larger than the London penetration depth
L
. Eq. (5.5.36) is derived from the Ferrell-
Prange equation [?]. Typical values for
J
exceed those for
L
by at least one order of magnitude.
2013
82 R. HACKL AND D. EINZEL The Josephson Effect
e
f
Magnetic field B =0 Magnetic field B =B
0
/2
Magnetic field B =B
0
Magnetic field B =3B
0
/2
-4 -3 -2 -1 0 1 2 3 4
0.0
0.5
1.0

I
m
a
x
/
I
c
/
0
Figure 5.7: Current in a long junction. (a)(d) Current distribution for different eld/ux values where
B
0
corresponds to one ux quantum. (e) Maximal current (in units of the critical current) as a function
of the ux penetrating the junction according to Eq. (5.5.35).
0
is the ux quantum. (f) Experimental
results from a Pb/YBa
2
Cu
3
O
7
junction. From [?].
c _ Walther-Meiner-Institut
Chapter 6
Unconventional Materials
The entire eld of superconductivity is driven by the discovery and development of new compounds
bringing the materials scientists into the center of the game. As opposed to semiconductor physics
it is extremely hard to predict or at least guess the tendency towards superconductivity. In addition,
any prediction of transition temperatures from rst principles, even for known materials, is difcult
and was successful only for a few elements and simple compounds with conventional electron-phonon
coupling [9]. To our knowledge BKBO is the only material having a T
c
close to 30 K which was predicted
on the basis of the isostructural low-T
c
sister compound BaPbBiO3 [?]. In all other cases empirical rules
such as those of Bernd Matthias [?] or highly complex quantum chemistry considerations have little or
no predictive power yet meaning that our understanding is still insufcient and requires new and bright
ideas. Hence, the purpose of this chapter is twofold in that the zest for understanding on the theoretical
side is highlighted and the inventions of the materials scientist are celebrated by describing some of the
fascinating properties of unconventional superconductors.
6.1 Classication
Electron-phonon coupling was considered the only mechanism leading to superconductivity for a long
time although magnetic polarizability was recognized as possible alternative [24]. However, only the
discovery of systems with heavy electrons [25] and superconductivity therein [26] opened new vistas.
1
It was pointed out from the beginning that the heavy mass (large heat capacity) may originate in the
Kondo-like interaction of the conduction electrons with the spins of the 4 f -electrons of cerium. The dis-
covery of superconductivity at approximately 0.5 K in CeCu
2
Si
2
in 1979 [26] was a new paradigm and
opened a completely new eld of research that remained active and vibrant until now [27]. The heavy
fermion systems were certainly the rst examples of materials belonging to the class of unconventional
superconductors which include now also the copper-oxygen [6] and the iron-based compounds [7] as
compiled in Table 6.1. What means unconventional? The distinction between conventional and uncon-
ventional superconductors is not clear-cut. Sometimes all electron-phonon superconductors carry the
name conventional while all others are unconventional. In the terminology of Anderson [?] dividing rule
is given by the Fermi surface average of the gap
k
: Whenever
k

k
= 0 superconductivity is uncon-
ventional. Considering the modern developments it makes also sense to ask as to whether or not there
1
The name heavy electrons (or fermions) derives from measurements of the electronic heat capacity c
p
= T with the
the Sommerfeld constant. As shown rst for CeAl
3
c
p
is approximately three orders of magnitude larger than that of a usual
metal. Since is proportional to the density of states at the Fermi energy N
F
and, hence, to the effective mass m

of the
electrons a large was associated with a large m

. Maurice Rice pointed out early that, by denition, N


F
[v
F
[
1
implying a
large to be more indicative of slow rather than heavy electrons. He added that slow electrons might sound less exciting and
important.
83
84 R. HACKL AND D. EINZEL Unconventional Materials
Table 6.1: The main catagories of superconductors. Conventional includes all metallic compounds with
electron-phonon coupling where a lattice instability driven by strong coupling is the only additional
phase. MgB
2
is listed separately for holding the record of the conventional systems due to a mixture
of intra- and interband pairing. CuO
2
represents all superconductors on copper-oxygen basis. Fe-based
systems include also those with Fe replaced by Rh. f -electron systems are typically those having a
high electronic specic heat at low temperature and are also known as heavy fermion compounds. The
following abbreviations are used: electron (el.), instability (inst.), antiferromagnetism (AF), density wave
(DW), spin density wave (SDW), ferromagnetism (FM), and cylinder (cyl.). The missing entry in the last
column reects the existence of various types of pairing states in f -electron systems. The years in
brackets indicate the rst chance for observing superconductivity (see text). References can be found in
the text.
conventional MgB
2
CuO
2
Fe-based f -electron
structural anisotropy isotropic 10. . . 100 10. . . 10
5
5. . . 100 isotropic
normal state metal metal strange metal metal heavy el.
competing phases (lattice inst.) AF, DW SDW AF/FM
Fermi surface multi-band cyl. + sphere cyl. 1 band cyl. 5 bands multi-band
T
c,max
(K) 30 39 160 56 18.5
H
c2,max
(T) 50 50. . . 100 100. . . 250 100 27+

0
(nm) 10. . . 1000 1. . . 5 3 5 3. . . 30

GL
(nm) 0.01. . . 50 100 100 50 100
superconductivity isotropic s anisotropic s d
x
2
y
2 s

year of discovery 1911 2002(1952) 1986(1978) 2006(1995) 1979


is one or more other phases competing with superconductivity. The heavy fermions, the cuprates, the
Fe-based materials, and organic metals share this property. In the vast majority of the cases magnetism
competes with superconductivity. Less frequently, charge order plays a role. Very often the competing
phase is suppressed continuously as a function of a parameter r other than temperature such as applied
pressure or magnetic eld or doping. If the transition temperature to the ordered state T
o
(r) approaches
zero for a specic r
c
the material belongs to the large class of quantum critical systems [?, 28]. Here,
the physics is dominated by the uctuations of the spin or charge density in a wide range of temperature
above r
c
which are considered to contribute substantially to superconductivity in both the cuprates and
the Fe-based compounds [29, 30]. In fact, since the electron-phonon coupling is too small to explain
transition temperatures in the range of 100 K other types of exchange bosons have to take over at least so
long as a the interaction is retarded (meaning that the timescale is much longer than that of the electronic
screening).
Although the denitions are not generally accepted and/or used one has at least an intuitive idea along
which lines the divide runs. Nevertheless, when reading the literature one should be aware of the slightly
fuzzy denitions. For the present purposes it is enough to know that there are systems having properties
fundamentally different from those of the usual metallic superconductors. In most of the cases they are
characterized by rich phase diagrams and the proximity of magnetism and superconductivity. In the
following we provide an overview of the Fe-based and copper-oxygen systems and a brief summary of
the heavy fermions and the organic metals.
c _ Walther-Meiner-Institut
The iron-age of superconductivity SUPERCONDUCTIVITY 85
6.2 The iron-age of superconductivity
High-temperature superconductivity in compounds with iron
2
was not considered seriously before
Yoichi Kamihara and coauthors in Hideo Hosonos group discovered a transition at T
c
= 26 K in
La(O
1x
F
x
)FeAs. [31] It is interesting to note that LaFePO having T
c
= 6 K was synthesized already
in 1995 by Barbara Zimmer in the group of Wolfgang Jeitschko
3
. The observation of superconductivity
was only mentioned in Zimmers thesis. Since the publication of Kamiharas results in 2008 several
ten thousand papers have appeared, and the highest T
c
so far exceeds 50 K. Even though the materials
contain arsenic or other rather toxic elements many laboratories started with the preparation of poly-
and single-crystalline samples. Generally the Fe atoms form two-dimensional layers and are coordinated
with atoms from the fth column in most of the cases (with Se being an exception) as shown in Fig. 6.1.
The family name pnictides (Pn) comes from Greek i o (asphyxiant, suffocative) for the column
of nitrogen.
In particular Chinese scientists contributed a lot of important results and found Nd(O
1x
F
x
)FeAs with
the so far highest transition at 55 K [?]. While in the beginning many people believed the iron pnic-
tides to be another class of oxides Dirk Johrendt and his group [8] demonstrated that high transition
temperatures can also be obtained in purely intermetallic compounds. At optimal doping with x 0.4
Ba
1x
K
x
Fe
2
As
2
reaches a T
c
of 38 K. In contrast to the oxiuorides (see Fig. 6.1) large single crystals
can be grown although the homogeneity and quality is not in all cases satisfactory. However, crystals
of BaFe
2
(As
1x
P
x
)
2
and LaFePO are already sufciently clean for the observation of quantum oscilla-
tions [32, 33].
Figure 6.1: Structures of iron pnictide compounds (by courtesy of D. Johrendt). LaFeAsO
1x
F
x
(1111;
T
max
c
= 28 K) was the rst pnictide superconductor with high T
c
[31]. With Pr, Nd or Sm replacing
La T
c
exceeds 50 K [?]. LaFeAsO
1x
F
x
is isostructural to LaFePO having the transition at 6 K at the
stoichiometric composition. BaFe
2
As
2
(122) develops a spin-density wave (SDW). When doped with K
for Ba [8], Co for Fe [34] or P for As the SDW is suppressed and superconductivity appears. LiFeAs
(111) has a maximal T
c
of 18 K. The simplest of the materials, FeSe (11), with the same structural
elements but Se for As has a maximal T
c
of 8 K with 9% Se deciency at ambient pressure and reaches
27 K at 1.5 GPa [35, 36].
The question as to the origin of superconductivity arose immediately. Are the iron pnictides similar to
the cuprates or to MgB
2
with T
c
= 39 K due to electron-phonon coupling or are they a material class on
their own? The spin-density-wave (SDW) order of the parent compound indeed suggests a proximity to
the cuprates, where the superconducting phase emerges from a Mott insulator. With doping p away from
2
This section is partially copied from a contribution of one of us (R.H.) to the Annual Report 2009 of the WMI. The title is
borrowed from a Viewpoint by Michelle Johannes in Physics 1, 28 (2008).
3
Note that the undoped parent compounds of both the cuprates and the pnictides were well known quite some time ahead of
the rst observation of superconductivity in doped variants.
2013
86 R. HACKL AND D. EINZEL Unconventional Materials
half lling spin and charge uctuations as well as superconductivity follow antiferromagnetic long range
order. However, in contrast to the cuprates there is no universal phase diagram in the pnictides (Fig. 6.2).
In the essentially hole doped oxiuorides there is an abrupt transition from a magnetically ordered phase
to a superconducting one with T
c
only weakly depending on doping. The electron-doped intermetallic
compounds have a smooth transition, and SDW order and superconductivity may even coexist [37]. The
phase boundary of superconductivity is dome shaped.
J une 16, 2009; 4
BaFe
2
As
2
Chu et al. PRB 79, 014506 (2009)
Mandrus, Canfield, Bchner, Klauss, Dai,.
Ba(Fe
1-x
Co
x
)
2
As
2
Figure 6.2: Phase diagrams of LaFeAsO
1x
Fx (left) [38] and Ba(Fe
1x
Co
x
)
2
As
2
(right) [37]. While
SDW und superconductivity (SC) overlap in 122 there is a strict separation in La-1111. The structural
transition at T

always precedes spin density wave order at T

.
The differences in the phase diagrams of the pnictides are surprising since the electronic structures are
remarkably similar. There are 5 bands derived from the Fe 3d orbitals. Two (
1,2
) form concentric hole-
like Fermi cylinders around the center of the Brillouin zone (BZ), two (
1,2
) have FSs which encircle the
corner of the small BZ derived from the 2Fe crystallographic unit cell (Fig. 6.3). Since the cross sections
of the resulting Fermi surfaces are nearly equal the and sheets are nested with the vector Q(, ).
Consequently, the electronic susceptibility as described by the Lindhard function is strongly momentum
dependent [30] and is believed to be at the origin of the SDW. The pronounced peaks in the susceptibility
make the strong variations of the properties upon small changes of the electronic and lattice structures at
least plausible.
The real part of the susceptibility is also considered a possible origin of superconductivity [30, 3941]
while the electron-phonon coupling is probably weak [42]. From this point of view, the pnictides and the
cuprates appear to be cousins in the same family even if the strong metallicity of the parent phases of the
FePn compounds may argue otherwise. But how can the coordinates of the pnictides be determined? In
a recent optical transport study the authors conclude from the reduced band width that the pnictides are
half way between normal metals and the cuprates [43]. However, the related spectral redistribution to be
expected upon doping is not observed by angle-resolved photoemission (ARPES) and x-ray absorption
(XAS) [44, 45]. Perhaps one of the most telling similarities would be if the pnictides had the signature
property of all cuprates an energy gap
k
having nodes and a sign change along the Fermi surface [19].
In the 6 years after the discovery the understanding and the experimental knowledge has advanced sub-
stantially. The results from the usual spectroscopic methods such as tunneling and photoemission start
to converge. In most of the compounds the modulation of the energy on the Fermi surface is not sup-
portive of gap nodes on individual Fermi surfaces [47]. However, it becomes more likely that the sign
change between the hole and the electron surfaces leaves imprints in the tunneling spectra in an applied
magnetic eld [?] supporting the early suggestion of Mazin and coworkers [30] of an s

gap. Here the


phases of the gaps on the hole- and electron-like Fermi surfaces differ by . This would imply that, for
c _ Walther-Meiner-Institut
The iron-age of superconductivity SUPERCONDUCTIVITY 87
Generic band structure and Fermi surfaces
1Fe

k
y
1Fe
2F

2Fe
M

X
Q
k
x

SPP 1458- Iron Pnictides page 14 November 30, 2009


Figure 6.3: Iron plane (left), Brillouin zone (BZ), and real part of the non-interacting susceptibility

0
(q, ) [30] (right) of FePn materials. The cell relevant for the electronic structure contains 1 Fe atom
(dashes) and is smaller by a factor of 2 and rotated by 45

with respect to the crystal cell (full line and


axes a and b). The BZ of the unit cell (full line) and the rst quadrant of the Fe plane (dashed line) are
shown along with the FS cross sections at /c =0 (adopted from Ref. [46]). The dotted FSs are obtained
by downfolding the 1 Fe BZ. Even with the ellipsoidal elongation of the M barrels the and bands are
approximately nested. Re
0
(q, ) controls the pairing strength [30].
ideal conditions of equal cross sections of the respective Fermi surfaces and magnitudes of the gaps, the
Josephson current in c-direction perpendicular to the Fe planes would vanish for a conventional counter
electrode. While this Gedankenexperiment is good for visualizing the consequences of the s

gap (and
also other gaps) the technical realization is hampered by the real shape of the Fermi surfaces and the
remaining variation of the gaps. Here the coherence factors come into play.
As was shown in problem 2 of set 5 external perturbations having the potentials or A are either in-
dependent of the particle momentum or proportional to the momentum or its square. This and the spin
introduce a sensitivity to the sign of k in the response since the Bogoliubov particles in a superconductor
are coherent superpositions of electrons and holes and the amplitudes of the perturbation matrix elements
have to be added before being squared. (The problem is outlined in the books of Tinkham and de Gennes
and will be discussed in detail in a later version of chapter 3 of this manuscript.) The bottom line is that
two particle response functions given by a superposition of occupied and unoccupied states such as the
spin susceptibility, the ultrasound absorption, the NMR properties, the optical conductivity or the light
scattering response assume characteristic energy and temperature dependences depending on the related
perturbation operator. In the case of the spin susceptibility and of the longitudinal ultrasound absorption
there is no momentum involved. For NMR and optical absorption (one dipole transition) k enters lin-
early via k A and for light scattering the perturbation is proportional to k
2
(two dipole transitions) via
A
2
. Since the full momentum dependence of the gap enters the coherence factors the different types of
responses assume energy and temperature dependences characteristic for individual
k
.
In the Fe-based superconductors a resonance in the spin correlation has been observed by neutron scatter-
ing [48] which may in fact be indicative of gaps having opposite sign on the electron and hole bands. The
NMR and optical experiments are not conclusive yet. Also the angle-resolved photoemission (ARPES)
is more complicated than in the cuprates due to the relatively strong electronic k
z
dispersion following
from the weak anisotropy (see Table 6.1). In fact ARPES and Andreev tunneling favor large, essentially
constant gaps on all Fermi surfaces [49, 50]. In some compounds such as Ba
1x
K
x
Fe
2
As
2
the modu-
lation of the gap is stronger along k
z
than in the k
x
k
y
-plane [?]. Nevertheless, some of the materials
have very small gaps on parts of the Fermi surface in addition to the possible sign change [47, 51, 52].
Some experiments indicate a strong modulation of
k
and even true nodal behavior [53, 54]. The strong
material dependence inferred from the experiments is also expected from theoretical consideration that
indicate the close proximity of various ground states [?].
2013
88 R. HACKL AND D. EINZEL Unconventional Materials
The possible anisotropies in the superconducting state are accompanied by band dependent carrier dy-
namics in the normal state as observed by quantum oscillatory phenomena [32, 33] and the analysis of
Hall data [55]. For the clarication of these fundamental questions at the heart of the physics of the
pnictides bulk sensitive spectroscopies with band and momentum resolution will be instrumental.
Recently, the critical current density was determined in magnetic elds up to 45 T and found to be in
the range of 10
4
Acm
2
up to at least 30 T (Fig. 6.4). These properties open the possibility to use
Ba(Fe
1x
Co
x
)
2
As
2
for static elds in the range above 21 T the limiting value of Nb
3
Sn. It became
Figure 6.4: Critical currents j
c
and pinning forces of Ba(Fe
1x
Co
x
)
2
As
2
in a magnetic eld. The trans-
port current is in the a b plane. (a) The eld B that points in z-direction. (b) The eld B is in the
ab-plane. From Ref. [?].
apparent relatively early that the pnictides are much harder a problem to solve than, e.g., MgB
2
. In fact,
electron-spin or direct electron-electron interactions moved into the main focus of research. Therefore,
the people working on the CuO
2
compounds were naturally attracted. The hope is that the results in
the pnictides pave the way also towards a better understanding of the cuprates and of high-temperature
superconductivity in general.
6.3 Copper-oxygen compounds
When superconductivity was discovered in LaBaCuO [6] in 1986 and the proper phase La
2x
Ba
x
CuO
4
isolated soon thereafter [?] an unprecedented goldrush started. Now almost 3 decades and some 300,000
publications later the puzzle is still among the most important issues of solid state research but the knowl-
edge of superconductivity, competing phases, strange metals, Mott and charge transfer insulators and the
theoretical description thereof has grown tremendously. In addition, several experimental methods were
pushed close to perfection and opened completely new insights. Here, we just give a brief summary of
what has been revealed in this still exciting eld.
4
4
The section is partially copied from the article in Zeitschrift f ur Kristallographie published by one of us (R.H.) [56].
c _ Walther-Meiner-Institut
Copper-oxygen compounds SUPERCONDUCTIVITY 89
6.3.1 History
It is enlightening to look at the activities preceding the nal successful synthesis of Ba
x
La
5x
Cu
5
O
5(3y)
[6]. As in many other cases there was a program where to search for higher transition temperatures. The
famous empirical rules of Bernd Matthias, who succeeded to discover more than 1000 new supercon-
ducting materials, are probably the best documented example. They include, among other statements,
Stay away from oxides which had its origin in the unsuccessful attempt to nd an oxide with a T
c
sub-
stantially above 10 K. Therefore, the search for new superconductors was focused on cubic intermetallic
compounds with the hope to possibly reach a T
c
close to 40 K, the theoretically expected upper limit for
isotropic metals [57]. In fact, such a compound would have been very advantageous for applications with
liquid hydrogen sufcing as a cryogen.
Beyond the very successful mainstreamresearch on intermetallic superconductors there were also various
innovative ideas on the market. Little, picking up a comment of F. London, theoretically studied the pos-
sibility of superconductivity in organic materials with the conducting and polarizable structural elements
being separated in space [58]. Allender, Bray, and Bardeen continued partially in this direction [59] but
considered also metal-semiconductor hetero-structures [60]. The basic idea is that electron-phonon cou-
pling is expected to be stronger in highly polarizable poorly conducting materials. Then, the electrons in
an adjacent thin metallic layer are being coupled to Cooper pairs via the polarizability of the insulator or
semiconductor over a length scale on the order of the superconducting coherence length
0
as dened in
the microscopic theory of Bardeen, Cooper, and Schrieffer (BCS) [2]. Alternatively, the role of polarons
was explored for essentially homogeneous materials with low carrier density [61]. On this substrate and
with a sound background in ferroelectricity, Bednorz and M uller (Fig. 6.5) started their search for strong
coupling superconductors in the early 1980ies which lead to the discovery of the cuprates.
Figure 6.5: J. Georg Bednorz (left) and K. Alex M uller. By courtesy of the Nobel Foundation.
6.3.2 Materials
Within a few years almost ten families of cuprates have been discovered. They have rather different
crystal structures with occasionally huge unit cells accommodating between 8 and over 100 atoms. The
most popular families are compiled in Table 6.2 along with their maximal superconducting transition
temperatures T
max
c
.
2013
90 R. HACKL AND D. EINZEL Unconventional Materials
Table 6.2: Important families of copper oxygen superconductors. R is for Nd, Sm or even La in the case
of thin lms. n is the number of adjacent CuO
2
planes; three seems to be best for T
c
[62]. Hg-1223 has
the highest T
c
so far. With applied pressure, 150 K can be reached [63].
number of CuO
2
planes n 1 2 3 4 5
chemical formula nickname T
max
c
(n) [K] Ref.
La
2x
Ba
x
CuO
4
LBCO 30 [?, 6]
La
2x
Sr
x
CuO
4
LSCO 39 [64, 65]
R
2x
Ce
x
CuO
4
RCCO 29 [66, 67]
YBa
2
Cu
3
O
6+x
Y-123 93 [68]
Bi
2
Sr
2
Ca
n1
Cu
n
O
2n+4+
Bi-22(n-1)(n) 39 98 110 [?]
Tl
2
Ba
2
Ca
n1
Cu
n
O
2n+4+
Tl-22(n-1)(n) 95 118 123 112 105 [69]
TlBa
2
Ca
n1
Cu
n
O
2n+3+
Tl-12(n-1)(n) 70 103 125 112 107 [69]
HgBa
2
Ca
n1
Cu
n
O
2n+2+
Hg-12(n-1)(n) 96 128 136 123 100 [62, 69]
Structure and chemistry
The copper-oxygen compounds have nearly tetragonal crystal structures with b/a ratios between 1.00
and 1.02. The parent compounds La
2
CuO
4
and Nd
2
CuO
4
(see insets of Fig. 6.8) have been known for
a long time [7072] and have K
2
NiF
4
structure [73] with 3 approximately cubic perovskite-like blocks
per unit cell stacked along the crystallographic c-axis. The resulting c/a ratio is close to 3. According to
the valence count (and also to band structure calculations) the compounds should be metals with a half
lled conduction band (1 electron per unit cell) but the strong electronic correlations block the transport
(for details see section 6.3.3). Therefore, the number of free carriers is zero at half lling (n = p = 0).
If Nd is partially replaced by Ce or La by Sr(Ba) the materials are doped away from half lling, 1+n or
1p, respectively, and become conductors with a small number n or p of mobile carriers directly given
by x. The solubility limits for Ce and Sr(Ba) are x 0.18 and 0.33, respectively. Both compounds have
CuO
2
planes at distances c/2 which are offset by (1/2,1/2) in tetragonal lattice units a. In La
2
CuO
4
the
Cu atom is in the center of an oxygen octahedron, in Nd
2
CuO
4
there is no oxygen in apex position above
and below the CuO
2
plane. In contrast to K
2
NiF
4
, La
2
CuO
4
is tetragonal only at temperatures above
530 K [74]. Below, the structure is slightly orthorhombic with the octahedra tilted about their basal axis
which corresponds to the orthorhombic a

-axis. The transition temperature to the orthorhombic structure


decreases with Sr doping and vanishes at x 0.20. Nd
2
CuO
4
remains tetragonal for all temperatures. In
doped Nd
2x
Ce
x
CuO
4
(NCCO) and in La
2
CuO
4
there is excess oxygen in the structures which has to
be removed for optimal physical properties. La
2x
Sr
x
CuO
4
(LSCO) for x > 0.05 has an oxygen decit
after preparation. It is not clear yet whether fully stoichiometry La
2
CuO
4
and Nd
2
CuO
4
exist at all.
Nevertheless, La
2x
Sr
x
CuO
4
is one of the most intensively studied cuprates, since the entire doping
range is accessible with a single relatively well ordered compound.
YBa
2
Cu
3
O
6+x
(Y-123) is the only material which exists at the stoichiometric composition. Crystals with
the exact cation ratio 1:2:3 can be grown from the ux. When prepared in BaZrO
3
crucibles they are vir-
tually free of defects [7577]. The structures for the limiting doping levels YBa
2
Cu
3
O
6
and YBa
2
Cu
3
O
7
are shown in Fig. 6.6. YBa
2
Cu
3
O
6
has a half lled conduction band but is an antiferromagnetic (AF)
insulator for the same reasons as La
2
CuO
4
and Nd
2
CuO
4
. The copper atoms in the chains have valence
1+, and holes on the CuO
2
planes are generated only when one chain Cu has two oxygen neighbors. For
certain intermediate compositions in the range 0 < x < 1 highly ordered phases can be prepared with do-
main sizes of several 100

A in all 3 crystallographic directions [7880]. Within small limits doping can
also be achieved by replacing Y
3+
by Ca
2+
[81]. Y-123 has two neighboring CuO
2
planes at a distance
c _ Walther-Meiner-Institut
Copper-oxygen compounds SUPERCONDUCTIVITY 91
c
a
b
YBa
2
Cu
3
O
6
YBa
2
Cu
3
O
7
a
Figure 6.6: Structure of YBa
2
Cu
3
O
6+x
(Y-123). (left) YBa
2
Cu
3
O
6
and (right) fully oxygenated
YBa
2
Cu
3
O
7
. The small (blue) sphere in the center is Y, all other small (red) spheres are Cu. Ba and
O are represented as large (yellow) spheres and open circles, respectively. In spite of a half lled con-
duction band (1 electron per unit cell) YBa
2
Cu
3
O
6
is insulating due to electronic correlations and has no
mobile carriers (p = 0). Doping in Y-123 is achieved by adding O to the CuO chains along the crystallo-
graphic b-axis. In this way electrons are removed from the CuO
2
planes. YBa
2
Cu
3
O
7
has 0.82 electrons
or 0.18 mobile holes per CuO
2
formula unit (p = 0.18). This is already slightly above the doping level
optimal for superconductivity (see Fig. 6.8). The CuO
2
planes (shaded, see Fig. 6.9) are common to all
cuprates with Y-123 having two of them at a distance c/3. The maximal T
c
is reached in triple-layer
compounds (see Table 6.2). Details about most of the crystal structures of the cuprates can be found in
Shaked et al. [69].
c/3 a which share one octahedron and are in symmetric positions with respect to the central Y atom
of the conventional unit cell.
Bi-based cuprates, in particular Bi
2
Sr
2
CaCu
2
O
8+
(Bi-2212) [?], are very popular although they grow
only far off stoichiometry and have complex modulated crystal structure according to Ref. [82]. However,
since the two adjacent BiO
2
layers are bound by van der Waals interaction, they can be cleaved easily.
The surfaces obtained in this way are stable, electrically neutral, and typically atomically at over mm
ranges. This makes Bi-2212 the workhorse of surface sensitive experimental methods such as angle-
resolved photoemission spectroscopy (ARPES) and scanning tunneling spectroscopy (STS). The doping
level can be changed via the oxygen content in the Bi O layer and by replacing Ca
2+
by Y
3+
. The
highest T
max
c
at p 0.16 is obtained with 8 % Y doping and reaches typically 96 K [83]. By annealing
the crystals at oxygen partial pressures in excess of 1000 bar [84] p = 0.23 can be obtained. In order to
reach the underdoped range below p = 0.16 annealing protocols with low oxygen partial pressure and
low temperatures are required in contrast to what is expected from the phase diagram [85].
In addition to these three extremely well studied families, the thallium and mercury-based cuprates
are also important since the highest transition temperatures and the most extreme doping levels can
be reached here in materials with T
max
c
above 90 K. In either case, compounds with up to ve neigh-
boring CuO
2
planes exist [69] all of them becoming superconducting above 90 K. The record breaking
HgBa
2
Ca
2
Cu
3
O
8+
has three adjacent CuO
2
planes and reaches a T
c
close to 150 K at 25 GPa applied
pressure [63]. Three CuO
2
planes are apparently optimal for T
c
, for four and more layers T
c
decreases
again. Very recently, HgBa
2
CuO
4+
(Hg-1201) was demonstrated to be another model system in that the
doping can be tuned in the range 0.07 p 0.24 while a very high degree of order can be maintained
manifesting itself in a Meissner effect of almost 100 % [86].
Single layer Tl
2
Ba
2
CuO
6+
(Tl-2201) can be driven non-superconducting metallic on the overdoped
side by high-pressure oxygen annealing [87]. Since optimal doping is slightly below the accessible
2013
92 R. HACKL AND D. EINZEL Unconventional Materials
range the maximal transition temperature at p = 0.16 can only be extrapolated to be close to 95 K in the
latest generation of crystals [88]. Hence, Tl-2201 is expected to display the properties of a true high-T
c
compound in contrast to LSCO.
Growth of crystals
All materials were rst prepared in crucibles. The main complications are that the liquid phases and
uxes corroded all standard materials and that the desired compositions do not normally correspond to
stable points in the high-temperature phase diagrams. Y-123 alone can be grown stoichiometrically from
the ux. Instead of Y other rare earth metals can be used but only crystals with Y are free of site defects.
In all other cases a nite number of rare earth atoms change position with Ba thus reducing the degree
of order. The only stable crucible material is BaZrO
3
which had to be developed rst [7577]. All other
materials get partially dissolved and contaminate the crystals. Crystals grown in BaZrO
3
are shown in
Fig. 6.7.
Figure 6.7: Y-123 single crystals in BaZrO
3
crucibles. BaZrO
3
is not corroded by the ux, which is
aggressive to all conventional crucible materials, and had to be developed rst [75]. After decanting
the ux freestanding crystals, many of them with atomically at surfaces, are left behind on the walls.
From [76] with permission.
Single crystals of NCCO, LSCOand of the Bi family were relatively soon grown in optical furnaces using
the [traveling solvent] oating zone technique ([TS]FZ) with a small volume of appropriate ux between
the feed and the seed rod . When the ux is liqueed the material in the feed rod is dissolved and diffuses
to the colder seed where it crystallizes. Normally, several tricks are necessary to obtain a single crystal
or at least sufciently large grains. The temperature of the melt, the external pressure of the atmosphere
and the oxygen partial pressure are all very critical for a successful growth and have to be determined
systematically [83, 89, 90]. In this way large high-quality single crystals can be prepared in several places
of the world having very similar properties. However, in contrast to Y-123 full reproducibility cannot be
achieved.
For the high vapor pressures of Hg and Tl at elevated temperatures the TSFZ technique cannot be applied
for the preparation of single crystal of the respective families. Using sealed crucibles and a second
containment the volatility and also the toxicity can be controlled reasonably well at the price of a reduced
parameter space for the growth conditions. In spite of these challenges, single crystals of Tl-2201 and
Hg-1201 have nowa reproducibly high quality and reach sizes in the mmrange [8688,91]. The potential
of these materials was already demonstrated and can hardly be overestimated.
c _ Walther-Meiner-Institut
Copper-oxygen compounds SUPERCONDUCTIVITY 93
Thin lms
Along with these bulk methods thin-lm techniques were developed and applied successfully. With a
few exceptions all applications rely on thin lms. There are various methods, (i) chemical vapor depo-
sition (CVD), (ii) sputtering, (iii) pulsed laser deposition (PLD), (iv) molecular beam epitaxy (MBE),
(v) thermal co-evaporation, and (vi) jet printing. All have been known before the advent of the cuprates,
(ii)-(vi) are still used for the cuprates. In a PLD machine a high-energy UV laser evaporates material
from a target with the desired composition. The substrate on which the lm is supposed to be grown is
in the center of the plume of the evaporating atoms. PLD is very efcient and fast and is probably the
most widely used technique. In contrast to PLD, the MBE technique allows full access to the ratio of
the cations. The regulation of the respective ion currents is practically instantaneous facilitating layer-
by-layer growth and an extremely high lm perfection [92]. Articial multi-layers with digital interfaces
can be produced.
Thermal co-evaporation is a comparably simple and cheap approach which turned out to be very suc-
cessful for the production of large-scale lms of Y-123. Since the crucibles with the starting materials
are heated resistively the regulation is only important for the long-term control of the cation ratio. In all
cases the substrates have to be heated to the appropriate temperature. Usually one uses mono-crystalline
material with lattice parameters close to those of the deposited lms. Within certain limits the atmo-
sphere in the preparation chamber can be adjusted. In most of the cases the deposited cuprate lms have
oxygen decits which are xed by in-situ post-annealing protocols [93]. The quality of MBE lms can
come very close to those of single crystals although the structures are not normally fully relaxed. On the
other hand, materials can be prepared which do not exist in equilibrium. This holds particularly true for
electron-doped cuprates [66].
6.3.3 Physical properties
As mentioned above the purity of the samples has crucial inuence on the properties. In particular, the
integrity of the CuO
2
planes is essential to reach the maximal transition temperature of a material class.
In turn, the effect of impurities in the CuO
2
plane can be studied systematically [94] and demonstrates
that the majority of the properties derives from the planes in accordance with band structure calculations.
Here, I try to focus on the properties of virtually clean materials and widely ignore defect-related and
preparation problems.
Phase diagram
The most remarkable property of the cuprates, in my opinion right after the high T
c
, is the universality
of the phase diagrams as shown in Fig. 6.8. Here, the phases are only described separately; possible
interrelations will be postponed to section 6.3.3.
At least on the hole-doped side, the shape of the superconducting dome is almost independent of the
material class. In the doping range 0.05 p 0.27 the transition temperature is well reproduced by [98]
T
c
T
max
c
= 182.6(p0.16)
2
. (6.3.1)
The maximum is always close to 16 % doping but T
max
c
can vary between 38 K for La
1.84
Sr
0.16
CuO
4
[64]
and approximately 150 K in HgBa
2
Ca
2
Cu
3
O
8+
at 25 GPa [63]. A variety of materials following the
universal curve are compiled in Ref. [95]. The relation is valid only in clean materials. In disordered
samples the dome shrinks in that superconductivity occurs in a narrower doping range and at lower
2013
94 R. HACKL AND D. EINZEL Unconventional Materials
insulating metallic metallic
La
2-x
Sr
x
CuO
4
Nd
2-x
Ce
x
CuO
4
300
insulating metallic metallic
2-x x 4 2-x x 4
T*
(
K
)
Cu
O
T* 200
a
t
u
r
e
(
Ln
AF
T
O
100
e
m
p
e
r
a
T
e
SC
SC
0 0.30
0
0.30
Doping p n
Figure 6.8: Generic phase diagrams of electron (left) and hole-doped (right) cuprates. The insets show
the tetragonal unit cells of the prototypical compounds Nd
2
CuO
4
and La
2
CuO
4
, respectively. The shaded
range in the center of the diagram indicates long-ranged antiferromagnetism (AF). The superconducting
dome on the hole-doped side reaches from 0.05 to 0.27 in all clean cuprates [95] independent of the
maximal transition temperature T
max
c
close to 0.16. T
max
c
varies between 30 K in La
2x
Ba
x
CuO
4
[?] and
150 K in HgBa
2
Ca
2
Cu
3
O
8+
at an applied pressure of 25 GPa [63]. T

and T
0
schematically indicate the
onset temperatures of the pseudogap range [96] and of charge and spin ordering [97].
temperatures. This effect can be studied systematically by substituting Zn for Cu, for instance, or by
irradiation [99102]. Also single-layer Bi
2
Sr
1.61
La
0.39
CuO
6+
has a narrower dome and is not super-
conducting for p < 0.10 [103]. In the following we discuss only the clean limit, where all correlation
lengths of ordering phenomena are much smaller than the electronic mean free path . As has been shown
for Y-123 depends more on the sample quality than on the doping and can be in the micrometer range
in YBa
2
Cu
3
O
6.53
with alternating completely lled and empty chains [104].
Given that zero doping (half lling) can be reached, the antiferromagnetism is similarly universal as su-
perconductivity. The maximal N eel temperatures T
N
range between 280 and 420 K. As-grown La
2
CuO
4
has T
N
280 K and only post-annealing in Ar yields T
N
= 325 K [74, 105]. In Y-123 T
N
reaches
420 K [97]. There is no direct scaling between T
N
and T
c
.
Long-ranged AF disappears rapidly on the hole-doped side and somewhat slower for electron-doping, in
any case much faster than one would expect from the percolation limit [106]. However, spin correlations
without long-ranged order can be observed well above T
N
and up to very high doping levels. In LSCO,
Wakimoto and coworkers nd them to disappear along with superconductivity above p
sc2
= 0.27 [107].
Presently, there is not enough experimental material available to decide whether or not spin uctuations
and superconductivity generally coexist up to p
sc2
.
In the range up to p 0.20 there is another transition or crossover at a doping dependent temperature
T

(p) which is usually referred to as the pseudogap line [96, 108]. The nomenclature pseudogap is
related to the observation of a reduced spectral weight in the ARPES spectra measured below T

(p) on
c _ Walther-Meiner-Institut
Copper-oxygen compounds SUPERCONDUCTIVITY 95
fractions of the Fermi surface [109, 110]. This gap in the single-particle properties leaves imprints on
practically all types of responses. In some material classes, and here individual behavior sets in, other
crossover phenomena can be observed below T

(p). Particularly in the La-based compounds, static spin


and charge textures develop [97,111113]. More recently, indications of uctuating ordering phenomena
were also observed in Bi-2212 and Y-123 [114120]. Finally, in some compounds a spin glass phase at
low doping can be observed below some 10 K.
For the rich variety of materials available, the hole-doped side is more exhaustively studied than electron-
doped cuprates. Nevertheless, it is well established that the AF range is broader and the superconducting
dome is narrower for n-doping [66, 90]. A pseudogap was observed by various methods [66, 121123]
and may be a signature of the back-folding of the conduction band at the AF Brillouin zone [124]. The
temperature range of the results from different methods is not yet consistent. While ARPES indicates
the opening of the pseudogap above T
c
[125], tunneling reveals the pseudogap only below T
c
when
superconductivity is suppressed by a high magnetic eld [121]. Systematic transport studies on thin
lms uncover a quantum critical point at n 0.165 [123] which could be the end point of the T

(n) line.
Similarly, a crossover from a small to a large Fermi surface close to n = 0.17 may be interpreted as a
closing of the SDW-like gap [126].
Electronic structure
In conventional superconductors, the band width, the Fermi, phonon, and gap energies are well sepa-
rated, and superconductivity can be treated as a small perturbation of the normal state. In the cuprates,
all energy scales are in close proximity including the magnetic exchange coupling J. The correlation
effects originate in the large Coulomb repulsion U, lead to a substantial incoherent part of the electronic
spectral functions, and k is not a good quantum number any further. Consequently, interaction effects
are interrelated and cannot be observed independently, a fact which still creates confusion.
Instead of dealing with these complications it appears more fruitful to search for the origin of the
variation of T
max
c
in an otherwise rather universal phase diagram. A natural starting point seems the
electronic structure of the CuO
2
plane as the basic building unit in the individual environment of a given
material class.
High energies The most transparent access to the electronic structure of the CuO
2
plane is through
La
2
CuO
4
since with the valences of La and O, given as 3+ and 2, respectively, Cu, residing only in the
plane, is in a 2+ state. The 4s orbital is not relevant for the plane and, on an atomic level, we are dealing
with oxygen 2p and copper 3d states. In the tetragonal environment of the cuprates the degeneracy of the
nine 3d electrons is lifted and the d
x
2
y
2 orbital happens to be the highest occupied one hybridizing with
the oxygen p
x,y
orbitals as shown in Fig. 6.9 (a). The resulting conduction band is half lled and therefore,
on this level of sophistication, La
2
CuO
4
should be a metal. This is also predicted by band structure
calculation in local density approximation (LDA). However, already a Hartree-Fock calculation shows
that the exchange energy is higher than the kinetic energy and blocks the metallic transport. This effect
is usually referred to as a Mott metal-insulator transition. The appropriate description is the Hubbard
model which sets the kinetic and the Coulomb energy in relation. If next-nearest neighbor hopping t
/
is
included in addition to the nearest neighbor integral t for a more realistic description of the cuprates [see
Eq. (6.3.3)] the one-band Hubbard Hamilonian reads [127],
H =

i, j
t(c

i
c
j
) +t
/
(c

i
c
j
) +U

i
n
i
n
i
, (6.3.2)
where c

i
and c
i
creates and, respectively, annihilates an electron with spin on site i and n
i
= c

i
c
i
is the density operator. i, j indicates that the sum is restricted to nearest- and next nearest neighbor
hopping in the case of t and t
/
, respectively.
2013
96 R. HACKL AND D. EINZEL Unconventional Materials
Cu O
M

(b) (a)

k
y

y
k
x
a

Figure 6.9: Electronic structure at E


F
= (T 0) of an idealized quadratic CuO
2
plane. Panel (a)
shows the orbital character of Cu and O (without phases) at low quasiparticle energy
k
=
k
0.
(b) Brillouin zone and Fermi surface of a single CuO
2
plane. The Fermi surface encircles the empty
states around the M points at (, ) for unit lattice spacing a. The diagonal (M) is usually called
the nodal direction since the superconducting gap
k
crosses 0 here (see section 6.3.4). Correspondingly,
the neighborhood of X is called antinode. The number of carriers, electrons or holes, correspond to the
imbalance of the areas around M and . At half lling (1 electron per CuO
2
) the areas equal. The plane
should be metallic but the correlations make it insulating.
If one considers a situation with one hole per copper site the hole tries to hop from site i via the bridging
oxygen to one of the four nearest-neighbor sites j. The kinetic energy which can be expressed in terms
of the transfer integral t is much smaller than the Coulomb repulsion U for double occupancy. In this
way the transport is blocked. In addition, since the Pauli principle is also at work, hopping is only
possible when the spins on sites i and j are anti-parallel hence antiferromagnetically ordered. For t U
the magnetic coupling between the neighboring spins is given as J = t
2
/U in second order perturbation
theory. If one constructs a tight-binding Fermi surface with only nearest-neighbor hopping (t
/
= 0) and
half lling in the idealized quadratic Brillouin zone (BZ) of the CuO
2
plane a square results covering half
of the BZ area and being rotated by 45

[see Fig. 6.9 (b)] which coincides with the magnetic BZ. Due to
the extended parallel parts, the conguration is unstable and a gap would open at the Fermi energy such
as in charge or spin density wave (CDW/SDW) systems [5].
It is instructive to have a closer look at the effect of the Hubbard U on the electronic structure as shown in
Fig. 6.10. On an atomic level, the Cu 3d and the O2p orbitals at
d
and
p
, respectively, are split by some
2 eV. Crystal eld splitting and hybridization broaden the atomic levels considerably, and the Cu 3d
x
2
y
2
and O2p
x,y
orbitals are mixed covalently. The Fermi energy E
F
is in the middle of the anti-bonding
(AB) band indicating metallicity. The dispersionless non-bonding (NB) and the bonding (B) bands are
approximately 3 eV below E
F
. The correlation energy U opens a gap at E
F
, which was rst introduced
by Mott, and the system becomes an insulator. In the case of the cuprates U is approximately 8 eV hence
much larger than [
p

d
[, and all three bands are needed for a proper description.
Now the system will be doped by replacing La
3+
by Sr
2+
. In a one-band picture, part of the copper is
nominally transformed into Cu
3+
for compensation, and hopping becomes possible into empty d
x
2
y
2
orbitals thus opening a channel for transport. This picture is quite useful but according to the preceding
paragraph not quite true. In fact the rst hole is created on oxygen as demonstrated experimentally very
early [128, 130] since the charge-transfer energy [
p

d
[ is smaller than U [?]. More precisely, the extra
positive charge forms a cloud around the central Cu
2+
which compensates also the copper spin and is
therefore called a Zhang-Rice singlet (ZRS) [129]. With minor modication, it still obeys the dynamics
of the one-band Hubbard model as if the charge would reside on the copper.
The fact that [
p

d
[ < U can only be taken into account properly - rather than approximately as
in the one-band Hubbard model sketched in the preceding paragraph - if the Cu 3d
x
2
y
2 , O2p
x
, and
c _ Walther-Meiner-Institut
Copper-oxygen compounds SUPERCONDUCTIVITY 97

UHB
AB E
F

d |
p

d
| U
NB
ZRS
AB
E
E
F

p
B
E
F
LHB
crystal field
covalency
t U
LHB covalency
t U
Figure 6.10: Schematic band structure of the CuO
2
plane at eV energies in electron representation. The
levels
p
and
d
correspond, respectively, to the O2p and Cu 3d atomic levels. The bonding band (B) at
approximately 6 eV below the Fermi level E
F
(long dashes), the dispersionless non-bonding (NB) and
the anti-bonding (AB) band originate from the covalently overlapping Cu 3d
x
2
y
2 and O2p
x,y
orbitals.
All other orbitals are neglected. Since the hopping integral t is much smaller than the Coulomb energy
U the conguration is insulating at half lling, and the AB conduction band is split by U into the upper
(UHB) and lower (LHB) Hubbard band separated by the Mott gap U. For small U the rst hole goes on
copper. If U exceeds [
p

d
[ the material becomes a charge transfer insulator [?] and the rst doped
hole is created on the oxygen [128]. The new state is called a Zhang-Rice singlet (ZRS) [129]. (The
gure is inspired by lecture notes of D. Einzel.)
O2p
y
orbitals are included in a three-band Hubbard model [131, 132]. It has been shown via numerical
solutions of the two models that both give rather similar results for the phase diagrams at T =0 including
superconductivity [133], since the physics apparently does not change qualitatively if the role of the
Hubbard U is taken over by [
p

d
[. Beyond the phase diagrams on either side of half lling also the
electron-hole asymmetry [see Fig. 6.8] can be captured. The holes go on the oxygen atoms and quench
the superexchange coupling between the antiferromagnetically ordered Cu spins while the electrons
appear rst on the Cu site and just dilute the spins. Consequently, the antiferromagnetism survives much
longer on the electron-doped side. To which extent the substantial differences between the material
classes in T
max
c
and other ordering phenomena can be explained through the material dependent ne
tuning of [
p

d
[ is a matter of present research [133].
Energies in the range k
B
T
In order to arrive at energies in the range of a few k
B
T around E
F
, where the relevant physics such as
superconductivity occurs, the high-energy degrees of freedom have to be integrated out [133, 134]. This
procedure is a further idealization and is only qualitative in that individual properties cannot be captured.
Nevertheless, a thorough understanding of the coherent part of the electrons spectral properties is at the
origin of the explanation of the relevant interactions and of superconductivity.
In this spirit, a downfolded LDA band structure [135] can be derived which reproduces the experimental
results from ARPES [136, 137]. This holds particularly true for the shape of the Fermi surface which can
be obtained to within a few percent from the tight binding band structure (now in momentum space),

k
=2t(cos(k
x
) +cos(k
y
)) +4t
/
cos(k
x
)cos(k
y
) . (6.3.3)
Data from slightly overdoped Bi-2212 can be tted satisfactorily using t = 250 meV, t
/
/t = 0.35 and
neglecting band-splitting effects. With /t = 1.1 one arrives at a lling close to p = 0.16. With minor
changes in the parameters the CuO
2
Fermi surfaces of all other cuprates can be reproduced. The Brillouin
zone with the Fermi surface for a single CuO
2
plane is shown in Fig. 6.9. The parametrization of the
2013
98 R. HACKL AND D. EINZEL Unconventional Materials
band structure in terms of hopping integrals is frequently used as a basis for phenomenological modeling
[138141] and microscopic calculations [133, 142, 143].
The experimental dispersion yields a Fermi velocity v
F
2 10
7
cm/s [144] substantially smaller than
the LDA prediction indicating strong interaction effects. A hallmark is the kink-like change of the
slope v
k
= h
1

k
/k at approximately 70 meV observed in the nodal ARPES spectra [145, 146] which
turns out to be quite universal [144]. In the presence of conventional electron-phonon interaction a
kink is expected in the strong coupling limit [14]. In the cuprates, the origin of the interactions is still
controversial. The usual electron-phonon coupling [146150], the coupling to spin excitations [151,152],
uctuations of the charge density [142, 153, 154] and orbital currents [155, 156] have been proposed.
The strong renormalisation is in accordance with the short lifetime
k
() of the electrons which de-
creases rapidly away from the Fermi surface [145, 157]. For a while it appeared that the imaginary part

//
k
() = 2[
k
()]
1
of the electronic self energy =
/
+i
//
is scale free and varies linearly with en-
ergy and temperature T [145] as expected when the electrons scatter on critical uctuations of any
origin (such as spin or charge density or orbital currents). As a consequence, the quasiparticle weight
Z
k
= [1
/
k
()]
1
at the Fermi energy E
F
vanishes and the electrons spectral function is distributed
over very large energy scales in contrast to the properties of a Landau Fermi liquid with 0 <Z
k
<1. This
phenomenology is called marginal Fermi liquid (mFL) [158] and has a big share in the discussion of the
cuprates.
With the continuously improved resolution of ARPES experiments, various substructures were found at
low energies and analyzed in terms of phonons [147, 148] and spin uctuations [151, 157]. In all cases,

//
k
() varies faster than linear at low energies and crosses over to the more linear behavior in the 50 meV
range. This phenomenology is expected if the electrons couple to a dispersionless (Einstein) mode which
must not necessarily be a lattice vibration. These recent observations let it appear very likely that several
interactions contribute to the coupling spectrum.
At a given energy ,
k
() depends strongly on the position on the Fermi surface. Along the diagonal
M line of the BZ (nodal direction)
//
k
() is relatively small and does not depend substantially on
doping [136, 137]. In the vicinity of the X point (antinode) things are more complicated. At very
high doping as in Tl-2201, the lifetime of antinodal quasiparticles may even be longer than that of
nodal ones [159]. With decreasing doping the lifetime and the weight of the antinodal quasiparticles
decreases continuously. Slightly below optimal doping the interactions along the principle directions
become strong enough to completely suppress coherence even in the superconducting state [160, 161].
At elevated temperatures no quasiparticle develops any more for p < 0.18 indicating a loss of coherence
in the pseudogap regime on parts of the Fermi surface [137].
Bridging the gap At rst glance one would expect that the renormalized conduction band with a width
in the range of an eV and the UHB and LHB split by U 8 eV do not overlap. However, the strong
correlations leave only a small fraction Z
k
1 of the quasiparticles spectral weight close to E
F
and
distribute 1Z
k
over large energy scales. This incoherent part can be observed experimentally between
the LHB and the conduction band as a faint structure with almost vertical dispersion [162164] and is
reproduced theoretically using Monte-Carlo techniques for solving the Hubbard model [165, 166]. Thus,
there is one more indication that the Hubbard model is a reasonable starting point for the description of
the CuO
2
planes in the cuprates.
Two-particle dynamics
The CuO
2
planes determine the majority of the physical properties and, in particular, carry the currents in
the normal and in the superconducting states. Owing to the layered structure, the cuprates are electrically
c _ Walther-Meiner-Institut
Copper-oxygen compounds SUPERCONDUCTIVITY 99
highly anisotropic [167]. The anisotropies of the resistivities
c
/
ab
, with the subscripts indicating the
crystallographic directions, range from 30 to 10
5
[168, 169].
Transport At optimal doping all clean cuprates have an essentially linear resistivity in the a b plane
down to T
c
,
ab
(T) T, which saturates only at very low temperature, as demonstrated for LSCO in
magnetic elds [170]. Below a doping dependent temperature T

(p) T

(p) there is a reduction of

ab
(T) below the linear variation which is associated with a pseudogap [108, 168, 169, 171, 172]. If the
high-temperature part is extrapolated to T = 0 the residual resistivity is very small and approaches 0 for
the cleanest optimally doped crystals. If superconductivity is suppressed by high magnetic elds
ab
(T)
saturates at a nite value [170]. For p > 0.16, also the out-of-plane resistivity becomes purely metallic
and
c
(T)
ab
(T). The ratio
c
/
ab
(T 1.5T
c
) at optimal doping is approximately 30 in Y-123 [168]
and 5000 in Bi-2212 [169] having the same T
max
c
and close to 500 in LSCO with T
max
c
= 38 K [170].
There is a dichotomy between the under- and the over-doped ranges on the hole-doped side which be-
comes particularly clear at low temperatures when superconductivity is suppressed by magnetic elds.
Various systematic studies have been carried out recently by Ando and coworkers [64]. The results for
LSCO and Y-123 are shown in Fig. 6.11. In both Y-123 and LSCO the resistivity generally turns in-
(d)
(c)
(a)
(b)
Figure 6.11: Resistivity vs T for different doping levels p for LSCO (a,b) and Y-123 (c,d). The compari-
son indicates the similarity of the different classes as long as the crystals are sufciently clean. Note the
deviation from linearity below a doping dependent temperature T

(p) which is particularly clearly seen


in Y-123 at 0.11 p 0.15 corresponding to 6.60 y 6.85. From Ref. [64] with permission.
sulating, corresponding to d
ab
(T)/dT < 0, at low temperatures and doping before superconductivity
appears at p
sc1
0.05. For p > p
sc1
d
ab
(T)/dT becomes essentially positive for T > T
c
. If a magnetic
eld is applied which is high enough to suppress superconductivity completely an upturn is observed
with a logarithmic divergence towards zero temperature. With increasing p the minimum shifts to lower
temperature and approaches T = 0 close to p = 0.17 [170]. For p > 0.17 the resistivity remains metallic
and exhibits a T

variation over extended temperature ranges with 1.5. Apparently, full metal-
licity develops above optimal doping. In strongly over-doped Tl-2201 with T
c
15 K, 1.5 < < 2 is
found [173]. If the resistivity is tted to (T) =
0
+AT +BT
2
the coefcient A of the linear term
approaches zero as [p
sc2
p[ for p p
sc2
[174, 175].
The details become more transparent when the dynamics is studied as a function of the electron mo-
mentum k as discussed already briey in the context of the electronic structure and single-particle life-
times in paragraph 6.3.3. Concerning transport, two-particle properties have to be considered where an
electron is scattered from an occupied into an empty state leading to the usual restrictions and cor-
rections. In an early nuclear magnetic resonance (NMR) experiment, deviations from the Korringa
law, (T
1
T)
1
=const, with T
1
the spin lattice relaxation time and T the temperature was found well
above T
c
for p 0.17 [176], indicating the loss of a relaxation channel for the electrons. The NMR
2013
100 R. HACKL AND D. EINZEL Unconventional Materials
B B
2000
T = 200K
2g
B
1g
B

opt.
doping
(
K
)
Bi-2212
T =200 K
1000
X M X M

0
X
-M
(a)
0
2
K
/
K
)
(a)
-M
2
0

/
T

(
K

0
X
0.05 0.10 0.15 0.20 0.25
-4
-2

(b)
doping p
Figure 6.12: In-plane anisotropy of the electronic relaxation

= [

]
1
at 200 K as seen by electronic
Raman scattering [183]. (11.6K = 8cm
1
= 1meV)

is closely related to a resistivity , and


1
=
ne
2
m
in a Drude model. As an additional information from Raman scattering, the regions around the X
points (diamonds) and along the M line (squares) of the Brillouin zone (see Fig. 6.9) can be projected
independently with different light polarizations [181]. (a) There is little doping dependence along the
nodal directions (M). At X there is an abrupt change at p = 0.210.01. The crossover seems to be
universal since it is also seen in LSCO and in Tl-2201 [184] and by NMR [185]. It is predicted by the
Hubbard-Holstein model [143]. (b) The temperature dependence of

/T[
200K
is isotropic above
p = 0.210.01. Below the crossover, the antinodal derivative decreases continuously and changes sign
close to p = 0.16.
form factors suggest that particles close to (, 0) may experience a gap which was directly observed by
ARPES [109, 110]. Similarly, optical transport (IR) results in Y-123 show that the electrons with mo-
menta along the diagonal relax differently from those at the X points of the BZ. The distinction is pos-
sible for the specic crystal and band structure of Y-123 which facilitates to project diagonal and (, 0)
momenta for in-plane and out-of-plane polarizations, respectively [135]. For this reason, the pseudo-
gap as an anti-nodal property was discovered rst by c-axis polarized IR spectroscopy [177]. However,
the projection in optical spectroscopy with in-plane polarizations is incomplete with nite sensitivity
everywhere in the Brillouin zone, and the pseudogap is clearly visible below optimal doping also for
E|a, b [178, 179].
The electronic Raman response measures a quantity similar to the conductivity [180, 181] but has in-
plane selection rules which facilitate independent access to nodal and antinodal electrons by appropri-
ately selecting the light polarizations [181, 182]. Results for Raman relaxation rates

( is for the
polarizations corresponding to symmetry projections [181]) of differently doped Bi-2212 at 200 K are
shown in Fig. 6.12. In the nodal conguration the doping dependence of the spectra and of the corre-
sponding carrier relaxation is weak for p
sc1
< p < p
sc2
[183, 186, 187]. Above p 0.21 no polarization
dependence corresponding to a relaxation anisotropy can be observed. The relaxation rate for antinodal
electrons increases abruptly below p 0.21, and approximately 30 % of the spectral weight is lost in the
energy range up to 250 meV. This was traced back to a doping and momentum dependent correlation gap
c _ Walther-Meiner-Institut
Copper-oxygen compounds SUPERCONDUCTIVITY 101
extrapolating to 2
C
200 meV at p = 0 [183]. A similar phenomenology emerges with a progressive
loss of quasiparticle coherence starting at X and proceeding to the node upon reducing p as observed
by ARPES [109, 110, 137, 188] and studied also in the context of light scattering [186, 189, 190]. The
correlation gap, the pseudogap and, nally, the superconducting gap (see section 6.3.4), have different
energy scales and their interrelation has to be determined yet.
The onset of anomalies in the doping range around p = 0.16 are also seen in the Hall effect [191] and,
particularly clearly in the Nernst signal [175, 192194] constituting, respectively, a transverse voltage
in response to a charge and heat current in a perpendicular magnetic eld. The Hall effect exhibits a
maximum close to T

[175, 194] but only the quantum oscillations of the Hall resistivity at very low
temperature indicate that the anomaly may be related to a reconstruction of the Fermi surface [117].
Very recently, indications of a Fermi surface reconstruction were also discovered on the electron-doped
side in NCCO [126].
The Nernst effect is sensitive to superconducting vortices [195] and density-wave order [196]. For a long
time the Nernst signal observed between T

and T
c
in LSCO was considered a signature of vortex motion
above the coherence temperature T
c
in the spirit of a 2D Kosterlitz-Thouless transition [192]. Only recent
results in Eu-doped LSCO showed that the onset of the Nernst voltage coincides with the charge-ordering
temperature [194, 197] found in various other experiments [?, 113]. In Y-123 the onset of the anisotropic
Nernst signal [175, 192, 193] coincides with various other indications of broken rotational symmetry
such as Kerr rotation [198] or incommensurable peaks in the dynamic spin susceptibility [118, 119].
New frequencies in the quantum oscillations indicate that a partial reconstruction of the Fermi surface
goes along with the ordering phenomena. Hence, the superstructures found rst in the the spin channel
in LSCO [199, 200] and Nd-doped LSCO [111] seem to be a generic phenomenon of all cuprates that is
accompanied by charge order. Since the superstructures are static only in exceptional cases signatures
of them in the transport escaped observation for a long time, in particular in the compounds with high
T
c
. Their importance is being unveiled only slowly. Further details will be discussed at the end of the
following subsection and in sections 6.3.3 and 6.3.4.
Spin dynamics Homogeneous magnetism exists in wide doping ranges. At 0 p 0.03 the antiferro-
magnetism is long-ranged in LSCO. In NCCO three-dimensional (3D) antiferromagnetism exists below
n = 0.13. The exchange coupling J 130 meV is among the largest ones existing. The order is truly 3D
but the coupling along the c-axis is orders of magnitude smaller than along a.
The dynamics at high energy was studied early by Raman scattering. The photon ips essentially two
neighboring spins breaking six bonds with energy J [201]. More accurately, a two-magnon density of
states is measured and projects the at parts of the dispersion in the vicinity (, 0). The maximal energy
observed is therefore at E
2M
6Js. In a spin 1/2 system the peak is close to 3J (2.7J for quantum
corrections) [?, 201, 202]. In LSCO, Y-123 and Bi-2212 spin correlations can be observed by Raman
scattering up to approximately p = 0.20 [?, 187, 203]. In LSCO and NCCO magnetic short-range order
was observed by inelastic neutron scattering up to p 0.27 [107] and n 0.17 [204], respectively.
For the large magnitude of J the full dispersion of the spin excitations was studied with neutrons only
recently. The spectrum extends beyond J well above the energy of thermal neutrons. Results for various
compounds up to approximately 200 meV are shown in Fig. 6.13 [97]. If the energy axis is normalized to
the exchange coupling J the dispersions collapse on top of each other lending evidence to the universality
of the spin excitations.
At intermediate and low energies spin excitations were studied in detail by neutron scattering and NMR.
The decrease of the Knight shift below T
c
indicated spin singlet pairing [207] (see section 6.3.4). For
T >T
c
, the spin-lattice relaxation rate T
1
1
is proportional to T compatible with Fermi liquid-like carriers
[176,207] only close to optimal doping and above. At low doping a spin gap is found below T
c
by neutron
scattering [208] and magnetic resonance [176, 209] putting magnetism and superconductivity in relation.
It was conjectured early that most of the spin susceptibility results from itinerant electrons rather than
2013
102 R. HACKL AND D. EINZEL Unconventional Materials
-0.2 -0.1 0 0.1 0.2
0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
(0.5+h, 0.5) (rlu)

E

/

J
La
1.90
Sr
0.10
CuO
4
La
1.875
Ba
0.125
CuO
4
La
1.84
Sr
0.16
CuO
4
YBa
2
Cu
3
O
6.5
YBa
2
Cu
3
O
6.6
Figure 6.13: Dispersion of the spin excitations of La-based cuprates and in Y-123. If the energies are nor-
malized to the exchange coupling J the spectra are universal. Low and high-energy parts are compatible
with uctuating charge and spin order [205, 206]. From [97] with permission.
localized Cu moments. This inspired the model of a nearly antiferromagnetic Fermi liquid [210]. The
interaction of carriers and spin uctuations can be studied systematically in the uctuation exchange
(FLEX) scheme [?, ?]. However, in which way spin uctuations participate in the Cooper pairing is still
an open question (see section 6.3.4).
With polarized neutrons an intriguing narrow mode with wavevector Q = (, ) was found at low tem-
peratures [208, 211215] which is usually referred to as the -resonance. For p 0.16 the energy of the
resonance E
R
is proportional to T
c
. In the underdoped range the mode appears already between T

and
T
c
when the pseudogap opens up. The spectral weight of the mode is between 1 and 6 % of the integrated
spectral weight of the spin susceptibility [216] and its origin is controversial. Its role as a mediator of
superconductivity has been explored in various studies [152, 217219]. The results, however, did not
generate general agreement yet. In spite of that the -mode is characteristic of the cuprates and possibly
other superconductors in close proximity to a magnetic phase [?].
In the La-based compounds incommensurate peaks shifted by (0, ) and (, 0) (in the square unit cell
of the CuO
2
planes) away from the AF reex at (/2, /2) were discovered early in the inelastic channel
indicating a dynamic superstructure on top of the antiferromagnetic order for p = 0.075 and 0.14 [199].
If part of the La is replaced by Nd the superstructure becomes static [111]. Charge order accompanied
by a lattice distortion with a periodicity of four unit cells appear before the eight unit cell superstructure
of the spins is established. Below the onset point of superconductivity, p p
sc1
0.05, static diagonal
stripe order is observed in LSCO. At p
sc1
the stripes rotate by 45

and start to uctuate meaning they


can only be observed at nite energy [199, 220]. Generally, charge order precedes spin order upon
cooling [111, 113, 221]. Fluctuating modulations of the charge density cannot normally be observed
by neutron scattering but can be visualized by tunneling spectroscopy due to interference effects [114].
Fig. 6.14 shows that charge and spin order in Bi-2212 and LSCO have the same orientation above p
sc1
.
Recently, equally oriented nematic order was also observed in underdoped superconducting Y-123 in the
dynamic spin susceptibility [118, 119]. Assuming a stripe-like superstructure of the spins the dispersion
(Fig. 6.13) can be predicted quantitatively [205, 206, 222]. Hence, evidence mounts that dynamic phase
c _ Walther-Meiner-Institut
Copper-oxygen compounds SUPERCONDUCTIVITY 103

Bi
2
Sr
2
CaCu
2
O
8+
La
2-x
Sr
x
CuO

(c)
Figure 6.14: Spin and charge ordering in LSCO (a,b) and Bi-2212 (c), respectively. Parts (a,b) display
incommensurate neutron reexes at (h, k) = (0.5 , 0.5 ) below p
sc1
0.05 and at (0.5 , 0.5)
and (0.5, 0.5) above p
sc1
. They originate from a static and dynamic spin superstructure, respectively,
corresponding to static and uctuating stripes [111, 199, 220]. Above p
sc1
the spin and charge superstruc-
tures are observed to have the same orientation. Superstructures with similar orientations including the
rotation [see panel (b)] are also seen in Y-123 [116, 118, 119, 175]. From [220] with permission. The re-
sult on Bi-2212 (c) is a ltered STS image and shows a modulation of the charge density which becomes
visible due to interferences. The arrows indicate the orientation of the CuO
2
planes. From [114] with
permission.
separation and the related ordering phenomena are generic properties of the cuprates contributing to
anomalies such as critical uctuations, Fermi surface reconstruction and the pseudogap in the electronic
excitation spectrum.
Competing phases
The pseudogap range is one of the most intensively studied areas of the phase diagram being observed
below the T

line (see Fig. 6.8) and for doping levels below approximately p = 0.21 [96, 108, 172,
223]. Various properties discussed above indicate a gap in the quasiparticle excitation spectrum and,
hence, an instability above the transition to superconductivity. It is clear from ARPES [109, 110] that
the Fermi surface is not fully gapped above T
c
. Concomitantly, the materials remain metallic, and the
resistivity even decreases slightly below T

since the strongly interacting quasiparticles are gapped out


(see Fig. 6.11).
To some extent the resistivity of the cuprates (Fig. 6.11) indicates similarities to CDW or SDW systems
such as the recently discovered FeAs superconductors [7, 8, 31], where the resistivity also drops upon en-
tering the ordered phase in the undoped parent compounds [37]. Similarly, in several f -electron systems
a magnetically ordered phase is suppressed as a function of either doping or pressure giving room for
superconductivity [27]. In either case, superconductivity is in close proximity to other ordered phases, in
complete contrast to conventional materials [?]. Clearly, there is another instability above the supercon-
ducting phase, and it is of pivotal importance to understand the relationship of the phases as to whether
they compete or cooperate [96].
In most of the cases, in particular in the compounds with high transition temperatures, T

is a crossover
rather than a phase transition. There are various indications of a broken symmetry [114, 142, 223228]
2013
104 R. HACKL AND D. EINZEL Unconventional Materials
with the uctuations of incipient order widely considered important and responsible for many of the
anomalous properties of the cuprates including superconductivity. Very early Varma and coworkers in-
troduced the concept of a quantum critical point above which temperature is the only energy scale rather
than collective excitations such as spin waves or phonons. The related uctuations lead to an almost com-
plete collapse of Landaus Fermi liquid model and to the marginal Fermi liquid phenomenology [158],
where an electron has vanishingly small coherent weight even at E
F
. The quantumcritical point (QCP), at
which the uctuations suppress any phase transition above absolute zero, is somewhere buried below the
superconducting dome in the range 0.15 < p < 0.22 on the hole-doped side [175] and close to n = 0.165
for electron-doped systems [123]. Order or incipient order (for reviews see Ref. [28] or the book by
Sachdev [229]) are expected below T

(p) and between n = p = 0 and the critical doping. Similarly as


in many other systems the QCP cannot be accessed directly since it is protected by superconductiv-
ity. The appearance of superconductivity above a QCP is one of the reasons why the uctuations are
considered a possibility to mediate Cooper pairing.
It is controversial which types of uctuations dominate. Inspired by the NMR results Anderson proposed
the resonating valence bond (RVB) model where uctuating spin singlets are formed at high temperature
and condense below T
c
[230]. A similar phenomenology follows if polarons condense into bi-polarons
[61]. Ong and coworkers interpreted the onset of the Nernst signal between T
c
and up to maximally 3 T
c

at p 0.10 in terms of superconducting uctuations which survive even below p


sc1
[192]. Recent
studies in La
2xy
Eu
y
Sr
x
CuO
4
(LEuSCO) show that the Nernst signal sets in along with the formation
of a CDW-like superstructure and may originate from the related charge ordering [194, 197] rather than
from vortex motion [196]. While the superstructure is static in LEuSCO, LNdSCO [111], LBCO and
La
2x
(Ba
1y
Sr
y
)
x
CuO
4
[221, 231] below T
0
(Fig. 6.8) uctuating order is observed in LSCO [220] and
also in Y-123, at least at specic doping levels [116, 117, 119, 191]. Fermi surface reconstruction has
also been observed recently in NCCO. It can be described in terms of band folding resulting from the
AF order [89, 126]. Yet, the details and the origin behind the reconstruction remain important problems
to solve on either side of zero doping.
As already mentioned, to some extent there is a similarity to CDW and SDW materials with dimension
d greater than one, where only part of the Fermi surface is gapped while the rest sustains metallicity
[232, 233] or even superconductivity such as in 2H-NbSe
2
[234]. Beyond these similarities the type of
order in the cuprates has many new features. In particular in the high-T
c
compounds, an ordered phase is
not established, and it is probably sensible to speak of nematic order [114, 228], including spontaneous
deformations of the Fermi surface [235], with only the rotational symmetry broken.
The lattice appears to play a crucial role in stabilizing the order. It has been shown for La-based com-
pounds that the tilting angle
t
of the copper-oxygen octahedra may be a parameter to quantify the
proximity to static order [112]. If
t
exceeds a critical value static order is established by kind of a
lock-in transition and superconductivity disappears. Hence, in realistic models electron-phonon interac-
tion should be included in the Hubbard model [142] to bring the derived phase diagrams closer to the
experiments [143, 185].
There is no evidence whether and which uctuations contribute to superconductivity (see also sec-
tion 6.3.4). However, it was shown for LBCO at 1/8 doping and for LEuSCO that static order quenches
the 3D phase transition [112, 236]. In the case of LBCO the CuO
2
planes decouple and the phase transi-
tion to 2D superconductivity is suppressed by uctuations [236]. Upon applied pressure the static order
becomes nematic and superconductivity is restored [237].
There are two conclusions. (i) The pseudogap phase has many signatures of a broken symmetry other
than the gauge symmetry of superconductivity. There are many experimental indications that the rota-
tional symmetry is broken and that the electronic states partially reconstruct due to incipient charge order
driven by the strong correlations. The contribution of superconducting uctuations to the pseudogap is
c _ Walther-Meiner-Institut
Copper-oxygen compounds SUPERCONDUCTIVITY 105
small as can be seen independently from the spectral weight redistribution in the optical conductiv-
ity [238]. (ii) As soon as the order becomes static superconductivity is quenched. In this sense there
is a competition between the two phases, and the possible coexistence is clearly different from that in
conventional CDW and SDW systems [234, 239]. To which extent the critical uctuations of incipient
order contribute to or drive superconductivity needs to be claried [96].
6.3.4 Superconductivity
Superconductivity in the cuprates mesmerized nearly all solid state scientists for its unprecedented ro-
bustness. The essential parameters of Y-123 are summarized in Table 6.3 and compared with those of Al
and Nb
3
Sn. Applications on a large scale were expected to be realized within a few years. The transition
temperatures were high enough to make cooling with liquid nitrogen an option. Upper critical elds in
the 100 T range and critical current densities of j
c
10
7
Acm
2
close to those of the best metallic alloys
and three to four orders of magnitude above the maximal capacity of Cu triggered expectations of com-
pletely replacing power transmission lines, storing energy or constructing magnets with elds in excess
of 30 or even 40 T virtually free of energy consumption. However, a brief look at Table 6.3 shows were
the problems are buried. Nevertheless, substantial progress could be made since 1986 and several of the
ideas have become commercial products (see chapter 7).
Table 6.3: Superconducting parameters of Al, Nb
3
Sn, and Y-123. The data for Al and Nb
3
Sn are taken
from Ref. [10]. The references for Y-123 are indicated in the last column. Some entries are estimated
using the relations
BCS
= hv
F
(
0
)
1
and B
c2
=
0
(

2
GL
)
2
with v
F
the Fermi velocity and
0
=
2.07 10
15
Wb the ux quantum. The critical current for Al is determined via the Silsbee criterion [10]
for a wire with 1 mm diameter. All derived quantities should be considered order of magnitude estimates.
Note that
BCS
and
GL
are related but different quantities. has no index since it can be measured with
some accuracy (see, e.g., Ref. [104]) Also in the case of , the London and the Ginzburg-Landau (GL)
denitions should be distinguished.
quantity unit Al Nb
3
Sn Y-123 comment Ref.
T
max
c
K 1.19 18 93 [81]

0
meV 0.18 4.3 35 [240]

0
k
B
T
c
1.76 2.7 4.3

ab
BCS

A 2 10
4
100 20
B
c
c1
T 0.01 0.1 0.05 B|c
B
c
c2
T 24 130 B|c [241]
B
ab
c2
T 240 B|ab [10]

ab
GL

A 15

c
GL

A 40 1.5

ab

A 500 800 900100 B ab [104]

GL
0.5 20 80
j
c
(5K, 10T) A/cm
2
1.6 10
3
10
6
5 10
7
[242]
2013
106 R. HACKL AND D. EINZEL Unconventional Materials
Experiments
One can spot the short coherence length
0
(used if specialization to
BCS
or
GL
is not necessary) to
be among the major problems to deal with, since it prevents effectively the pinning of ux-lines. The
resulting ux ow goes along with energy dissipation and kills all applications in high elds and with
large currents. Typically, one pinning center per coherence volume
2
0,ab

0,c
is needed. Hence, in con-
ventional alloys ux ow can be suppressed by a moderate density of defects (see Table 6.3). In addition,
(non-magnetic) impurities have little impact on superconductivity in an s-wave superconductor [243]. In
contrast, T
c
is rapidly reduced in the cuprates, and a high density of pinning centers is required. Fortu-
nately, the structure helps, but one had to learn to keep the pinning centers away from the CuO
2
planes.
For instance, oxygen clusters on the CuO chain sites of Y-123 pin effectively [?] and have only mild
inuence on T
c
[78]. For practical purposes the proper distribution of pinning centers is among the major
challenges, and the dynamics of ux lines remains an important eld of research [?,?]. What is the origin
of the short coherence length and of the sensitivity to disorder?
One could phrase it this way: you have the choice between Skylla and Charybdis. The high transition
temperature goes along with a large energy gap which results in a short coherence length,
BCS
=
hv
F
()
1
. The high transition temperatures in turn, come from an exotic coupling mechanism which
makes superconductivity in the cuprates unconventional. Following the denition proposed by Pitaevski
[17] and Brueckner et al. [16] unconventional means that

k
2
_

2
k
+[
k
[
2
= 0. (6.3.4)
Hence the gap
k
is strongly anisotropic, changes sign and has nodes on the Fermi surface making T
c
highly susceptible to defects. The sign change is topologically different from a strongly anisotropic but
generally positive gap with vanishingly small minima and goes along with a discontinuous transition
from a four-fold to a two-fold rotational symmetry which implies a change in the phase of the gap
similar to the structure of atomic orbitals with l 1. Since spin singlet pairing was identied early by
NMR [207], odd internal angular momentum of the Cooper pairs going along with spin triplet states can
be excluded. Hence, l = 2 is the lowest possible angular momentum and d
x
2
y
2 is realized.
Experimentally, the sign change of the gap can be demonstrated only in a phase-sensitive experiment
[18, 19] and not by spectroscopy probing the magnitude of the gap [
k
[. The d
x
2
y
2 character was pinned
down by Wollman and coworkers [18] and consecutively corroborated in various ways for both electron
and hole-doped cuprates [19].
On the Fermi surface k = k
F
the d-wave gap is simply given by

=
0
cos(2) with
0
the gap
maximum and the azimuthal angle which is zero on the M-X line [see Fig. 6.9] with the origin in M.
On the tight-binding band structure, as given in Eq. (6.3.3), the gap is parameterized as

k
=

0
2
[cos(k
x
a) cos(k
y
a)] (6.3.5)
for a quadratic unit cell with lattice parameter a.
It is an enchanting coincidence that the paper on the unconventional gap in UBe
13
[13] directly precedes
the article on superconductivity in La-Ba-Cu-O [6]. There, the magnetic penetration depth (T) was
used as a diagnostic tool which also brought the break-through for the cuprates [20]. Although there
were very early indications that the gap is strongly anisotropic and may even have nodes [244247], only
the experiment of Hardy and coworkers on high-quality Y-123 single crystals [20] triggered an avalanche
of activities including the rst phase sensitive experiment by Wollman et al. [18].
c _ Walther-Meiner-Institut
Copper-oxygen compounds SUPERCONDUCTIVITY 107
To map out the magnitude of the gap in the cuprates spectroscopically, resolution in k-space is needed.
For this reason ARPES became particularly important since one can map electronic single-particle ener-
gies with a resolution E 2 meV as a function of the in-plane momentum k
|
with a resolution of the
Fermi surface angle of better than one degree. The cuprates and ARPES proted mutually from each
other, in kind of a symbiosis, since Bi-2212, due to its extremely two-dimensional structure and the ex-
ceptional cleaving plane between the Bi-O layers, facilitated deep insights into the physics of the cuprates
by photoemission and thus enormously fueled the method itself [137]. The most important results are
the observation of the Fermi surface and of the momentum dependence of [
k
[ [248, 249] as shown in
Fig. 6.15, of the pseudogap [109, 110] having the same momentum dependence as [
k
[, the doping de-
pendence of the dispersion [136, 188], and various renormalization effects on the band structure which
are believed to be in close but not yet understood relationship with the Cooper pairing [146, 151, 250].
Fermi-surface angle (deg)
Figure 6.15: The magnitude of the gap [
k
[ as a function of the Fermi surface angle as dened in Fig. 6.9.
Note that the labels on the Brillouin zone in the inset correspond to the reciprocal lattice of Bi-2212. The
inset shows experimental points for the Fermi surface (circles) and the tight-binding t (full line). The
hairlines indicate the replica originating from the superstructure of the Bi-O layers. From [249] with
permission.
Electronic Raman scattering [251] is among the few other possibilities to see the gap anisotropy directly
since different parts of the Fermi surface are projected independently by appropriately adjusting the
polarizations of the incoming and outgoing photons [181, 182]. Since light scattering is a two-particle
method both the gap in the excitation spectrum and the condensate are seen. At optimal doping the results
agree with those fromARPES. Additional information is obtained predominantly at more extreme doping
levels closer to the onset points of superconductivity. It turns out that the gap close to the nodal direction
scales with T
c
in very wide doping ranges [?,190,240,252256] in qualitative agreement with low-energy
tunneling [257] and recent ARPES results [258]. In the latter experiment the particle-hole mixing typical
for (k,-k) pairing can be seen below but close to T
c
at E > E
F
. This identies the observed gap as the
superconducting one.
On the electron-doped side ARPES [125] and Raman scattering [259, 260] reveal gap magnitudes

0
/k
B
T
c
in the range 2 to 2.5, much smaller than for hole-doping. Phase sensitive experiments show
that the gap changes sign similarly as on the hole-doped side [19, 261]. The gap appears to vary non-
monotonically [125, 260]. However, the interference with the pseudogap may inuence the magnitude of
2013
108 R. HACKL AND D. EINZEL Unconventional Materials
the superconducting gap and needs to be claried further. The relatively small gap ratios signal interme-
diate to weak coupling, and it is not overly surprising that
0
approximately follows T
c
in the relatively
small doping range where superconductivity exists [262].
For p-doping, the approximate scaling of the gap extracted from extended portions of the Fermi surface
around the nodes as
0
4.5k
B
T
c
in the under- and over-doped ranges [181, 190, 254, 257] is rather
surprising as one would expect the gap to reect the supposedly doping dependent coupling strength. In
fact, most of the electron-electron interactions such as those from spin uctuations or Coulomb repulsion
increase towards p = 0. However, all these coupling potentials are not isotropic but dominate along the
principle axes so that the nodal part could be less inuenced. It has been argued that the increase of
the gap may be compensated by the loss of major parts of the Fermi surface due to the interaction
itself [137, 189, 190, 263265]. The reduction of the superuid density towards low doping may be a a
ngerprint of this phenomenon [172, 266, 267] but there is no quantitative understanding yet.
One would expect that this problem could be claried by looking at the environment of the X points (anti-
node) where the strong interactions prevail. However, the loss of coherent quasi-particles and the opening
of the pseudogap for p 0.19 progressively shroud the pairing dynamics. This is further complicated by
the emergence of inhomogeneities which can be observed by scanning tunneling spectroscopy (STS) as
shown in Fig. 6.16. In the spectra of Bi-2212 large and small gaps are spatially separated, and the large
gaps line up with oxygen defects where the doping level is expected to be reduced [268]. The pseudogap
and the superconducting gap may even mix in some doping ranges [223, 269, 270].
Sample Bias (mV)
d
I
/
d
V

(
a
r
b
)
0 5 0 0 1 -
(a)
1
2
3
4
5
6
-50 0
(b)
Figure 6.16: Inhomogeneity of the tunneling spectra of slightly underdoped (p = 0.15 0.01) Bi-2212
as seen by STS. The spectra in panel (a) are measured at the spots in panel (b) having the same color.
Note that the slope close to zero bias depends only little on the position. The asymmetry for positive and
negative bias results from the charge order. From [268] with permission.
Affairs do not simplify on the overdoped side. While the condensation energy has a maximum close to
p = 0.19 and the superuid density tends to saturate [172, 267] the nodal and anti-nodal gaps continue
to develop independently. In addition, the interpretation of the coherence peaks remains controversial.
There are particularly enlightening experiments. (i) Electronic Raman scattering with applied pressure
demonstrates that the superconductivity-induced anti-nodal structures decouple from T
c
already at op-
timal doping [271]. (ii) In STS the energy of the coherence peaks in the range of
0
is lower than
one would expect from the slope close to zero bias [257, 268, 270] and depend on the location on the
sample [268].
The -mode at energy E
R
(see section 6.3.3) follows T
c
(p) and E
R
(p) 1.3
0
(p) on the overdoped side
[?, 211, 213215, 272]. On the underdoped side, data on Y-123 [213, 272] show that the scaling with T
c
is
not valid any more. As to whether or not the proportionality to
0
takes over depends on the denition of
the gap which, in my opinion, remains problematic, in particular in the presence of the pseudogap. Hence,
the -resonance proves to be in close relationship to unconventional superconductivity [?, 216, 219],
while the more stringent question as to its relationship with the magnitude of the gap and with the origin
of the Cooper pairing is not settled.
c _ Walther-Meiner-Institut
Copper-oxygen compounds SUPERCONDUCTIVITY 109
In summary, not all microscopic properties in the superconducting state are claried. While the symmetry
of the energy gap is found to be universally d
x
2
y
2 [19] the magnitude, which could be instrumental to
identify the origin of Cooper pairing, is hard to pin down. Close to the node scaling with T
c
seems likely
but around X discrepancies as large as 50 % between different probes are typical. The condensation
energy seems to peak at p = 0.19 [172] making superconductivity most robust slightly above optimal
doping.
Origin of superconductivity
The basic notions of the superconducting state are a d-wave gap and a universal dome-shaped dependence
of the transition temperature T
c
on doping p. Whenever p =0 is accessible the cuprates are AF insulators
highlighting the importance of strong electronic correlations. In the range 0 < p < 0.20 a competition
of ordering phenomena is observed. These facts should be captured at least qualitatively by a theoretical
approach towards superconductivity.
In conventional superconductors the condensation of electrons into Cooper pairs is an instability of the
normal metallic state. Cooper showed that an innitesimally weak attractive potential, V
0
, which is
non-zero only for energies
k
h
0
with h
0
the energy of the coupling boson (phonon in conventional
metals), makes two electrons with opposite momenta and spins living above a lled Fermis sphere,

k
=
k
E
F
> 0, to pair and to reduce their energy by [273]. Bardeen, Cooper and Schrieffer (BCS)
derived how N = O(10
23
) electrons can exploit this energy gain by forming a condensate which is
characterized by a single wave function similar to that of an electron in an isolated atom or an innite
plane wave [2]. The energy gain and the transition temperature T
c
depend linearly on the cutoff h
0
and exponentially on the coupling strength = N
F
V
0
with N
F
the density of electronic states at E
F
.
In the BCS approximation, 1, there is no direct relationship to real materials. This open problem
was solved by Eliashberg who showed how can be derived from the phonon spectrum and reach
values in excess of 1 [3, 274]. Coupling spectra
2
()F() with
2
() the energy dependent electron-
phonon interaction and F() the phonon density of states have been derived for elements and alloys
from electron tunneling spectra [275]. Since F() can be measured directly by neutron scattering
and F can be derived independently and serve as a basis for a quantitative comparison with theoretical
predictions [276] and, therefore, provide key information for a microscopic understanding of the pairing.
Generalizations including momentum dependent coupling and bosonic excitations other than phonons
have been put forward but require various approximations (similary as Eliashbergs original approach)
[151, 218, 250, 277, 278]. In all cases the characteristic bosonic energy must be much smaller than the
electronic energies, h
0
E
F
, [279] implying that the interaction is retarded. This means that one deals
with two fairly different time scales. The electronic one reacts instantaneously to a perturbation. The
other one maintains the polarization eld created by one electron sufciently long so as to allow a second
electron to experience it. This condition holds excellently in conventional metals having h
0
/E
F
=
O(10
2
).
In the beginning (see section 6.3.1) strong electron phonon coupling with > 1 was considered to lead
to sufciently stable Cooper pairing in the cuprates. The limiting case is the formation of polarons
which, at low temperature, condense into bi-polarons [61]. Polaronic behavior was indeed observed at
low doping [280, 281] but it is hard to pin down at optimal doping and beyond. This does not imply
that electron-phonon coupling can be disregarded. Actually, there are experimental indications such as
doping dependent shifts in the phonon spectra [282, 283], isotope effects at low doping [284], kinks in
the electronic dispersion in the entire doping range [146, 147, 149] or strong coupling effects of specic
phonons [285, 286]. However, the derived overall coupling constants are considered to be too small to
support superconductivity in the 100 K range [62, 250], and the way the electron-lattice interaction enters
is probably different from the situation in conventional superconductors [62, 143].
2013
110 R. HACKL AND D. EINZEL Unconventional Materials
Coupling mechanisms other than phononic can arise from low-energy spin [?, ?, 210, 287290] or charge
uctuations [153] or from high-energy instantaneous interactions such as the Coulomb repulsion U or
the exchange coupling J [230, 291, 292]. In all cases the Hubbard model is a useful starting point that
predicts antiferromagnetism, phase separation and superconductivity [29, 133] as shown in Fig. 6.17.
t
e
r
antiferromagnetic
superconducting
a
r
a
m
e
t
superconducting
r
d
e
r

p
a
AF+SC pseudogap SC
o
r
doping p
Figure 6.17: Phases predicted by the 3-band Hubbard model on the p-doped side. Similar to the 1-band
version (oxygen orbitals not explicitly taken into account) antiferromagnetic (AF) and superconducting
(SC) correlations corresponding to nite order parameters are found. In the overlap region a pseudogap
appears in the derived spectral functions. SC vanishes only at p =0. Additional crossover lines are found
at higher doping if electron-phonon interaction is included [142, 143]. From [133] with permission.
For t <U superconductivity can be obtained even though the interaction is repulsive, since the d-
wave gap changes sign and facilitates a solution of the gap equation. In a real-space argument one would
say that the two electrons avoid the repulsive part of U by arranging in a d-wave pair function which
vanishes when the potential is repulsive [29, 133] as already pointed out by Pitaevskii and Brueckner
et al. [16, 17]. This type of interaction is instantaneous since it is purely electronic and on a very high
energy or short time scale [292].
In the limit U the Coulomb repulsion is integrated out and J 130meV becomes the highest en-
ergy scale right after the band width. For p 0.03 the nearest-neighbor coupling J leads to the usual
Heisenberg-type long-ranged AF order (given that there is nite coupling in c-direction). At higher
doping only short range order and paramagnetism survive. There are indications that the coupling be-
tween 2 spins survives beyond optimal doping [107]. On this basis Anderson formulated the RVB ap-
proach [230], where local singlet pairs start to couple well above T
c
via the exchange energy J and
condense into Cooper pairs below T
c
. Then, the pseudogap is the energy reduction in the RVB state and
phase uctuations prevent the singlets from condensing above T
c
.
Upon proceeding to lower energies the pairing becomes more conventional in the sense that the inter-
action is retarded. Then, the Eliashberg theory can be applied [3, 274, 277], and a coupling spectrum
should be derivable from any type of electronic response or, turning the argument around, observable
by an appropriate independent experimental method such as neutron scattering in the case of spin uc-
tuations. There are quantitative studies of how the spin spectrum could provide enough coupling for
the cuprates [152] but there is no consensus yet since the interaction between a spin uctuation and an
electron can be treated only phenomenologically. Alternatively, the electrons can also interact via uc-
tuations of orbital currents [155, 156] or of the charge density [153, 154]. Small magnetic moments as
possible indications of orbital currents were discovered recently below T

[?] but it is as complicated as


in the case of spins to determine the coupling. Traces of charge uctuations are even harder to pin down
since there is no independent probe for the related excitations [114]. Caprara et al. propose to study the
c _ Walther-Meiner-Institut
Copper-oxygen compounds SUPERCONDUCTIVITY 111
related Aslamazov-Larkin uctuations by light scattering [293]. Since charge uctuations couple to the
lattice, the phonons may be back in the game in an indirect fashion [142].
Presently, spectra measured with different methods are analyzed in order to nd ngerprints of the
relevant bosons. These include ARPES [146, 147, 151, 152], neutron scattering [152], STS [138],
IR [218, 294296], and Raman spectroscopy [?, 139, 184]. There are indications for both electron-
phonon [146, 147] and electron-spin [151] interaction in the nodal ARPES spectra. With adjusted cou-
pling constants the nodal electron dispersion can be reproduced by and large on the basis of the spin sus-
ceptibility [152]. Away from the node high-resolution ARPES data do not exist yet. The optical and Ra-
man spectra always show two well separated energy scales at approximately 50 and 200 meV [184, 296]
which cannot a priori be identied with specic excitation. Possibly the strong polarization dependence
in the Raman results [184] may help to assign the modes via the selection rules. Then charge and spin
uctuations would dominate at low and high energies, respectively [?] making the exchange coupling
J an important player, as suggested by Anderson [292], along with coupled charge-phonon excitations.
However, as in the other cases the experiments have been done above T
c
and the coupling constants can
at best be guessed. Is there any independent criterion to foster a decision?
Poilblanc and Scalapino derived a partial sum rule for the complex Eliashberg gap function (k, ),
I(k, ) = f
k
()[
()
, varying between 0 for = 0 and approximately 1 for which measures
the contributions to the pairing interaction at a given momentum k as a function of the cut-off energy
[297]. When all the coupling (attractive or repulsive) is exhausted at high energies I(k, ) approaches
1. In Pb for instance, I() increases most rapidly at the transversal and longitudinal phonon frequencies,
exceeds unity above the energy range of the phonons due to the unretarded repulsion in normal metals
and approaches 1 asymptotically in the high energy limit [298]. This type of analysis requires high
resolution data for the gap function (k, ) which do not exist for the cuprates yet. However, theoretical
models can be studied, and in the Hubbard model 80 % of the coupling occurs in the energy range of
the spin uctuations. In contrast to Pb, I() does not exceed unity in the Hubbard model indicating
instantaneous pairing interactions at higher energies such as J and U. Their relative weight has still to be
claried [133, 298300].
It will be an important step forward if the relevant interactions or energy scales in the cuprates can be
identied. In spite of enormous progress both experimental and theoretical, the main question as to the
relative contribution of the various possible pairing mechanisms is not yet answered. Probably, it is the
right mixture which allows one to explain not only the phase diagram but also the material dependence.
It has been noticed that the maximal T
c
and the ratio t
/
/t depend in a systematic way on the distance of the
apical oxygen from the CuO
2
plane reecting the individual electronic structure of the compounds [301].
Johnston and coworkers derived the corresponding electron-phonon coupling
ph
and nd that T
c
can be
tuned substantially by varying the ratio
s
/
ph
with
s
the coupling via spin uctuations [62]. Similarly,
since the three-band Hubbard model includes p d charge uctuations which depend sensitively on
details of the materials it may supplement the plain vanilla one-band model [302] by adding a channel
for tuning T
max
c
[133].
In any case, we have to further sharpen our diagnostic tools to nally tackle the proper origin(s) of
superconductivity in the cuprates and, maybe, get ideas towards novel materials. Although new su-
perconductors have usually been found through the intuition of the materials scientists, the search was
always guided by concepts.
6.3.5 Summary and perspectives
The highest transition temperatures to superconductivity so far are observed in copper-oxygen com-
pounds with CuO
2
planes as the common building elements. The planes are separated by perovskite-like
2013
112 R. HACKL AND D. EINZEL Unconventional Materials
blocks. T
c
can reach 135 K under normal conditions and exceed 150 K with 25 GPa applied pressure.
The upper critical eld B
c2
(T 0) reaches values well above 100 T in compounds with T
c
100 K.
The cuprates share the proximity of superconductivity and other ordered phases or states with incipient
order with various materials such as f -electron systems, organic metals and the recently discovered iron-
based superconductors. It is a crucial question as to which extent superconductivity gets fueled by the
neighboring instabilities and their uctuations. Apparently, the dimensionality plays a role since low
dimensions favor uctuations and reduce the screening [303].
At present the Hubbard model seems a viable way towards a microscopic description since it captures
antiferromagnetism, competing phases and superconductivity. However, it is an open question which
of the interactions emerging from the model, i.e. spin uctuations, exchange coupling and Coulomb
repulsion, dominate in driving superconductivity. Even phonons may enter in a couple of ways yet
different from those in conventional systems.
So far neither theory nor experiment are in a state to suggest search strategies for new superconducting
materials. However, from what we know from the comparison of the cuprates, iron-pnictides and f -
electron systems, nearly planar systems at the brink of stability of a magnetic or charge-ordered phase
seem to be favorable.
After almost three decades of research into cuprates various applications emerge (see next chapter).
The most popular passive devices are frequency lters, fault current limiters, power transmission lines,
dynamical capacitors, high eld magnets, efcient eddy current heaters, and low-noise pick-up coils.
Presently, only SQUID magnetometers are among the cuprate-based applications using the dynamic
properties (Josephson effect). Although cycle frequencies in the THz range would be possible computing
with Josephson junctions appears very unlikely an application of the cuprates at the moment for the
enormous and continuous progress of semiconductor devices relying on established technology.
c _ Walther-Meiner-Institut
Chapter 7
An overview of applications
Materials and applications have a strong impact on the development of a eld since they generate funding
at a high rate and from many sources. The potential for applications of superconductivity was recognized
already by Kamerlingh-Onnes but it was a long way to go before the rst solenoid was presented in 1955
by Yntema [304], providing a eld of 0.71 T. The conductor was Nb, a weak type II superconductor
having 0.9. The development of the Ginzburg-Landau theory was one major step towards the un-
derstanding of the kind of problems that had to be solved for this application: high currents in high
elds. In addition, the metallurgy of superconducting materials, connections between superconductors
and between superconductors and normal metals, and efcient cooling at the temperature of liquid he-
lium were on the agenda. NbTi was a big leap since it has an upper critical eld of approximately 16 T
while maintaining a high ductility. It is still the material that dominates high eld applications. Only for
elds in excess of 12 T Nb
3
Sn [305] and YBa
2
Cu
3
O
7
conductors
1
are used and developed, respectively.
Apart from big magnetic coils high-power applications on a large scale are still at their infancy. On the
other hand active devices for detecting small voltages and magnetic elds are widely used and commer-
cially available. These include SQUID magnetometers for research, exploration and military, lters for
mobile telecommunication, and Josephson elements for metrology. To get superconducting technology
started the advantages must be striking, and other solutions have to be outperformed technically and
economically.
7.1 Potential areas
7.1.1 Economic considerations
By and large
2
new technologies prevail only if the production and operation costs are substantially
smaller and if the reliability is comparable or better in comparison to an established solution. In addition,
the inertia of running systems can hardly be overestimated. While the operation costs can be estimated
and controlled reasonably well the production costs may include a completely new infrastructure. A
well known example is hydrogen technology. Similarly, for power transmission over larger distances
with superconductors requiring cooling with liquid He the costs He recovery and liquefaction systems
are prohibitively high. Although operative cables were and are available there was no progress in instal-
lation before the cuprate superconductors permitted cooling with liquid nitrogen. Here, the liquefaction
and distribution is orders of magnitude simpler and cheaper than in the case of He. Consequently, also the
1
http://www.magnet.fsu.edu/usershub/publications/sciencehighlights/2010/28/0654118 Boebinger ASC1 coil nal.pdf
2
For details search for the Peter principle.
113
114 R. HACKL AND D. EINZEL An overview of applications
reliability is much higher. For the grid, the requirements are approximately 1/365 days for maintenance
as can be realized with nitrogen technology but not with He.
If, on the other hand, the operating costs exceed the those for installation new approaches are desirable.
As an example we compare the costs for a conventional and a superconducting magnet for a laboratory.
Similar considerations apply for a magnet used for magnetic resonance imaging (MRI) in medical ap-
plications or chemical analysis. Table 7.1 shows a comparison. This example shows that experiments in
Table 7.1: Comparison of the approximated overall cost of a laboratory magnet providing a eld of 10 T
in a bore of 50 mm. k e is 1,000 e, LHe is for liquid helium. In modern MRI machines there is no
LHe rell any more since the evaporating gas is reliqueed with a closed-cycle refrigerator consuming
approximately 5 kW corresponding to 40 e/d for electricity.
NbTi solenoid Cu solenoid
installation 100 k e 30 k e
infrastructure 0 10 k e
cooling 3 l LHe/d 1 m
3
water/min
30 e/d 3 k e
energy 0 5,000 kW
0 40,000 k e/d
maintenance LHe rell electricity/water
reliability 1/365 d ?
conductor 1 e(kAm)
1
0.2 e (kAm)
1
high elds were out of reach for standard laboratories before the advent of superconducting coils sim-
ilarly as medical examination using MRI. There is just one exception for which resistive magnets are
still needed: the generation of continuous elds above 21 T. The power consumption of the 45 T magnet
in the National High Field Laboratory in Tallahassee (FL, USA) or of the 33 T magnet in the CNRS
High Field Laboratory in Grenoble (France) is accordingly high and reaches values in the 30 MW range
requiring cooling facilities with capacities on the order of 1,000 m
3
per hour.
We may conclude that superconductivity will preferably be used for local applications since cooling of
extended installations is complicated and costly. However, with the advent of the cuprates the restrictions
become more relaxed since LHe can be replaced by liquid nitrogen (LN2) entailing a cost reduction by
approximately a factor of 100. For instance, cooling with LN2 costs approximately 0.5 e(md)
1
for a
high power cable. Nevertheless, the production costs for the conductors remain high.
7.1.2 Areas of application
It is clear that the so far available superconductors will not completely replace conventional materials.
Even if a material with a T
c
above room temperature will successfully be synthesized in the future prob-
lems such as ux pinning (see below), AC losses, connection with other materials, and, last but not least,
production costs will be an issue. On the basis of current technology the following applications are
realized or feasible. Some will be discussed in more detail below.
High current applications
This includes high-eld magnets for research, medicine, fusion, and accelerators, the improve-
ment of local grids, the interconnection of independent grids, fault current limiters, dynamic syn-
chronous condensers, motors, and generators.
c _ Walther-Meiner-Institut
Passive applications SUPERCONDUCTIVITY 115
Filter technology
Filter characteristics can be improved substantially with an enhanced conductivity. Here the dy-
namic conductivity in the MHz and predominantly GHz range is relevant. Using cuprates at
40. . . 77 K at least an order of magnitude improvement in performance and size can be reached
in comparison to copper. Here mobile telecommunication and defense prevail.
Sensors
SQUID magnetometers are widely used in research and development but also in Geology for ex-
ploration. The Josephson effect is used for metrology for instance.
Superconducting computers
In the beginning one was thinking of superconducting logic circuits switching between zero and
nite resistance. This concept was outperformed by Si technology before any realization. The
rapid single ux quantum logics (RSFQ) is still competitive concerning frequency but is not pur-
sued any further. Currently SQUIDS are being studied as to whether they can be used as Qbits in
a quantum computer.
Prior to all applications many physical concepts had to be developed and technical challenges had to be
solved. We shall discuss the important aspects before providing an overview of realized applications in
a eld. A broad discussion will follow in the lecture Applied Superconductivity in the Summer Term.
7.2 Passive applications
In most passive applications one wants to exploit the dissipation free transport of high currents sometimes
in a magnetic eld. In type I superconductors the critical currents and elds are very small limiting the
range of applications to the screening of small static elds and electromagnetic noise in the sub GHz
range ( h < where is the gap). In type II materials the elds destroying superconductivity are much
higher but in the presence of a transport current the vortices may move giving rise to dissipation. We
discuss no the origin of the dissipation and possible remedies.
7.2.1 Physical and technical challenges
If a transport current j
tr
ows in a superconductor in the presence of a magnetic eld as shown in Fig. 7.1
there is a force on the vortices originating from the Lorentz force. We saw in the discussion of the
Abrikosov lattice that the vortices assume a symmetric arrangement without a transport current since the
net forces vanish. For a nite transport current perpendicular to the eld B the symmetry is broken since
the transport and the screening currents add vectorially, j(r)
total
= j(r)
tr
+j(r)
screen
resulting in a net
projection of the currents around the ux lines on the y-direction and a force F
L
in positive x-direction.
The symmetry breaking can also be seen in a different way: the applied current j
tr
creates a eld around
the superconductor which enhances the eld on the l.h.s. and leads to a reduction of B on the opposite
side. As a consequence the vortex density starts to become location dependent thus establishing a force
to restore equilibrium. The density of the Lorentz force reads
f =
F
V
= nev
tr
B j
tr
B. (7.2.1)
If we integrate over the volume of one ux line, c
2
, we get
F
1
c
_

2
[j
tr
[[B[ (7.2.2)
2013
116 R. HACKL AND D. EINZEL An overview of applications
x
z
y
a
b
c
j
tr
B
F
L
Figure 7.1: Superconductor in a magnetic eld. The eld B that points in z-direction is in the range
B
c1
< B < B
c2
. The transport current j
tr
and the screening currents add resulting in an imbalance of the
screening currents around the ux lines. The net force on the ux lines originates in this asymmetry.
since most of the ux is in a cylinder having a radius of order , and since the current is perpendicular
to the eld. Inside the cylinder the force is approximately constant, and we arrive at the result derived
already earlier,
F
1
= c[j
tr
[
_

2
[B[d
2
r. (7.2.3)
or, more generally, F
1
/c = j
tr
e
z

0
since a ux line carries one ux quantum. The power dissipated by
one ux line is then
dQ
1
dt
=
d
dt
_
x
0
F
1
dx
/
F
1
dx
dt
= F
1
[v
FL
[ (7.2.4)
where v
FL
is the (unknown) drift velocity of the ux lines in x-direction. The last part of the derivation
is counting the number of ux lines N, and the total dissipation is P = N

Q
1
P =UI =
[B[ab

0
F
1
[v
FL
[ (7.2.5)
with U the voltage drop along y and I = ac [j
tr
[. Using Eq. (7.2.3) the voltage drop over the length b of
the superconductor can be expressed as
U = B[v
FL
[b, (7.2.6)
meaning that, for homogeneous conditions, the electric eld is E = B[v
FL
[. The drift velocity v
FL
can be
measured by virtue of Eq. (7.2.6). We can estimate [v
FL
[ by dening the maximally acceptable voltage
drop. Usually one denes the onset of the resistive state by occurrence of one V along the direction of
the transport current implying that in a eld of 1 T and and b = 10
2
m the drift velocity reaches only
[v
FL
[ = 10
4
m/s.
Whenever v
FL
,= 0 energy is dissipated. For understanding how the energy is dissipated we can follow
two ways. Qualitatively we observe that Cooper pair are broken up in the center of the vortex and turn
into normal particles. As long as the vortex is at rest the normal electrons do so as well making the mixed
state dissipationless after the relaxation time of the normal electrons. However, if the vortices move the
unpaired electrons are dragged on with velocity v
FL
and dissipate energy via collisions. Finally, vortices
appear on one side and get annihilated on the opposite side giving rise to an ac component in U. If the
spacing of the vortices is d the average voltage drop over the distance d is

U
d
=B[v
FL
[d, and a frequency
= 2
[v
FL
[
d
(7.2.7)
c _ Walther-Meiner-Institut
Passive applications SUPERCONDUCTIVITY 117
can be found. If we solve this for [v
FL
[ and substitute it into

U
d
we nd

U
d
=
Bd
2
2
. (7.2.8)
Bd
2
is the ux through one vortex that is just the ux quantum
0
, and we nd

U
d
=
h
2e
(7.2.9)
and recover the second Josephson equation. Very sloppily one can say that the current trajectory gets
interrupted by a vortex when it moves through and causes a phase slip of 2 that gives rise to a voltage
pulse.
The important question is as to whether or not the vortices can be prevented from moving since oth-
erwise the mixed state (Shubnikov phase) is next to useless for applications. We now discuss how the
pinning of the vortices can in fact be accomplished. The most direct way of describing pinning (see
V. V. Schmidt) is by realizing that a vortex has a normal core inside which the condensation energy van-
ishes implying that the energy density inside the core is enhanced by approximately B
2
c
/2
0
where B
c
is
the thermodynamic critical eld or by
B
2
c
2
0

2
(7.2.10)
per unit length over that of the superconducting material at distances larger than from the center of
the core. The same loss of condensation energy occurs if a hole with diameter 2 is drilled into the
superconductor. This can be realized experimentally by shooting accelerated heavy atoms on the super-
conductor. Now, if the vortex core is centered around this columnar defect, no additional condensation
energy is lost. If, however, the vortex is moved away from the hole the free energy of the material is
pushed up again by the same amount. Therefore a restoring force f
p
(per unit length) exists trying to pull
the vortex back into the cylindrical cavity. At the edge of the cavity f
p
should be approximately equal
to the condensation energy,
f
p

B
2
c
2
0
, (7.2.11)
yielding
f
p,c

B
2
c
2
0
c, (7.2.12)
for the slab of thickness c (see Fig. 7.1). Whenever f
p,c
<F
1
(see above) the restoring force is big enough
to the vortex pinned. If we equate f
p,c
and F
1
we nd for the current density
j
tr
=
B
2
c
2
0

0
, (7.2.13)
or, by using the GL expression (4.4.52) for the thermodynamic critical eld,
j
tr
=
B
c

2
0

, (7.2.14)
which is comparable to the pair-breaking critical current (not derived yet from the GL theory). In other
words, given efcient pinning a superconductor in in eld can carry a high current. However, the total
eld is substantially enhanced by the current. Therefore, the critical current is reduced substantially upon
approaching B
c2
.
2013
118 R. HACKL AND D. EINZEL An overview of applications
7.2.2 Examples
Conventional materials
The most popular application with the biggest market are solenoids for MRI, and many of us had an
opportunity to experience one of the machines. A eld of up to 3 T with well-dened small gradients in
a volume of approximately 1 m
3
is provided by a system of superconducting coils. The multilamentary
wires are made of NbTi embedded in a matrix of alloyed Cu. They can be fabricated by established
extrusion techniques and wound up [10]. Hydrogen is the only nucleus being probed here and the spatial
resolution is approximately 0.5 mm. Coils for 48 T are in the experimental phase since the resolution
can be improved and other nuclei such as phosphorus can be used as a probe.
The other big market is laboratory magnets. As mentioned above relatively inexpensive NbTi standard
laboratory magnets are in wide use. For chemical analysis high-eld NMR has an increasing share.
After a very long experimental phase Nb
3
Sn coils providing elds in excess of 21 T became available
in 1980ies and are standard for high-eld NMR magnets only since the nineties [306]. Here, hybrid
coils are used. The outer coil is a NbTi solenoid providing some 10 T the inner part is a Nb
3
Sn solenoid
adding another 10 or 11 T, and a eld 21 T can be reached in a bore of 50 mm having a homogeneity
of better than 10
6
in a volume of 1 cm
3
. Similarly, the 45 T magnet in Tallahassee is a hybrid having
a NbTi, a Nb
3
Sn, and a resistive coil with the latter one providing an additional eld of some 25 T at
the price of 30 MW electrical power and 1440 m
3
per hour cooling water. On of the biggest installations
of superconducting magnets is the accelerator of the CERN. In Karlsruhe coils for fusion reactors were
developed and built. The coils have complicated shapes and are as big as a two-story building. Since a
while prototypes for toroidal coils for energy storage are being developed.
Beyond the conventional metallic systems there is an increasing number of applications of high-T
c
cuprates but none of the established techniques can be used for the cuprates.
3
Cuprates with high transition temperature
It is not only the higher transition temperatures which relax the requirements for cooling but also the
enhanced robustness of superconductivity as quantied by the condensation energy F B
2
c
T
2
c
with
B
c
the thermodynamical critical eld. In Y-123 for instance, the upper critical eld B
c2
(T), at which
superconductivity collapses, is in the range 140 T in the low-T limit and still some 40 T at 77 K[241]. The
critical current densities exceed 100 Amm
2
and 10,000 Amm
2
at 77 and 4.2 K, respectively. For active
devices a switching frequency in the THz range can be reached for the large energy gap,
1
/ h.
Given these advantages, why took it more than 20 years until the rst application was commercialized?
It is rather complicated to get sufciently homogeneous large-scale products at competitive costs since
the materials have to be synthesized at 600-800 C and are brittle. Y-123, the workhorse in applications,
is a quaternary compound and small deviations from stoichiometry reduce T
c
substantially. On the other
hand, defects at average distances close to the coherence length
0
20

A are necessary for high critical
elds and currents. The enormous ac-anisotropy and the critical current of in-plane grain boundaries
which decreases exponentially with the misalignment angle [307] require essentially mono-crystalline
specimens. For coils or for power transmission lines the quality must be maintained over hundreds of
meters.
In spite of the difculties various products are in the test phase now (for recent overviews see, e.g.,
Ref. [308310] or the web-links in the references). Wires with Bi-2223 can be bought from the shelf
with lengths up to 1.5 km [311]. For this product Bi-2223 powder is lled in metal tubes, mostly Ag, and
3
The following section is an almost literal quote from a publication of one of us (R.H.) [56]
c _ Walther-Meiner-Institut
Active devices SUPERCONDUCTIVITY 119
then rolled and reheated several times to get the platelet-like micro-crystals aligned. Very good results are
obtained with oriented lms of Y-123, which are deposited on strained metallic ribbons on top of various
buffer layers [312, 313]. There are various techniques to grow thin lms.For industrial production costs
play a central role. Thermal co-evaporation of the metallic elements followed by in-situ post-annealing
in oxygen [93] is among the promising methods. It was developed in a spin-off company of the Technical
University Munich which now sells superconductors as well as production and test equipment [312]. For
some applications jet-printing is used to deposit liquid precursors which are processed afterwards [314].
In this way complicated geometries can be realized, possibly at the price of a slightly reduced critical
current. On the other hand, the technique is extremely simple and cost effective.
In wider collaborations fault current limiters have been developed. In contrast to conventional technology
the device can reversible interrupt the connection between the generator and the grid in the case of an
overload. Inductive fault current limiters can cut spikes without fully interrupting the transmission. The
rst device produced by Zenergy Power was delivered in 2010 to CE Electric in the United Kingdom.
In the United States dynamic synchronous condensers are used to compensate variable inductive and
capacitive loads resulting from and enhancing rapid uctuations of the power consumption in the grid.
Filters for base stations of mobile communication have a much better selectivity and are substantially
smaller. The superconducting lter element has an area of only a few quare centimeters. Thousands of
these lters which t into a 19
//
rack have been installed already [315]. Motors and generators having
rotors with superconducting coils have a slightly better efciency and substantially reduced size and
weight. This helps to reduce the material consumption and makes the technology favorable for, e.g.,
ships or upcoming applications such as wind turbines. It has been demonstrated that thin-lm Y-123
receiver coils improve the signal-to-noise ratio of MRI by a factors between 2 and 9 compared to that
achievable with copper [316] as shown in Fig. 7.2 At the National High Field Laboratory in Tallahasse
solenoids for the 30 to 40 T range are under development. For many experiments these magnets can
replace the 45 T Hybrid magnet with a superconducting outer and a conventional inner coil at a price of
approximately 20 MW power consumption.
All these applications are local, and cryogen-free cooling is possible and is actually used widely. The
development and optimization of pulse-tube cryo-coolers in the last decade simplied the refrigeration
substantially and improved the reliability. The maintenance intervals are years and the the base tem-
perature of a single stage engine is close to 30 K. For the lters miniature Stirling coolers have been
developed with a power consumption in the 100 W range [315].
For non-local application such as power transmission cooling is still an issue. It is very much relaxed
by the use of liquid nitrogen but still complicated and subject to failure [310]. Therefore, supercon-
ductors are mainly considered for specialized applications, where conventional techniques cannot carry
the increased load, although the development of cables is very advanced . Nevertheless, using a ca-
ble manufactured by AmSC, the Long Island Power Authority started to transmit electricity for 300,000
households in April 2008 as shown in Fig. 7.3 [313]. Demonstration projects have been started in various
other places of the world.
7.3 Active devices
The sensing technique with SQUIDS on the basis of conventional superconductors was already mature
at the end of the 1980ies while the use of cuprates operated at 77 K has still to overcome some prob-
lems [308]. Due to low pinning potentials the noise is the main problem to x. While the noise of
approximately 50 fTHz
1/2
at 1 Hz is still too high for magneto-encephalography the study of the heart
is feasible and has enough resolution [317]. There exist small start-up companies producing integrated
solutions [318]. SQUIDs are now used mainly in the laboratory but also in military applications, in ge-
ology, and in quality control. Bits for quantum computing are so far only made of conventional metals.
2013
120 R. HACKL AND D. EINZEL An overview of applications
tissue
05 m
B
0.5 m
c-shapedopen magnet
plastic
p p g
plastic
tube
He compressor He compressor
Figure 7.2: Schematic viewof a setup for magnetic resonance imaging (MRI). The c-shaped open magnet
is fabricated with Bi-2223 tape wire. The sensor has an Y-123 receiver. Magnet and receiver are cooled
cryogen free with closed-cycle He refrigerators. The lower image on the r.h.s. shows the MRI image of
the little nger with a conventional Cu receiver. The image on top is the result obtained with the Y-123
receiver. The improvement of the signal to noise ratio for the same scanning time is approximately a
factor of 5. Reproduced from [316].
Figure 7.3: Superconducting power transmission line (front) operated by the Long Island Power Author-
ity (LIPA). The cables have a core of Y-123 conductor including Cu for protection, a channel for liquid
nitrogen and thermal insulation. The superconductors carry the same energy as the over-head lines in the
background. The system is operative since April 2008. Courtesy of American Superconductors [313].
c _ Walther-Meiner-Institut
Bibliography
[1] H. Kamerlingh-Onnes, , Comm. Leiden 120b, (1911).
[2] J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Theory of Superconductivity, Phys. Rev. 108, 1175
(1957).
[3] G. M. Eliashberg, , Zh. Eksp. Teor. Fiz. 38, 966 (1960).
[4] A. J. Leggett, A theoretical description of the new phases of liquid
3
He, Rev. Mod. Phys. 47, 331
(1975).
[5] G. Gr uner, in Density waves in Solids, edited by D. Pines (Addison-Wesley, ADDRESS, 1994).
[6] J. G. Bednorz and K. A. M uller, Possible high T
c
superconductivity in the Ba-La-Cu-O System,
Z. Phys. B 64, 189 (1986).
[7] Y. Kamihara, H. Hiramatsu, M. Hirano, R. Kawamura, H. Yanagi, T. Kamiya, and H. Hosono,
Iron-Based Layered Superconductor: LaOFeP, J. Am. Chem. Soc.. 128, 10012 (2006).
[8] M. Rotter, M. Tegel, and D. Johrendt, Superconductivity at 38 K in the Iron Arsenide
(Ba
1x
K
x
)Fe
2
As
2
, Phys. Rev. Lett. 101, 107006 (2008).
[9] M. L uders, M. A. L. Marques, N. N. Lathiotakis, A. Floris, G. Profeta, L. Fast, A. Continenza,
S. Massidda, and E. K. U. Gross, Ab initio theory of superconductivity. I. Density functional
formalism and approximate functionals, Phys. Rev. B 72, 024545 (2005).
[10] W. Buckel and R. Kleiner, Superconductivity (Wiley, Weinheim 2004, Weinheim, 2004).
[11] P. Drude, Zur Elektronentheorie der Metalle, Annalen der Physik 306, 566 (1900).
[12] P. W. Higgs, Broken Symmetries and the Masses of Gauge Bosons, Phys. Rev. Lett. 13, 508
(1964).
[13] F. Gross, B. S. Chandrasekhar, D. Einzel, K. Andres, P. J. Hirschfeld, H. R. Ott, J. Beuers, Z. Fisk,
and J. L. Smith, Anomalous temperature dependence of the magnetic eld penetration depth in
superconducting UBe
13
, Z. Phys. B: Condens. Matter 64, 175 (1986).
[14] N. Ashcroft and N. Mermin, Solid State Physics (Saunder College, Philadelphia, ADDRESS,
1976), p. 848.
[15] W. Prestel, Study of the Interaction Processes in Cuprate Superconductors by a Quantitative Com-
parison of Spectroscopic Experiments, Ph.d. thesis, Technische Universit at M unchen, 2012.
[16] K. A. Brueckner, T. Soda, P. W. Anderson, and P. Morel, Level Structure of Nuclear Matter and
Liquid
3
He, Phys. Rev. 118, 1442 (1960).
121
122 R. HACKL AND D. EINZEL BIBLIOGRAPHY
[17] L. P. Pitaevskii, , Sov. Phys. JETP 10, 1267 (1960).
[18] D. A. Wollman, D. J. Van Harlingen, W. C. Lee, D. M. Ginsberg, and A. J. Leggett, Experimental
determination of the superconducting pairing state in YBCO from the phase coherence of YBCO-
Pb dc SQUIDs, Phys. Rev. Lett. 71, 2134 (1993).
[19] C. C. Tsuei and J. R. Kirtley, Pairing symmetry in cuprate superconductors, Rev. Mod. Phys. 72,
969 (2000).
[20] W. N. Hardy, D. A. Bonn, D. C. Morgan, R. Liang, and K. Zhang, Precision measurements of the
temperature dependence of in YBa
2
Cu
3
O
6.95
: Strong evidence for nodes in the gap function,
Phys. Rev. Lett. 70, 3999 (1993).
[21] B. M uhlschlegel, Die thermodynamischen Funktionen des Supraleiters, Z. Phys. 155, 313 (1959).
[22] D. Einzel, Supraleitung und Superuidit at, Lexikon der Physik (2000).
[23] H. F. Hess, R. B. Robinson, R. C. Dynes, J. M. Valles, and J. V. Waszczak, Scanning-Tunneling-
Microscope Observation of the Abrikosov Flux Lattice and the Density of States near and inside
a Fluxoid, Phys. Rev. Lett. 62, 214 (1989).
[24] P. G. de Gennes, Superconductivity of Metals and Alloys (W. A. Benjamin, ADDRESS, 1964),
iSBN 0-7382-0101-4.
[25] K. Andres, J. E. Graebner, and H. R. Ott, 4 f -Virtual-Bound-State Formation in CeAl
3
at Low
Temperatures, Phys. Rev. Lett. 35, 1779 (1975).
[26] F. Steglich, J. Aarts, C. D. Bredl, W. Lieke, D. Meschede, W. Franz, and H. Sch afer, Supercon-
ductivity in the Presence of Strong Pauli Paramagnetism: CeCu
2
Si
2
, Phys. Rev. Lett. 43, 1892
(1979).
[27] C. Peiderer, Superconducting phases of f -electron compounds, Rev. Mod. Phys. 81, 1551
(2009).
[28] M. Vojta, Quantum phase transitions, Rep. Prog. Phys. 66, 2069 (2003).
[29] D. J. Scalapino, The case for d
x2y2
pairing in the cuprate superconductors, Physics Reports 250,
329 (1995).
[30] I. I. Mazin, D. J. Singh, M. D. Johannes, and M. H. Du, Unconventional Superconductivity with a
Sign Reversal in the Order Parameter of LaFeAsO
1x
F
x
, Phys. Rev. Lett. 101, 057003 (2008).
[31] Y. Kamihara, T. Watanabe, M. Hirano, and H. Hosono, Communication Iron-Based Layered Su-
perconductor La[O
1x
F
x
]FeAs (x = 0.05-0.12) with T
c
= 26 K, J. Am. Chem. Soc. 130, 3296
(2008).
[32] A. I. Coldea, J. D. Fletcher, A. Carrington, J. G. Analytis, A. F. Bangura, J.-H. Chu, A. S. Erickson,
I. R. Fisher, N. E. Hussey, and R. D. McDonald, Fermi Surface of Superconducting LaFePO
Determined from Quantum Oscillations, Phys. Rev. Lett. 101, 216402 (2008).
[33] J. G. Analytis, C. M. J. Andrew, A. I. Coldea, A. McCollam, J.-H. Chu, R. D. McDonald, I. R.
Fisher, and A. Carrington, Fermi Surface of SrFe
2
P
2
Determined by the de Haasvan Alphen
Effect, Phys. Rev. Lett. 103, 076401 (2009).
[34] A. S. Sefat, R. Jin, M. A. McGuire, B. C. Sales, D. J. Singh, and D. Mandrus, Superconductivity
at 22 K in Co-Doped BaFe
2
As
2
Crystals, Phys. Rev. Lett. 101, 117004 (2008).
c _ Walther-Meiner-Institut
BIBLIOGRAPHY SUPERCONDUCTIVITY 123
[35] F.-C. Hsu, J.-Y. Luo, K.-W. Yeh, T.-K. Chen, T.-W. Huang, P. M. Wu, Y.-C. Lee, Y.-L. Huang,
Y.-Y. Chu, D.-C. Yan, and M.-K. Wu, Superconductivity in the PbO-type structure -FeSe, PNAS
105, 14262 (2008).
[36] Y. Mizuguchi, F. Tomioka, S. Tsuda, T. Yamaguchi, and Y. Takano, Superconductivity at 27 K in
tetragonal FeSe under high pressure, Appl. Phys. Lett. 93, 152505 (2008).
[37] J.-H. Chu, J. G. Analytis, C. Kucharczyk, and I. R. Fisher, Determination of the phase diagram of
the electron-doped superconductor Ba(Fe
1x
Co
x
)
2
As
2
, Phys. Rev. B 79, 014506 (2009).
[38] H. Luetkens, H.-H. Klauss, M. Kraken, F. J. Litterst, T. Dellmann, R. Klingeler, C. Hess, R.
Khasanov, A. Amato, C. Baines, M. Kosmala, O. J. Schumann, M. Braden, J. Hamann-Borrero,
N. Leps, A. Kondrat, G. Behr, J. Werner, and B. Buchner, The electronic phase diagram of the
LaO
1x
F
x
FeAs superconductor, Nat. Mater. 8, 305 (2009).
[39] A. V. Chubukov, D. V. Efremov, and I. Eremin, Magnetism, superconductivity, and pairing sym-
metry in iron-based superconductors, Phys. Rev. B 78, 134512 (2008).
[40] S. Graser, T. Maier, P. Hirschfeld, and D. Scalapino, Near-degeneracy of several pairing channels
in multiorbital models for the Fe pnictides, New J. Phys. 11, 025016 (2009).
[41] F. Wang, H. Zhai, and D.-H. Lee, Antiferromagnetic correlation and the pairing mechanism of the
cuprates and iron pnictides: A view from the functional renormalization group studies, Europhys.
Lett. 85, 37005 (2009).
[42] L. Boeri, O. V. Dolgov, and A. A. Golubov, Is LaFeAsO
1x
F
x
an Electron-Phonon Superconduc-
tor?, Phys. Rev. Lett. 101, 026403 (2008).
[43] M. M. Qazilbash, J. J. Hamlin, R. E. Baumbach, L. Zhang, D. J. Singh, M. B. Maple, and D. N.
Basov, Electronic correlations in the iron pnictides, Nat. Phys. 5, 647 (2009).
[44] D. H. Lu, M. Yi, S.-K. Mo, A. S. Erickson, J. Analytis, J.-H. Chu, D. J. Singh, Z. Hussain,
T. H. Geballe, I. R. Fisher, and Z.-X. Shen, Electronic structure of the iron-based superconductor
LaOFeP, Nature 455, 81 (2008).
[45] E. Z. Kurmaev, R. G. Wilks, A. Moewes, N. A. Skorikov, Y. A. Izyumov, L. D. Finkelstein, R. H.
Li, and X. H. Chen, X-ray spectra and electronic structures of the iron arsenide superconductors
RFeAsO
1x
F
x
(R = La,Sm), Phys. Rev. B 78, 220503 (2008).
[46] K. Terashima, Y. Sekiba, J. H. Bowen, K. Nakayama, T. Kawahara, T. Sato, P. Richard, Y.-M. Xu,
L. J. Li, G. H. Cao, Z.-A. Xu, H. Ding, and T. Takahashi, Fermi surface nesting induced strong
pairing in iron-based superconductors, PNAS 106, 7330 (2009).
[47] P. J. Hirschfeld, M. M. Korshunov, and I. I. Mazin, Gap symmetry and structure of Fe-based
superconductors, Rep. Prog. Phys. 74, 125508 (2011).
[48] A. D. Christianson, E. A. Goremychkin, R. Osborn, S. Rosenkranz, M. D. Lumsden, C. D. Malli-
akas, I. S. Todorov, H. Claus, D. Y. Chung, M. G. Kanatzidis, R. I. Bewley, and T. Guidi, Un-
conventional superconductivity in Ba
0.6
K
0.4
Fe
2
As
2
from inelastic neutron scattering, Nature 456,
930 (2008).
[49] D. V. Evtushinsky, D. S. Inosov, V. B. Zabolotnyy, A. Koitzsch, M. Knupfer, B. Buchner, M. S.
Viazovska, G. L. Sun, V. Hinkov, A. V. Boris, C. T. Lin, B. Keimer, A. Varykhalov, A. A. Kordyuk,
and S. V. Borisenko, Momentumdependence of the superconducting gap in Ba
1x
K
x
Fe
2
As
2
, Phys.
Rev. B 79, 054517 (2009).
2013
124 R. HACKL AND D. EINZEL BIBLIOGRAPHY
[50] P. Samuely, Z. Pribulov a, P. Szab o, G. Prist a s, S. L. Budko, and C. P.C., Point contact Andreev re-
ection spectroscopy of superconducting energy gaps in 122-type family of iron pnictides, Physica
C 469, 507 (2009).
[51] B. Muschler, W. Prestel, R. Hackl, T. P. Devereaux, J. G. Analytis, J.-H. Chu, and I. R. Fisher,
Band- and momentum-dependent electron dynamics in superconducting Ba(Fe
1x
Co
x
)
2
As
2
as
seen via electronic Raman scattering, Phys. Rev. B 80, 180510 (2009).
[52] M. A. Tanatar, N. Ni, C. Martin, R. T. Gordon, H. Kim, V. G. Kogan, G. D. Samolyuk,
S. L. Budko, P. C. Caneld, and R. Prozorov, Anisotropy of the iron pnictide superconductor
Ba(Fe
1x
Co
x
)
2
As
2
(x = 0.074, T
c
= 23 K), Phys. Rev. B 79, 094507 (2009).
[53] J. D. Fletcher, A. Seran, L. Malone, J. G. Analytis, J.-H. Chu, A. S. Erickson, I. R. Fisher, and A.
Carrington, Evidence for a Nodal-Line Superconducting State in LaFePO, Phys. Rev. Lett. 102,
147001 (2009).
[54] C. W. Hicks, T. M. Lippman, M. E. Huber, J. G. Analytis, J.-H. Chu, A. S. Erickson, I. R. Fisher,
and K. A. Moler, Evidence for a Nodal Energy Gap in the Iron-Pnictide Superconductor LaFePO
from Penetration Depth Measurements by Scanning SQUID Susceptometry, Phys. Rev. Lett. 103,
127003 (2009).
[55] L. Fang, H. Luo, P. Cheng, Z. Wang, Y. Jia, G. Mu, B. Shen, I. I. Mazin, L. Shan, C. Ren, and
H.-H. Wen, Roles of multiband effects and electron-hole asymmetry in the superconductivity and
normal-state properties of Ba(Fe
1x
Co
x
)
2
As
2
, Phys. Rev. B 80, 140508 (2009).
[56] R. Hackl, Superconductivity in copper-oxygen compounds, Z. Kristallogr. 226, 323 (2011).
[57] C. Varma, in Superconductivity in d- and f-Band Metals, edited by W. Buckel and W. Weber
(Kernforschungszentrum Karlsruhe, ADDRESS, 1982), p. 500.
[58] W. A. Little, Possibility of Synthesizing an Organic Superconductor, Phys. Rev. 134, A1416
(1964).
[59] D. Allender, J. W. Bray, and J. Bardeen, Theory of uctuation superconductivity from electron-
phonon interactions in pseudo-one-dimensional systems, Phys. Rev. B 9, 119 (1974).
[60] D. Allender, J. Bray, and J. Bardeen, Model for an Exciton Mechanism of Superconductivity,
Phys. Rev. B 7, 1020 (1973).
[61] A. Alexandrov and J. Ranninger, Bipolaronic superconductivity, Phys. Rev. B 24, 1164 (1981).
[62] D. Johnston, The Puzzle of High Temperature Superconductivity in Layered Iron Pnictides and
Chalcogenides , Adv. Phys. 59, 803 (2010).
[63] D. T. Jover, R. J. Wijngaarden, H. Wilhelm, R. Griessen, S. M. Loureiro, J.-J. Capponi, A.
Schilling, and H. R. Ott, Pressure dependence of the superconducting critical temperature of
HgBa
2
Ca
2
Cu
3
O
8+y
and HgBa
2
Ca
3
Cu
4
O
10+y
up to 30 GPa, Phys. Rev. B 54, 4265 (1996).
[64] Y. Ando, S. Komiya, K. Segawa, S. Ono, and Y. Kurita, Electronic Phase Diagram of High-T
c
Cuprate Superconductors from a Mapping of the In-Plane Resistivity Curvature, Phys. Rev. Lett.
93, 267001 (2004).
[65] R. J. Cava, B. Batlogg, R. B. van Dover, D. W. Murphy, S. Sunshine, T. Siegrist, J. P. Remeika,
E. A. Rietman, S. Zahurak, and G. P. Espinosa, Bulk superconductivity at 91 K in single-phase
oxygen-decient perovskite Ba
2
YCu
3
O
9
, Phys. Rev. Lett. 58, 1676 (1987).
c _ Walther-Meiner-Institut
BIBLIOGRAPHY SUPERCONDUCTIVITY 125
[66] N. P. Armitage, P. Fournier, and R. L. Greene, Progress and perspectives on electron-doped
cuprates, Rev. Mod. Phys. 82, 2421 (2010).
[67] Y. Tokura, H. Takagi, and S. Uchida, A superconducting copper oxide compound with electrons
as the charge carriers, Nature 337, 345 (1989).
[68] M. K. Wu, J. R. Ashburn, C. J. Torng, P. H. Hor, R. L. Meng, L. Gao, Z. J. Huang, Y. Q. Wang,
and C. W. Chu, Superconductivity at 93 K in a new mixed-phase Y-Ba-Cu-O compound system at
ambient pressure, Phys. Rev. Lett. 58, 908 (1987).
[69] H. Shaked, P. M. Kean, J. C. Rodriguez, F. F. Owen, R. L. Hitterman, and J. D. Jorgensen, Crystal
Structures of the High-Tc Superconducting Copper-Oxides (Elsevier Science B.V., ADDRESS,
1994), p. 71.
[70] J. Longo and P. Raccah, The structure of La
2
CuO
4
and LaSrVO
4
, J. Solid State Chem. 6, 526
(1973).
[71] B. Grande, H. M uller-Buschbaum, and M. Schweizer,

Uber Oxocuprate. XV Zur Kristallstruktur
von Seltenerdmetalloxocupraten: La
2
CuO
4
, Gd
2
CuO
4
, ZAAC 428, 120 (1977).
[72] H. M uller-Buschbaum and Wollschl ager,

Uber tern are Oxocuprate. VII. Zur Kristallstruktur von
Nd
2
CuO
4
, Z. Anorg. Allg. Chem. 414, 76 (1975).
[73] H. M uller-Buschbaum, Zur Kristallchemie der oxidischen Hochtemperatur-Supraleiter und deren
kristallchemischen Verwandten, Angew. Chem. 101, 1503 (1989).
[74] M. A. Kastner, R. J. Birgeneau, G. Shirane, and Y. Endoh, Magnetic, transport, and optical prop-
erties of monolayer copper oxides, Rev. Mod. Phys. 70, 897 (1998).
[75] A. Erb, E. Walker, and R. Fl ukiger, BaZrO3: The solution for the crucible corrosion problem
during the single crystal growth of high-T
c
superconductors REBa
2
Cu
3
O
7
; RE = Y, Pr, Physica
C 245, 245 (1995).
[76] A. Erb, E. Walker, and R. Fl ukiger, The use of BaZrO3 crucibles in crystal growth of the high-Tc
superconductors. Progress in crystal growth as well as in sample quality, Physica C 258, 9 (1996).
[77] R. Liang, D. A. Bonn, and W. N. Hardy, Growth of high quality YBCO single crystals using
BaZrO
3
crucibles, Physica C 304, 105 (1998).
[78] R. Liang, D. A. Bonn, and W. N. Hardy, Preparation and X-ray characterization of highly ordered
ortho-II phase YBa2Cu3O6.50 single crystals, Physica C 336, 57 (2000).
[79] R. Liang, D. A. Bonn, W. N. Hardy, J. C. Wynn, K. A. Moler, L. Lu, S. Larochelle, L. Zhou, M.
Greven, L. Lurio, and S. G. J. Mochrie, Preparation and characterization of homogeneous YBCO
single crystals with doping level near the SC-AFM boundary, Physica C 383, 1 (2002).
[80] R. Liang, D. A. Bonn, and W. N. Hardy, Evaluation of CuO
2
plane hole doping in YBa
2
Cu
3
O
6+x
single crystals, Phys. Rev. B 73, 180505 (2006).
[81] M. Kl aser, YBa
2
Cu
3
O
x
-Einkristalle dotiert mit Kalzium: Zchtung und Charakterisierung, Ph.D.
thesis, Universitt Karlsruhe, 1999.
[82] R. E. Gladyshevskii and R. Fl ukiger, Modulated structure of Bi
2
Sr
2
CaCu
2
O
8+
, a high-T
c
super-
conductor with monoclinic symmetry, Acta Crystallogr., Sect. B: Struct. Sci 52, 38 (1996).
2013
126 R. HACKL AND D. EINZEL BIBLIOGRAPHY
[83] H. Eisaki, N. Kaneko, D. L. Feng, A. Damascelli, P. K. Mang, K. M. Shen, Z.-X. Shen, and M.
Greven, Effect of chemical inhomogeneity in bismuth-based copper oxide superconductors, Phys.
Rev. B 69, 064512 (2004).
[84] C. Kendziora, M. C. Martin, J. Hartge, L. Mihaly, and L. Forro, Wide-range oxygen doping of
Bi
2
Sr
2
CaCu
2
O
8+
, Phys. Rev. B 48, 3531 (1993).
[85] G. Triscone, J. Y. Genoud, T. Graf, A. Junod, and J. Muller, Variation of the superconducting
properties of Bi
2
Sr
2
CaCu
2
O
8+x
with oxygen content, Physica C 176, 247 (1991).
[86] N. Bari si c, Y. Li, X. Zhao, Y.-C. Cho, G. Chabot-Couture, G. Yu, and M. Greven, Demonstrating
the model nature of the high-temperature superconductor HgBa
2
CuO
4+
, Phys. Rev. B78, 054518
(2008).
[87] A. P. Mackenzie, S. R. Julian, G. G. Lonzarich, A. Carrington, S. D. Hughes, R. S. Liu, and
D. S. Sinclair, Resistive upper critical eld of Tl
2
Ba
2
CuO
6
at low temperatures and high magnetic
elds, Phys. Rev. Lett. 71, 1238 (1993).
[88] D. Peets, R. Liang, M. Raudsepp, W. Hardy, and D. Bonn, Encapsulated single crystal growth and
annealing of the high-temperature superconductor Tl-2201, J. Cryst. Growth 312, 344 (2010).
[89] M. Lambacher, Crystal growth and normal state transport of electron doped high temperature
superconductors, Dissertation, Technische Universit at M unchen, 2008.
[90] M. Lambacher, T. Helm, M. Kartsovnik, and A. Erb, Advances in single crystal growth and an-
nealing treatment of electron-doped HTSC, Eur. Phys. J. Special Topics 188, 61 (2010).
[91] X. Zhao, G. Yu, Y.-C. Cho, G. Chabot-Couture, N. Barisic, P. Bourges, N. Kaneko, Y. Li, L.
Lu, E. Motoyama, O. Vajk, and M. Greven, Crystal Growth and Characterization of the Model
High-Temperature Superconductor HgBa
2
CuO
4+
, Adv. Mater. 18, 3243 (2006).
[92] I. Bozovic, Atomic-layer engineering of superconducting oxides: yesterday, today, tomorrow,
Appl. Supercond. 11, 2686 (2001).
[93] P. Berberich, W. Assmann, W. Prusseit, B. Utz, and H. Kinder, Large area deposition of
YBa
2
Cu
3
O
7
lms by thermal co-evaporation, J. Alloys Compd. 195, 271 (1993).
[94] G. Logvenov, A. Gozar, and I. Bozovic, High-Temperature Superconductivity in a Single Copper-
Oxygen Plane, Science 326, 699 (2009).
[95] J. L. Tallon, C. Bernhard, H. Shaked, R. L. Hitterman, and J. D. Jorgensen, Generic superconduct-
ing phase behavior in high-T
c
cuprates: T
c
variation with hole concentration in YBa
2
Cu
3
O
7
,
Phys. Rev. B 51, 12911 (1995).
[96] M. R. Norman, D. Pines, and C. Kallin, The pseudogap: friend or foe of high T
c
?, Adv. Phys. 54,
715 (2005).
[97] J. Tranquada, in Handbook of High-Temperature Superconductivity Theory and Experiment,
edited by J. Schrieffer and J. Brooks (Springer, Berlin-Heidelberg, 2007).
[98] R. Presland, M. L. Tallon, J. G. Buckley, R. S. Liu, R. and E. Flower, N. General trends in
oxygen stoichiometry effects on Tc in Bi and Tl superconductors, Physica C 176, 95 (1991).
[99] S. H. Naqib, J. R. Cooper, J. L. Tallon, R. S. Islam, and R. A. Chakalov, Doping phase diagram of
Y
1x
Ca
x
Ba
2
(Cu
1y
Zn
y
)
3
O
7
from transport measurements: Tracking the pseudogap below T
c
,
Phys. Rev. B 71, 054502 (2005).
c _ Walther-Meiner-Institut
BIBLIOGRAPHY SUPERCONDUCTIVITY 127
[100] F. Rullier-Albenque, H. Alloul, F. Balakirev, and C. Proust, Disorder, metal-insulator crossover
and phase diagram in high-T
c
cuprates, Europhys. Lett. 81, 37008 (2008).
[101] H. Alloul, J. Bobroff, M. Gabay, and P. J. Hirschfeld, Defects in correlated metals and supercon-
ductors, Rev. Mod. Phys. 81, 45 (2009).
[102] R. S. Islam, J. R. Cooper, J. W. Loram, and S. H. Naqib, Pseudogap and doping-dependent mag-
netic properties of La
2x
Sr
x
Cu
1y
Zn
y
O
4
, Phys. Rev. B 81, 054511 (2010).
[103] S. Ono and Y. Ando, Evolution of the resistivity anisotropy in Bi
2
Sr
2x
La
x
CuO
6+
single crystals
for a wide range of hole doping, Phys. Rev. B 67, 104512 (2003).
[104] R. Harris, P. J. Turner, S. Kamal, A. R. Hosseini, P. Dosanjh, G. K. Mullins, J. S. Bobowski, C. P.
Bidinosti, D. M. Broun, R. Liang, W. N. Hardy, and D. A. Bonn, Phenomenology of a[over] -axis
and b[over] -axis charge dynamics from microwave spectroscopy of highly ordered YBa
2
Cu
3
O
6.50
and YBa
2
Cu
3
O
6.993
, Phys. Rev. B 74, 104508 (2006).
[105] A. Erb, private communication (PUBLISHER, ADDRESS, 2008).
[106] O. P. Vajk, P. K. Mang, M. Greven, P. M. Gehring, and J. W. Lynn, Quantum Impurities in the
Two-Dimensional Spin One-Half Heisenberg Antiferromagnet, Science 295, 1691 (2002).
[107] S. Wakimoto, H. Zhang, K. Yamada, I. Swainson, H. Kim, and R. J. Birgeneau, Direct Relation
between the Low-Energy Spin Excitations and Superconductivity of Overdoped High-T
c
Super-
conductors, Phys. Rev. Lett. 92, 217004 (2004).
[108] T. Timusk and B. Statt, The pseudogap in high-temperature superconductors: an experimental
survey, Rep. Prog. Phys. 62, 61 (1999).
[109] A. G. Loeser, Z.-X. Shen, D. S. Dessau, D. S. Marshall, C. H. Park, P. Fournier, and A. Kapitulnik,
Excitation Gap in the Normal State of Underdoped Bi2Sr2CaCu2O8+, Science 273, 325 (1996).
[110] H. Ding, T. Yokoya, J. C. Campuzano, T. Takahashi, M. Randeria, M. R. Norman, T. Mochiku,
K. Kadowaki, and J. Giapintzakis, Spectroscopic evidence for a pseudogap in the normal state of
underdoped high-Tc superconductors, Nature 382, 51 (1996).
[111] J. Tranquada, B. Sternlieb, J. Axe, Y. Nakamura, and S. Uchida, Evidence for stripe correlations
of spins and holes in copper oxide superconductors, Nature 375, 561 (1995).
[112] H.-H. Klauss, W. Wagener, M. Hillberg, W. Kopmann, H. Walf, F. J. Litterst, M. H ucker, and
B. B uchner, From Antiferromagnetic Order to Static Magnetic Stripes: The Phase Diagram of
(La, Eu)2xSrxCuO4, Phys. Rev. Lett. 85, 4590 (2000).
[113] J. Fink, E. Schierle, E. Weschke, J. Geck, D. Hawthorn, V. Soltwisch, H. Wadati, H.-H. Wu,
H. A. D urr, N. Wizent, B. B uchner, and G. A. Sawatzky, Charge ordering in La
1.8x
Eu
0.2
Sr
x
CuO
4
studied by resonant soft x-ray diffraction, Phys. Rev. B 79, 100502 (2009).
[114] S. A. Kivelson, I. P. Bindloss, E. Fradkin, V. Oganesyan, J. M. Tranquada, A. Kapitulnik, and
C. Howald, How to detect uctuating stripes in the high-temperature superconductors, Rev. Mod.
Phys. 75, 1201 (2003).
[115] M. Vershinin, S. Misra, S. Ono, Y. Abe, Y. Ando, and A. Yazdani, Local Ordering in the Pseudogap
State of the High-Tc Superconductor Bi
2
Sr
2
CaCu
2
O
8+
, Science 303, 1995 (2004).
[116] L. Tassini, W. Prestel, A. Erb, M. Lambacher, and R. Hackl, First-order-type effects in
YBa
2
Cu
3
O
6+x
at the onset of superconductivity, Phys. Rev. B 78, 020511 (2008).
2013
128 R. HACKL AND D. EINZEL BIBLIOGRAPHY
[117] N. Doiron-Leyraud, C. Proust, D. LeBoeuf, J. Levallois, J. Bonnemaison, R. Liang, D. A. Bonn,
W. N. Hardy, and L. Taillefer, Quantum oscillations and the Fermi surface in an underdoped high-
Tc superconductor, Nature 447, 05872 (2007).
[118] V. Hinkov, D. Haug, B. Fauque, P. Bourges, Y. Sidis, A. Ivanov, C. Bernhard, C. Lin, and B.
Keimer, Electronic liquid crystal state in the high-temperature superconductor YBa
2
Cu
3
O
6
.45,
Science 319, 597 (2008).
[119] V. Hinkov, C. Lin, M. Raichle, B. Keimer, Y. Sidis, P. Bourges, S. Pailh` es, and A. Ivanov, Su-
perconductivity and electronic liquid-crystal states in twin-free YBa
2
Cu
3
O
6+x
studied by neutron
scattering, Eur. Phys. J. Special Topics 188, 113 (2010).
[120] R. Daou, J. Chang, D. LeBoeuf, O. Cyr-Choiniere, F. Laliberte, N. Doiron-Leyraud, B. J.
Ramshaw, R. Liang, D. A. Bonn, W. N. Hardy, and L. Taillefer, Broken rotational symmetry
in the pseudogap phase of a high-Tc superconductor, Nature 463, 519 (2010).
[121] L. Alff, Y. Krockenberger, B. Welter, M. Schonecke, R. Gross, D. Manske, and M. Naito, A hidden
pseudogap under the dome of superconductivity in electron-doped high-temperature superconduc-
tors, Nature 422, 698 (2003).
[122] N. P. Armitage, F. Ronning, D. H. Lu, C. Kim, A. Damascelli, K. M. Shen, D. L. Feng, H. Eisaki,
Z.-X. Shen, P. K. Mang, N. Kaneko, M. Greven, Y. Onose, Y. Taguchi, and Y. Tokura, Doping De-
pendence of an n-Type Cuprate Superconductor Investigated by Angle-Resolved Photoemission
Spectroscopy, Phys. Rev. Lett. 88, 257001 (2002).
[123] Y. Dagan, M. M. Qazilbash, C. P. Hill, V. N. Kulkarni, and R. L. Greene, Evidence for a Quantum
Phase Transition in Pr
2x
Ce
x
CuO
4
from Transport Measurements, Phys. Rev. Lett. 92, 167001
(2004).
[124] C. Kusko, R. S. Markiewicz, M. Lindroos, and A. Bansil, Fermi surface evolution and collapse of
the Mott pseudogap in Nd
2x
Ce
x
CuO
4
, Phys. Rev. B 66, 140513 (2002).
[125] H. Matsui, K. Terashima, T. Sato, T. Takahashi, M. Fujita, and K. Yamada, Direct Observation
of a Nonmonotonic d
x
2
y
2
-Wave Superconducting Gap in the Electron-Doped High-T
c
Supercon-
ductor Pr
0.89
LaCe
0.11
CuO
4
, Phys. Rev. Lett. 95, 017003 (2005).
[126] T. Helm, M. V. Kartsovnik, M. Bartkowiak, N. Bittner, M. Lambacher, A. Erb, J. Wosnitza, and
R. Gross, Evolution of the Fermi Surface of the Electron-Doped High-Temperature Superconduc-
tor Nd
2x
Ce
x
CuO
4
Revealed by Shubnikovde Haas Oscillations, Phys. Rev. Lett. 103, 157002
(2009).
[127] F. Gebhard, The Mott Metal-Insulator Transition - Models and Methods (Springer Tracts in Mod-
ern Physics, ADDRESS, 1997), Vol. 137.
[128] N. N ucker, J. Fink, J. C. Fuggle, P. J. Durham, and W. M. Temmerman, Evidence for holes on
oxygen sites in the high-T
c
superconductors La
2x
Sr
x
CuO
4
and YBa
2
Cu
3
O
7y
, Phys. Rev. B 37,
5158 (1988).
[129] F. C. Zhang and T. M. Rice, Effective Hamiltonian for the superconducting Cu oxides, Phys. Rev.
B 37, 3759 (1988).
[130] J. M. Tranquada, S. M. Heald, and A. R. Moodenbaugh, X-ray-absorption near-edge-structure
study of La
2x
(Ba,Sr)
x
CuO
4y
superconductors, Phys. Rev. B 36, 5263 (1987).
[131] V. J. Emery, Theory of high-T
c
superconductivity in oxides, Phys. Rev. Lett. 58, 2794 (1987).
c _ Walther-Meiner-Institut
BIBLIOGRAPHY SUPERCONDUCTIVITY 129
[132] C. M. Varma, S. Schmitt-Rink, and E. Abrahams, Charge transfer excitations and superconductiv-
ity in ionic metalls, Solid State Commun. 62, 681 (1987).
[133] W. Hanke, M. Kiesel, M. Aichhorn, S. Brehm, and E. Arrigoni, The 3-band Hubbard-model versus
the 1-band model for the high-Tc cuprates: Pairing dynamics, superconductivity and the ground-
state phase diagram, Eur. Phys. J. Special Topics 188, 15 (2010).
[134] C. Honerkamp, Iron pncitide superconductors studied by the functional renormalization group,
Eur. Phys. J. Special Topics 188, 33 (2010).
[135] O. K. Andersen, A. I. Liechtenstein, O. Jepsen, and F. Paulsen, LDA energy bands, low-energy
hamiltonians, t
/
, t
//
, t

(k), and J

, J. Phys. and Chem. Solids 56, 1573 (1995).


[136] A. A. Kordyuk, S. V. Borisenko, M. S. Golden, S. Legner, K. A. Nenkov, M. Knupfer,
J. Fink, H. Berger, L. Forr o, and R. Follath, Doping dependence of the Fermi surface in
(Bi, Pb)
2
Sr
2
CaCu
2
O
8+
, Phys. Rev. B 66, 014502 (2002).
[137] A. Damascelli, Z. Hussain, and Z.-X. Shen, Angle-resolved photoemission studies of the cuprate
superconductors, Rev. Mod. Phys. 75, 473 (2003).
[138] B. W. Hoogenboom, C. Berthod, M. Peter, O. Fischer, and A. A. Kordyuk, Modeling scanning
tunneling spectra of Bi
2
Sr
2
CaCu
2
O
8+
, Phys. Rev. B 67, 224502 (2003).
[139] F. Venturini, R. Hackl, and U. Michelucci, Comment on Nonmonotonic d
x
2
y
2 Superconducting
Order Parameter in Nd
2x
Ce
x
CuO
4
, Phys. Rev. Lett. 90, 149701 (2003).
[140] D. S. Inosov, R. Schuster, A. A. Kordyuk, J. Fink, S. V. Borisenko, V. B. Zabolotnyy, D. V.
Evtushinsky, M. Knupfer, B. B uchner, R. Follath, and H. Berger, Excitation energy map of high-
energy dispersion anomalies in cuprates, Phys. Rev. B 77, 212504 (2008).
[141] W. Prestel, F. Venturini, B. Muschler, I. T utt o, R. Hackl, M. Lambacher, A. Erb, S. Komiya, S.
Ono, Y. Ando, D. Inosov, B. Zabolotnyy, V. and V. Borisenko, S. Quantitative comparison of
single- and two-particle properties in the cuprates, Eur. Phys. J. Special Topics 188, 163 (2010).
[142] C. Castellani, C. Di Castro, and M. Grilli, Singular Quasiparticle Scattering in the Proximity of
Charge Instabilities, Phys. Rev. Lett. 75, 4650 (1995).
[143] S. Andergassen, S. Caprara, C. Di Castro, and M. Grilli, Anomalous Isotopic Effect Near the
Charge-Ordering Quantum Criticality, Phys. Rev. Lett. 87, 056401 (2001).
[144] X. J. Zhou, T. Yoshida, A. Lanzara, P. V. Bogdanov, S. A. Kellar, K. M. Shen, W. L. Yang, F.
Ronning, T. Sasagawa, T. Kakeshita, T. Noda, H. Eisaki, S. Uchida, C. T. Lin, F. Zhou, J. W.
Xiong, W. X. Ti, Z. X. Zhao, A. Fujimori, Z. Hussain, and Z.-X. Shen, High-temperature super-
conductors: Universal nodal Fermi velocity, Nature 423, 398 (2003).
[145] T. Valla, A. V. Fedorov, P. D. Johnson, B. O. Wells, S. L. Hulbert, Q. Li, G. D. Gu,
and N. Koshizuka, Evidence for Quantum Critical Behavior in the Optimally Doped Cuprate
Bi2Sr2CaCu2O8+, Science 285, 2110 (1999).
[146] A. Lanzara, P. V. Bogdanov, X. J. Zhou, S. A. Kellar, D. L. Feng, E. D. Lu, T. Yoshida, H.
Eisaki, A. Fujimori, K. Kishio, J.-I. Shimoyama, T. Noda, S. Uchida, Z. Hussain, and Z.-X. Shen,
Evidence for ubiquitous strong electron-phonon coupling in high-temperature superconductors,
Nature 412, 510 (2001).
2013
130 R. HACKL AND D. EINZEL BIBLIOGRAPHY
[147] X. J. Zhou, J. Shi, T. Yoshida, T. Cuk, W. L. Yang, V. Brouet, J. Nakamura, N. Mannella, S.
Komiya, Y. Ando, F. Zhou, W. X. Ti, J. W. Xiong, Z. X. Zhao, T. Sasagawa, T. Kakeshita, H.
Eisaki, S. Uchida, A. Fujimori, Z. Zhang, E. W. Plummer, R. B. Laughlin, Z. Hussain, and Z.-X.
Shen, Multiple Bosonic Mode Coupling in the Electron Self-Energy of (La
2x
Sr
x
)CuO
4
, Phys.
Rev. Lett. 95, 117001 (2005).
[148] T. Cuk, D. Lu, X. Zhou, Z.-X. Shen, T. Devereaux, and N. Nagaosa, A review of electron-phonon
coupling seen in the high-T
c
superconductors by angle-resolved photoemission studies (ARPES),
PSS 242, 11 (2005).
[149] W. Meevasana, N. J. C. Ingle, D. H. Lu, J. R. Shi, F. Baumberger, K. M. Shen, W. S. Lee, T.
Cuk, H. Eisaki, T. P. Devereaux, N. Nagaosa, J. Zaanen, and Z.-X. Shen, Doping Dependence of
the Coupling of Electrons to Bosonic Modes in the Single-Layer High-Temperature Bi
2
Sr
2
CuO
6
Superconductor, Phys. Rev. Lett. 96, 157003 (2006).
[150] W. Meevasana, X. J. Zhou, S. Sahrakorpi, W. S. Lee, W. L. Yang, K. Tanaka, N. Mannella, T.
Yoshida, D. H. Lu, Y. L. Chen, R. H. He, H. Lin, S. Komiya, Y. Ando, F. Zhou, W. X. Ti, J. W.
Xiong, Z. X. Zhao, T. Sasagawa, T. Kakeshita, K. Fujita, S. Uchida, H. Eisaki, A. Fujimori, Z.
Hussain, R. S. Markiewicz, A. Bansil, N. Nagaosa, J. Zaanen, T. P. Devereaux, and Z.-X. Shen,
Hierarchy of multiple many-body interaction scales in high-temperature superconductors, Phys.
Rev. B 75, 174506 (2007).
[151] A. A. Kordyuk, S. V. Borisenko, V. B. Zabolotnyy, J. Geck, M. Knupfer, J. Fink, B. B uchner, C. T.
Lin, B. Keimer, H. Berger, A. V. Pan, S. Komiya, and Y. Ando, Constituents of the Quasiparticle
Spectrum Along the Nodal Direction of High-T
c
Cuprates, Phys. Rev. Lett. 97, 017002 (2006).
[152] T. Dahm, V. Hinkov, S. V. Borisenko, A. A. Kordyuk, V. B. Zabolotnyy, J. Fink, B. Buchner, D. J.
Scalapino, W. Hanke, and B. Keimer, Strength of the spin-uctuation-mediated pairing interaction
in a high-temperature superconductor, Nat. Phys. 5, 217 (2009).
[153] A. Perali, C. Castellani, C. Di Castro, and M. Grilli, d-wave superconductivity near charge insta-
bilities, Phys. Rev. B 54, 16216 (1996).
[154] S. A. Kivelson, E. Fradkin, and V. J. Emery, Electronic liquid-crystal phases of a doped Mott
insulator, Nature 393, 550 (1998).
[155] C. M. Varma, Non-Fermi-liquid states and pairing instability of a general model of copper oxide
metals , Phys. Rev. B 55, 14554 (1997).
[156] V. Aji and C. M. Varma, Theory of the Quantum Critical Fluctuations in Cuprate Superconductors,
Phys. Rev. Lett. 99, 067003 (2007).
[157] A. A. Kordyuk, S. V. Borisenko, A. Koitzsch, J. Fink, M. Knupfer, B. B uchner, H. Berger, G. Mar-
garitondo, C. T. Lin, B. Keimer, S. Ono, and Y. Ando, Manifestation of the Magnetic Resonance
Mode in the Nodal Quasiparticle Lifetime of the Superconducting Cuprates, Phys. Rev. Lett. 92,
257006 (2004).
[158] C. M. Varma, P. B. Littlewood, S. Schmitt-Rink, E. Abrahams, and A. E. Ruckenstein, Phe-
nomenology of the normal state of Cu-O high-temperature superconductors, Phys. Rev. Lett. 63,
1996 (1989).
[159] M. Plat e, J. D. F. Mottershead, I. S. Elmov, D. C. Peets, R. Liang, D. A. Bonn, W. N. Hardy,
S. Chiuzbaian, M. Falub, M. Shi, L. Patthey, and A. Damascelli, Fermi Surface and Quasiparticle
Excitations of Overdoped Tl
2
Ba
2
CuO
6+
, Phys. Rev. Lett. 95, 077001 (2005).
c _ Walther-Meiner-Institut
BIBLIOGRAPHY SUPERCONDUCTIVITY 131
[160] X. J. Zhou, T. Yoshida, D.-H. Lee, W. L. Yang, V. Brouet, F. Zhou, W. X. Ti, J. W. Xiong,
Z. X. Zhao, T. Sasagawa, T. Kakeshita, H. Eisaki, S. Uchida, A. Fujimori, Z. Hussain, and Z.-X.
Shen, Dichotomy between Nodal and Antinodal Quasiparticles in Underdoped (La
2x
Sr
x
)CuO
4
Superconductors, Phys. Rev. Lett. 92, 187001 (2004).
[161] J. Chang, M. Shi, S. Pailh es, M. M ansson, T. Claesson, O. Tjernberg, A. Bendounan, Y. Sassa, L.
Patthey, N. Momono, M. Oda, M. Ido, S. Guerrero, C. Mudry, and J. Mesot, Anisotropic quasipar-
ticle scattering rates in slightly underdoped to optimally doped high-temperature La
2x
Sr
x
CuO
4
superconductors, Phys. Rev. B 78, 205103 (2008).
[162] J. Graf, G.-H. Gweon, K. McElroy, S. Y. Zhou, C. Jozwiak, E. Rotenberg, A. Bill, T. Sasagawa,
H. Eisaki, S. Uchida, H. Takagi, D.-H. Lee, and A. Lanzara, Universal High Energy Anomaly
in the Angle-Resolved Photoemission Spectra of High Temperature Superconductors: Possible
Evidence of Spinon and Holon Branches, Phys. Rev. Lett. 98, 067004 (2007).
[163] J. Chang, S. Pailh es, M. Shi, M. M ansson, T. Claesson, O. Tjernberg, J. Voigt, V. Perez, L. Patthey,
N. Momono, M. Oda, M. Ido, A. Schnyder, C. Mudry, and J. Mesot, When low- and high-energy
electronic responses meet in cuprate superconductors, Phys. Rev. B 75, 224508 (2007).
[164] D. S. Inosov, S. V. Borisenko, I. Eremin, A. A. Kordyuk, V. B. Zabolotnyy, J. Geck, A. Koitzsch,
J. Fink, M. Knupfer, B. Buchner, H. Berger, and R. Follath, Relation between the one-particle
spectral function and dynamic spin susceptibility of superconducting Bi
2
Sr
2
CaCu
2
O
8
, Phys.
Rev. B 75, 172505 (2007).
[165] R. Preuss, W. Hanke, and W. von der Linden, Quasiparticle Dispersion of the 2D Hubbard Model:
From an Insulator to a Metal, Phys. Rev. Lett. 75, 1344 (1995).
[166] B. Moritz, F. Schmitt, W. Meevasana, S. Johnston, E. Motoyama, M.Greven, D. Lu, C. Kim, R.
Scalettar, Z.-X. Shen, and T. Devereaux, Effect of strong correlations on the high energy anomaly
in hole- and electron-doped high-Tc superconductors, New J. Phys. 11, 093020 (2009).
[167] T. Ito, H. Takagi, S. Ishibashi, T. Ido, and S. Uchida, Normal-state conductivity between CuO2
planes in copper oxide superconductors, Nature 350, 596 (1991).
[168] K. Takenaka, K. Mizuhashi, H. Takagi, and S. Uchida, Interplane charge transport in
YBa
2
Cu
3
O
7y
: Spin-gap effect on in-plane and out-of-plane resistivity, Phys. Rev. B 50, 6534
(1994).
[169] L. Forro, Out-of-plane resistivity of Bi2Sr2CaCu2O8+x high temperature superconductor, Phys.
Lett. A 179, 140 (1993).
[170] G. S. Boebinger, Y. Ando, A. Passner, T. Kimura, M. Okuya, J. Shimoyama, K. Kishio, K.
Tamasaku, N. Ichikawa, and S. Uchida, Insulator-to-Metal Crossover in the Normal State of
La
2x
Sr
x
CuO
4
Near Optimum Doping, Phys. Rev. Lett. 77, 5417 (1996).
[171] T. Ito, K. Takenaka, and S. Uchida, Systematic deviation from T-linear behavior in the in-plane
resistivity of YBa
2
Cu
3
O
7y
: Evidence for dominant spin scattering, Phys. Rev. Lett. 70, 3995
(1993).
[172] J. L. Tallon and J. W. Loram, The doping dependence of T* - what is the real high-Tc phase
diagram?, Physica C 349, 53 (2001).
[173] M. Abdel-Jawad, J. G. Analytis, L. Balicas, A. Carrington, J. P. H. Charmant, M. M. J. French, and
N. E. Hussey, Correlation between the Superconducting Transition Temperature and Anisotropic
Quasiparticle Scattering in Tl
2
Ba
2
CuO
6+
, Phys. Rev. Lett. 99, 107002 (2007).
2013
132 R. HACKL AND D. EINZEL BIBLIOGRAPHY
[174] R. A. Cooper, Y. Wang, B. Vignolle, O. J. Lipscombe, S. M. Hayden, Y. Tanabe, T. Adachi, Y.
Koike, M. Nohara, H. Takagi, C. Proust, and N. E. Hussey, Anomalous Criticality in the Electrical
Resistivity of La
2x
Sr
x
CuO
4
, Science 323, 603 (2009).
[175] L. Taillefer, Scattering and Pairing in Cuprate Superconductors, Annu. Rev. Cond. Mat. Phys. 1,
51 (2010).
[176] H. Alloul, T. Ohno, and P. Mendels,
89
Y NMR evidence for a fermi-liquid behavior in
YBa
2
Cu
3
O
6+x
, Phys. Rev. Lett. 63, 1700 (1989).
[177] C. C. Homes, T. Timusk, R. Liang, D. A. Bonn, and W. N. Hardy, Optical conductivity of c axis
oriented YBa
2
Cu
3
O
6.70
: Evidence for a pseudogap, Phys. Rev. Lett. 71, 1645 (1993).
[178] A. V. Puchkov, D. N. Basov, and T. Timusk, The pseudogap state in high- superconductors: an
infrared study, J. Phys. Condens. Matter 8, 10049 (1996).
[179] D. N. Basov and T. Timusk, Electrodynamics of high- T
c
superconductors, Rev. Mod. Phys. 77,
721 (2005).
[180] B. S. Shastry and B. I. Shraiman, Theory of Raman scattering in Mott-Hubbard systems, Phys.
Rev. Lett. 65, 1068 (1990).
[181] T. P. Devereaux and R. Hackl, Inelastic light scattering from correlated electrons, Rev. Mod. Phys.
79, 175 (2007).
[182] T. P. Devereaux, D. Einzel, B. Stadlober, R. Hackl, D. H. Leach, and J. J. Neumeier, Electronic
Raman scattering in high-Tc superconductors: A probe of dx2-y2 pairing, Phys. Rev. Lett. 72,
396 (1994).
[183] F. Venturini, M. Opel, T. P. Devereaux, J. K. Freericks, I. T utt o, B. Revaz, E. Walker, H. Berger, L.
Forr o, and R. Hackl, Observation of an Unconventional Metal-Insulator Transition in Overdoped
CuO
2
Compounds, Phys. Rev. Lett. 89, 107003 (2002).
[184] B. Muschler, W. Prestel, L. Tassini, R. Hackl, M. Lambacher, A. Erb, S. Komiya, Y. Ando, D.
Peets, W. Hardy, R. Liang, and D. Bonn, Electron interactions and charge ordering in CuO
2
com-
pounds, Eur. Phys. J. Special Topics 188, 131 (2010).
[185] S. Billinge, M. Gutmann, and E. Bo zin, Structural Response to Local Charge Order in Underdoped
but Superconducting La
2x
(Sr,Ba)
x
CuO
4
, Int. J. Mod. Phys. B 17, 3640 (2003).
[186] S. Blanc, Y. Gallais, A. Sacuto, M. Cazayous, M. A. Measson, G. D. Gu, J. S. Wen, and Z. J.
Xu, Quantitative Raman measurement of the evolution of the Cooper-pair density with doping in
Bi
2
Sr
2
CaCu
2
O
8+
superconductors, Phys. Rev. B 80, 140502 (2009).
[187] B. Muschler, F. Kretzschmar, R. Hackl, J.-H. Chu, J. G. Analytis, and I. R. Fisher, Doping depen-
dence of the electronic properties of Ba(Fe
1x
Co
x
)
2
As
2
, 49 (2010).
[188] T. Yoshida, X. J. Zhou, K. Tanaka, W. L. Yang, Z. Hussain, Z.-X. Shen, A. Fujimori, S. Sahrakorpi,
M. Lindroos, R. S. Markiewicz, A. Bansil, S. Komiya, Y. Ando, H. Eisaki, T. Kakeshita, and S.
Uchida, Systematic doping evolution of the underlying Fermi surface of La
2x
Sr
x
CuO
4
, Phys.
Rev. B 74, 224510 (2006).
[189] X. K. Chen, J. G. Naeini, K. C. Hewitt, J. C. Irwin, R. Liang, and W. N. Hardy, Electronic Raman
scattering in underdoped YBa
2
Cu
3
O
6.5
, Phys. Rev. B 56, R513 (1997).
c _ Walther-Meiner-Institut
BIBLIOGRAPHY SUPERCONDUCTIVITY 133
[190] M. Le Tacon, A. Sacuto, A. Georges, G. Kotliar, Y. Gallais, D. Colson, and A. Forget, Two
energy scales and two distinct quasiparticle dynamics in the superconducting state of underdoped
cuprates, Nat. Phys. 2, 537 (2006).
[191] D. LeBoeuf, N. Doiron-Leyraud, J. Levallois, R. Daou, J.-B. Bonnemaison, N. E. Hussey, L.
Balicas, B. J. Ramshaw, R. Liang, D. A. Bonn, W. N. Hardy, S. Adachi, C. Proust, and L. Taillefer,
Electron pockets in the Fermi surface of hole-doped high-Tc superconductors, Nature 450, 533
(2007).
[192] Y. Wang, L. Li, and N. P. Ong, Nernst effect in high- T
c
superconductors, Phys. Rev. B 73, 024510
(2006).
[193] F. Rullier-Albenque, R. Tourbot, H. Alloul, P. Lejay, D. Colson, and A. Forget, Nernst Effect and
Disorder in the Normal State of High-T
c
Cuprates, Phys. Rev. Lett. 96, 067002 (2006).
[194] O. Cyr-Choiniere, R. Daou, F. Laliberte, D. LeBoeuf, N. Doiron-Leyraud, J. Chang, J.-Q. Yan,
J.-G. Cheng, J.-S. Zhou, J. B. Goodenough, S. Pyon, T. Takayama, H. Takagi, Y. Tanaka, and L.
Taillefer, Enhancement of the Nernst effect by stripe order in a high-Tc superconductor, Nature
458, 743 (2009).
[195] H.-C. Ri, R. Gross, F. Gollnik, A. Beck, R. P. Huebener, P. Wagner, and H. Adrian, Nernst,
Seebeck, and Hall effects in the mixed state of YBa
2
Cu
3
O
7
and Bi
2
Sr
2
CaCu
2
O
8+x
thin lms:
A comparative study, Phys. Rev. B 50, 3312 (1994).
[196] K. Behnia, The Nernst effect and the boundaries of the Fermi liquid picture, J. Phys. Condens.
Matter 21, 113101 (2009).
[197] C. Hess, E. Ahmed, U. Ammerahl, A. Revcolevschi, and B. B uchner, Nernst effect of stripe
ordering La
1.8x
Eu
0.2
Sr
x
CuO
4
, Eur. Phys. J. Special Topics 188, 103 (2010).
[198] J. Xia, E. Schemm, G. Deutscher, S. A. Kivelson, D. A. Bonn, W. N. Hardy, R. Liang, W.
Siemons, G. Koster, M. M. Fejer, and A. Kapitulnik, Polar Kerr-Effect Measurements of the High-
Temperature YBa
2
Cu
3
O
6+x
Superconductor: Evidence for Broken Symmetry near the Pseudogap
Temperature, Phys. Rev. Lett. 100, 127002 (2008).
[199] S.-W. Cheong, G. Aeppli, T. E. Mason, H. Mook, S. M. Hayden, P. C. Caneld, Z. Fisk, K. N.
Clausen, and J. L. Martinez, Incommensurate magnetic uctuations in La
2x
Sr
x
CuO
4
, Phys. Rev.
Lett. 67, 1791 (1991).
[200] A. Bianconi, N. L. Saini, A. Lanzara, M. Missori, T. Rossetti, H. Oyanagi, H. Yamaguchi, K. Oka,
and T. Ito, Determination of the Local Lattice Distortions in the CuO
2
Plane of La
1.85
Sr
0.15
CuO
4
,
Phys. Rev. Lett. 76, 3412 (1996).
[201] K. B. Lyons, P. A. Fleury, J. P. Remeika, A. S. Cooper, and T. J. Negran, Dynamics of spin
uctuations in lanthanum cuprate, Phys. Rev. B 37, 2353 (1988).
[202] P. E. Sulewski, P. A. Fleury, K. B. Lyons, S.-W. Cheong, and Z. Fisk, Light scattering from
quantum spin uctuations in R
2
CuO
4
(R=La, Nd, Sm), Phys. Rev. B 41, 225 (1990).
[203] S. Sugai, H. Suzuki, Y. Takayanagi, T. Hosokawa, and N. Hayamizu, Carrier-density-dependent
momentum shift of the coherent peak and the LO phonon mode in p-type high-T
c
superconductors,
Phys. Rev. B 68, 184504 (2003).
[204] E. M. Motoyama, G. Yu, I. M. Vishik, O. P. Vajk, P. K. Mang, and M. Greven, Spin correlations
in the electron-doped high-transition-temperature superconductor Nd
2x
Ce
x
CuO
4
, Nature 445,
05437 (2007).
2013
134 R. HACKL AND D. EINZEL BIBLIOGRAPHY
[205] M. Vojta and T. Ulbricht, Magnetic Excitations in a Bond-Centered Stripe Phase: Spin Waves Far
from the Semiclassical Limit, Phys. Rev. Lett. 93, 127002 (2004).
[206] G. S. Uhrig, K. P. Schmidt, and M. Gr uninger, Unifying Magnons and Triplons in Stripe-Ordered
Cuprate Superconductors, Phys. Rev. Lett. 93, 267003 (2004).
[207] S. E. Barrett, D. J. Durand, C. H. Pennington, C. P. Slichter, T. A. Friedmann, J. P. Rice, and D. M.
Ginsberg,
63
Cu Knight shifts in the superconducting state of YBa
2
Cu
3
O
7
(T
c
=90 K), Phys. Rev.
B 41, 6283 (1990).
[208] J. Rossat-Mignod, L. Regnault, C. Vettier, P. Bourges, P. Burlet, J. Bossy, J. Henry, and G. Laper-
tot, Neutron scattering study of the YBa
2
Cu
3
O
6+x
system, Physica C 185, 86 (1991).
[209] P. M. Singer and T. Imai, Systematic
63
Cu NQR and
89
Y NMR Study of Spin Dynamics in
Y
1z
Ca
z
Ba
2
Cu
3
O
y
across the Superconductor-Insulator Boundary, Phys. Rev. Lett. 88, 187601
(2002).
[210] A. J. Millis, H. Monien, and D. Pines, Phenomenological model of nuclear relaxation in the normal
state of YBa
2
Cu
3
O
7
, Phys. Rev. B 42, 167 (1990).
[211] H. A. Mook, M. Yethiraj, G. Aeppli, T. E. Mason, and T. Armstrong, Polarized neutron determi-
nation of the magnetic excitations in YBa
2
Cu
3
O
7
, Phys. Rev. Lett. 70, 3490 (1993).
[212] H. F. Fong, P. Bourges, Y. Sidis, L. P. Regnault, A. Ivanov, G. D. Gu, N. Koshizuka, and B. Keimer,
Neutron scattering from magnetic excitations in Bi
2
Sr
2
CaCu
2
O
8+
, Nature 398, 588 (1999).
[213] H. F. Fong, P. Bourges, Y. Sidis, L. P. Regnault, J. Bossy, A. Ivanov, D. L. Milius, I. A. Aksay, and
B. Keimer, Spin susceptibility in underdoped YBa
2
Cu
3
O
6+x
, Phys. Rev. B 61, 14773 (2000).
[214] H. He, P. Bourges, Y. Sidis, C. Ulrich, L. P. Regnault, S. Pailhs, N. S. Berzigiarova, N. N.
Kolesnikov, and B. Keimer, Magnetic Resonant Mode in the Single-Layer High-Temperature Su-
perconductor Tl2Ba2CuO6+, Science 295, 1045 (2002).
[215] G. Yu, Y. Li, E. M. Motoyama, X. Zhao, N. Bari si c, Y. Cho, P. Bourges, K. Hradil, R. A. Mole, and
M. Greven, Magnetic resonance in the model high-temperature superconductor HgBa
2
CuO
4+
,
Phys. Rev. B 81, 064518 (2010).
[216] H.-Y. Kee, S. A. Kivelson, and G. Aeppli, Spin1 Neutron Resonance Peak Cannot Account for
Electronic Anomalies in the Cuprate Superconductors, Phys. Rev. Lett. 88, 257002 (2002).
[217] D. Munzar, C. Bernhard, and M. Cardona, Does the peak in the magnetic susceptibility determine
the in-plane infrared conductivity of YBCO? A theoretical study, Physica C 312, 121 (1999).
[218] J. P. Carbotte, SchachingerE., and D. N. Basov, Coupling strength of charge carriers to spin uc-
tuations in high-temperature superconductors, Nature 401, 354 (1999).
[219] A. Abanov, A. V. Chubukov, M. Eschrig, M. R. Norman, and J. Schmalian, Neutron Resonance in
the Cuprates and its Effect on Fermionic Excitations, Phys. Rev. Lett. 89, 177002 (2002).
[220] M. Fujita, K. Yamada, H. Hiraka, P. M. Gehring, S. H. Lee, S. Wakimoto, and G. Shirane, Static
magnetic correlations near the insulating-superconducting phase boundary in La
2x
Sr
x
CuO
4
,
Phys. Rev. B 65, 064505 (2002).
[221] J. M. Tranquada, H. Woo, T. G. Perring, H. Goka, G. D. Gu, G. Xu, M. Fujita, and K. Yamada,
Quantum magnetic excitations from stripes in copper oxide superconductors, Nature 429, 534
(2004).
c _ Walther-Meiner-Institut
BIBLIOGRAPHY SUPERCONDUCTIVITY 135
[222] M. Vojta, Tendencies toward nematic order in YBa
2
Cu
3
O
6+
distortion vs. incipient charge
stripes, Eur. Phys. J. Special Topics 188, 49 (2010).
[223] L. Benfatto, S. Caprara, and C. D. Castro, Gap and pseudogap evolution within the charge-
ordering scenario for superconducting cuprates, Eur. Phys. J. B 17, 95 (2000).
[224] J. Zaanen and O. Gunnarsson, Charged magnetic domain lines and the magnetism of high-T
c
oxides, Phys. Rev. B 40, 7391 (1989).
[225] K. Machida, Magnetism in La
2
CuO
4
based compounds, Physica C 158, 192 (1989).
[226] V. J. Emery, S. A. Kivelson, and H. Q. Lin, Phase separation in the t-J model, Phys. Rev. Lett. 64,
475 (1990).
[227] E. Demler, W. Hanke, and S.-C. Zhang, SO(5) theory of antiferromagnetism and superconductiv-
ity, Rev. Mod. Phys. 76, 909 (2004).
[228] E. Fradkin, S. A. Kivelson, M. J. Lawler, J. P. Eisenstein, and A. P. Mackenzie, Nematic Fermi
Fluids in Condensed Matter Physics, Annu. Rev. Cond. Mat. Phys. 1, 153 (2010).
[229] S. Sachdev, Quantum Phase Transitions (Cambridge University Press, ADDRESS, 1999).
[230] P. W. Anderson, The Resonating Valence Bond State in La
2
CuO
4
and Superconductivity, Science
235, 1196 (1987).
[231] M. Fujita, H. Goka, K. Yamada, J. M. Tranquada, and L. P. Regnault, Stripe order, depinning,
and uctuations in La
1.875
Ba
0.125
CuO
4
and La
1.875
Ba
0.075
Sr
0.050
CuO
4
, Phys. Rev. B 70, 104517
(2004).
[232] S. V. Borisenko, A. A. Kordyuk, A. N. Yaresko, V. B. Zabolotnyy, D. S. Inosov, R. Schuster, B.
B uchner, R. Weber, R. Follath, L. Patthey, and H. Berger, Pseudogap and Charge Density Waves
in Two Dimensions, Phys. Rev. Lett. 100, 196402 (2008).
[233] V. Brouet, W. L. Yang, X. J. Zhou, Z. Hussain, R. G. Moore, R. He, D. H. Lu, Z. X. Shen, J.
Laverock, S. B. Dugdale, N. Ru, and I. R. Fisher, Angle-resolved photoemission study of the
evolution of band structure and charge density wave properties in RTe
3
( R =Y , La, Ce, Sm, Gd,
Tb, and Dy), Phys. Rev. B 77, 235104 (2008).
[234] P. Trey, S. Gygax, and J. P. Jan, Anisotropy of the Ginzburg-Landau parameter in NbSe
2
, J. Low
Temp. Phys. 11, 421 (1973).
[235] W. Metzner, D. Rohe, and S. Andergassen, Soft Fermi Surfaces and Breakdown of Fermi-Liquid
Behavior, Phys. Rev. Lett. 91, 066402 (2003).
[236] Q. Li, M. H ucker, G. D. Gu, A. M. Tsvelik, and J. M. Tranquada, Two-Dimensional Supercon-
ducting Fluctuations in Stripe-Ordered La
1.875
Ba
0.125
CuO
4
, Phys. Rev. Lett. 99, 067001 (2007).
[237] M. H ucker, M. v. Zimmermann, M. Debessai, J. S. Schilling, J. M. Tranquada, and G. D. Gu,
Spontaneous Symmetry Breaking by Charge Stripes in the High Pressure Phase of Superconduct-
ing La
1.875
Ba
0.125
CuO
4
, Phys. Rev. Lett. 104, 057004 (2010).
[238] A. Dubroka, L. Yu, D. Munzar, K. Kim, M. R ossle, V. Malik, C. Lin, B. Keimer, T. Wolf, and
C. Bernhard, Pseudogap and precursor superconductivity in underdoped cuprate high temperature
superconductors: A far-infrared ellipsometry study, Eur. Phys. J. Special Topics 188, 73 (2010).
[239] N. Toyota, M. Lang, and J. M uller, Low-Dimensional Molecular Metals (Springer, Berlin-
Heidelberg, 2007).
2013
136 R. HACKL AND D. EINZEL BIBLIOGRAPHY
[240] R. Nemetschek, M. Opel, C. Hoffmann, P. F. M uller, R. Hackl, H. Berger, L. Forr o, A. Erb, and
E. Walker, Pseudogap and Superconducting Gap in the Electronic Raman Spectra of Underdoped
Cuprates, Phys. Rev. Lett. 78, 4837 (1997).
[241] J. L. Smith, J. S. Brooks, C. M. Fowler, B. L. Freeman, J. D. Goettee, W. L. Hults, J. C. King, P. M.
Mankiewich, E. I. Obaldia, M. L. OMalley, D. G. Rickel, and W. J. Skocpol, Low-temperature
critical eld of YBCO, J. Supercond. 7, 269 (1994).
[242] B. Roas, L. Schultz, and G. Saemann-Ischenko, Anisotropy of the critical current density in epi-
taxial YBa
2
Cu
3
O
x
lms, Phys. Rev. Lett. 64, 479 (1990).
[243] P. Anderson, Theory of dirty superconductors, J. Phys. and Chem. Solids 11, 26 (1959).
[244] R. Hackl, W. Gl aser, P. M uller, D. Einzel, and K. Andres, Light-scattering study of the supercon-
ducting energy gap in YBa
2
Cu
3
O
7
single crystals, Phys. Rev. B 38, 7133 (1988).
[245] S. L. Cooper, F. Slakey, M. V. Klein, J. P. Rice, E. D. Bukowski, and D. M. Ginsberg, Gap
anisotropy and phonon self-energy effects in single-crystal YBa
2
Cu
3
O7, Phys. Rev. B 38,
11934 (1988).
[246] A. Yamanaka, T. Kimura, F. Minami, K. Inoue, and S. Takekawa, Superconducting Gap Excita-
tions in Bi-Sr-Ca-Cu-O Superconductor Observed by Raman Scattering, Jpn. J. Appl. Phys. 27,
L1902 (1988).
[247] H. Monien and A. Zawadowski, Theory of interband electron Raman scattering in YBa
2
Cu
3
O
7
:
A probe of unconventional superconductivity, Phys. Rev. Lett. 63, 911 (1989).
[248] Z.-X. Shen, D. S. Dessau, B. O. Wells, D. M. King, W. E. Spicer, A. J. Arko, D. Marshall,
L. W. Lombardo, A. Kapitulnik, P. Dickinson, S. Doniach, J. DiCarlo, T. Loeser, and C. H. Park,
Anomalously large gap anisotropy in the a-b plane of B
2
Sr
2
CaCu
2
O
8+
, Phys. Rev. Lett. 70, 1553
(1993).
[249] H. Ding, M. R. Norman, J. C. Campuzano, M. Randeria, A. F. Bellman, T. Yokoya, T. Taka-
hashi, T. Mochiku, and K. Kadowaki, Angle-resolved photoemission spectroscopy study of the
superconducting gap anisotropy in Bi
2
Sr
2
CaCu
2
O
8+x
, Phys. Rev. B 54, R9678 (1996).
[250] T. Cuk, F. Baumberger, D. H. Lu, N. Ingle, X. J. Zhou, H. Eisaki, N. Kaneko, Z. Hussain, T. P.
Devereaux, N. Nagaosa, and Z.-X. Shen, Coupling of the B
1g
Phonon to the Antinodal Electronic
States of Bi
2
Sr
2
Ca
0.92
Y
0.08
Cu
2
O
8+
, Phys. Rev. Lett. 93, 117003 (2004).
[251] M. V. Klein and S. B. Dierker, Theory of Raman scattering in superconductors, Phys. Rev. B 29,
4976 (1984).
[252] C. Kendziora and A. Rosenberg, a-b plane anisotropy of the superconducting gap in
Bi
2
Sr
2
CaCu
2
O
8+
, Phys. Rev. B 52, 9867 (1995).
[253] L. V. Gasparov, P. Lemmens, N. N. Kolesnikov, and G. G untherodt, Electronic Raman scattering
in Tl
2
Ba
2
CuO
6+
Symmetry of the order parameter, oxygen doping effects, and normal-state
scattering, Phys. Rev. B 58, 11753 (1998).
[254] M. Opel, R. Nemetschek, C. Hoffmann, R. Philipp, P. F. M uller, R. Hackl, I. T utt o, A. Erb, B.
Revaz, E. Walker, H. Berger, and L. Forr o, Carrier relaxation, pseudogap, and superconducting
gap in high-T
c
cuprates: A Raman scattering study, Phys. Rev. B 61, 9752 (2000).
[255] S. Sugai and T. Hosokawa, Relation between the Superconducting Gap Energy and the Two-
Magnon Raman Peak Energy in Bi
2
Sr
2
Ca
1x
Y
x
Cu
2
O
8+
, Phys. Rev. Lett. 85, 1112 (2000).
c _ Walther-Meiner-Institut
BIBLIOGRAPHY SUPERCONDUCTIVITY 137
[256] F. Venturini, M. Opel, R. Hackl, H. Berger, L. Forr, and B. Revaz, Doping dependence of the
electronic Raman spectra in cuprates, J. Phys. and Chem. Solids 63, 2345 (2002).
[257] C. Panagopoulos and T. Xiang, Relationship between the Superconducting Energy Gap and the
Critical Temperature in High- T
c
Superconductors, Phys. Rev. Lett. 81, 2336 (1998).
[258] W. S. Lee, I. M. Vishik, K. Tanaka, D. H. Lu, T. Sasagawa, N. Nagaosa, T. P. Devereaux, Z.
Hussain, and Z.-X. Shen, Abrupt onset of a second energy gap at the superconducting transition
of underdoped Bi2212, Nature 450, 81 (2007).
[259] B. Stadlober, G. Krug, R. Nemetschek, R. Hackl, J. L. Cobb, and J. T. Markert, Is Nd
2x
Ce
x
CuO
4
a High-Temperature Superconductor?, Phys. Rev. Lett. 74, 4911 (1995).
[260] G. Blumberg, A. Koitzsch, A. Gozar, B. S. Dennis, C. A. Kendziora, P. Fournier, and R. L. Greene,
Nonmonotonic d
x
2
y
2 Superconducting Order Parameter in Nd
2x
Ce
x
CuO
4
, Phys. Rev. Lett. 88,
107002 (2002).
[261] B. Chesca, K. Ehrhardt, M. M ole, R. Straub, D. Koelle, R. Kleiner, and A. Tsukada, Magnetic-
Field Dependence of the Maximum Supercurrent of La
2x
Ce
x
CuO
4y
Interferometers: Evidence
for a Predominant d
x
2
y
2 Superconducting Order Parameter, Phys. Rev. Lett. 90, 057004 (2003).
[262] M. M. Qazilbash, A. Koitzsch, B. S. Dennis, A. Gozar, H. Balci, C. A. Kendziora, R. L. Greene,
and G. Blumberg, Evolution of superconductivity in electron-doped cuprates: Magneto-Raman
spectroscopy, Phys. Rev. B 72, 214510 (2005).
[263] J. Orenstein and A. J. Millis, Advances in the Physics of High-Temperature Superconductivity,
Science 88, 468 (2000).
[264] A. Kaminski, H. M. Fretwell, M. R. Norman, M. Randeria, S. Rosenkranz, U. Chatterjee, J. C.
Campuzano, J. Mesot, T. Sato, T. Takahashi, T. Terashima, M. Takano, K. Kadowaki, Z. Z. Li,
and H. Raffy, Momentum anisotropy of the scattering rate in cuprate superconductors, Phys. Rev.
B 71, 014517 (2005).
[265] M. R. Norman, A. Kanigel, M. Randeria, U. Chatterjee, and J. C. Campuzano, Modeling the Fermi
arc in underdoped cuprates, Phys. Rev. B 76, 174501 (2007).
[266] Y. J. Uemura, G. M. Luke, B. J. Sternlieb, J. H. Brewer, J. F. Carolan, W. N. Hardy, R. Kadono,
J. R. Kempton, R. F. Kie, S. R. Kreitzman, P. Mulhern, T. M. Riseman, D. L. Williams, B. X.
Yang, S. Uchida, H. Takagi, J. Gopalakrishnan, A. W. Sleight, M. A. Subramanian, C. L. Chien,
M. Z. Cieplak, G. Xiao, V. Y. Lee, B. W. Statt, C. E. Stronach, W. J. Kossler, and X. H. Yu, Uni-
versal Correlations between T
c
and
n
s
m

(Carrier Density over Effective Mass) in High-T


c
Cuprate
Superconductors, Phys. Rev. Lett. 62, 2317 (1989).
[267] D. L. Feng, D. H. Lu, K. M. Shen, C. Kim, H. Eisaki, A. Damascelli, R. Yoshizaki, J.-i. Shi-
moyama, K. Kishio, G. D. Gu, S. Oh, A. Andrus, J. ODonnell, J. N. Eckstein, and Z.-X. Shen,
Signature of Superuid Density in the Single-Particle Excitation Spectrum of Bi2Sr2CaCu2O8+,
Science 289, 277 (2000).
[268] K. McElroy, D.-H. Lee, J. E. Hoffman, K. M. Lang, J. Lee, E. W. Hudson, H. Eisaki, S. Uchida,
and J. C. Davis, Coincidence of Checkerboard Charge Order and Antinodal State Decoherence in
Strongly Underdoped Superconducting Bi
2
Sr
2
CaCu
2
O
8+
, Phys. Rev. Lett. 94, 197005 (2005).
[269] R. Zeyher and A. Greco, Inuence of Collective Effects and the d Charge-Density Wave on Elec-
tronic Raman Scattering in High-T
c
Superconductors, Phys. Rev. Lett. 89, 177004 (2002).
2013
138 R. HACKL AND D. EINZEL BIBLIOGRAPHY
[270] A. N. Pasupathy, A. Pushp, K. K. Gomes, C. V. Parker, J. Wen, Z. Xu, G. Gu, S. Ono, Y. Ando,
and A. Yazdani, Electronic Origin of the Inhomogeneous Pairing Interaction in the High-Tc Su-
perconductor Bi2Sr2CaCu2O8+, Science 320, 196 (2008).
[271] A. Goncharov and V. Struzhkin, Raman spectroscopy of metals, high-temperature superconduc-
tors and related materials under high pressure, J. Raman Spectrosc. 34, 532 (2003).
[272] S. Pailhes, C. Ulrich, B. Fauqu e, V. Hinkov, Y. Sidis, A. Ivanov, C. T. Lin, B. Keimer, and P.
Bourges, Doping Dependence of Bilayer Resonant Spin Excitations in (Y,Ca)Ba
2
Cu
3
O
6+x
, Phys.
Rev. Lett. 96, 257001 (2006).
[273] L. N. Cooper, Bound Electron Pairs in a Degenerate Fermi Gas, Phys. Rev. 104, 1189 (1956).
[274] D. J. Scalapino, in Superconductivity - An Introduction To Fluid, Heat And Mass Transport
Processes, edited by R. Parks (Marcel Dekker, ADDRESS, 1969), p. 1456, iSBN-13: 978-
0824705961.
[275] W. L. McMillan and J. M. Rowell, Lead Phonon Spectrum Calculated from Superconducting
Density of States, Phys. Rev. Lett. 14, 108 (1965).
[276] W. Weber, The phonons in high T
c
A15 compounds, Physica B 126, 217 (1984).
[277] J. P. Carbotte, Properties of boson-exchange superconductors, Rev. Mod. Phys. 62, 1027 (1990).
[278] J. Carbotte and F. Marsiglio, in Handbook of High-Temperature Superconductivity Theory and
Experiment, edited by J. Schrieffer and J. Brooks (Springer, ADDRESS, 2007).
[279] A. Migdal, , Sov. Phys. JETP 7, 996 (1958).
[280] S. Lupi, P. Maselli, M. Capizzi, P. Calvani, P. Giura, and P. Roy, Evolution of a Polaron Band
through the Phase Diagram of Nd
2x
Ce
x
CuO
4y
, Phys. Rev. Lett. 83, 4852 (1999).
[281] P. Calvani, Optical properties of polarons, La Rivista del Nuovo Cimento 24, 1 (2001).
[282] B. Renker, F. Gompf, E. Gering, N. Ncker, D. Ewert, W. Reichardt, and H. Rietschel, Phonon
density-of-states for the high-Tc superconductor La
1.85
Sr
0.15
CuO
4
and its non-superconducting
reference La
2
CuO
4
, Z. Phys. B: Condens. Matter 67, 15 (1987).
[283] B. Renker, F. Gompf, E. Gering, G. Roth, W. Reichardt, D. Ewert, H. Rietschel, and H. Mutka,
Phonon density-of-states for high-T
c
(Y,RE)Ba
2
Cu
3
O
7
superconductors and non-superconducting
reference systems, Z. Phys. B: Condens. Matter 71, 437 (1988).
[284] R. Khasanov, A. Shengelaya, K. Conder, E. Morenzoni, I. Savic, and H. Keller, The oxygen-
isotope effect on the in-plane penetration depth in underdoped Y
1x
Pr
x
Ba
2
Cu
3
O
7
as revealed
by muon-spin rotation, J. Phys. Condens. Matter 15, L17 (2003).
[285] B. Friedl, C. Thomsen, and M. Cardona, Determination of the superconducting gap in
RBa
2
Cu
3
O
7
, Phys. Rev. Lett. 65, 915 (1990).
[286] A. Pashkin, M. Porer, M. Beyer, K. W. Kim, A. Dubroka, C. Bernhard, X. Yao, Y. Dagan, R. Hackl,
A. Erb, J. Demsar, R. Huber, and A. Leitenstorfer, Femtosecond Response of Quasiparticles and
Phonons in Superconducting YBa
2
Cu
3
O
7
Studied by Wideband Terahertz Spectroscopy, Phys.
Rev. Lett. 105, 067001 (2010).
[287] K. Miyake, S. Schmitt-Rink, and C. M. Varma, Spin-uctuation-mediated even-parity pairing in
heavy-fermion superconductors, Phys. Rev. B 34, 6554 (1986).
c _ Walther-Meiner-Institut
BIBLIOGRAPHY SUPERCONDUCTIVITY 139
[288] D. J. Scalapino, E. Loh, and J. E. Hirsch, d-wave pairing near a spin-density-wave instability,
Phys. Rev. B 34, 8190 (1986).
[289] A. Kampf and J. R. Schrieffer, Pseudogaps and the spin-bag approach to high-T
c
superconductiv-
ity, Phys. Rev. B 41, 6399 (1990).
[290] P. Monthoux and D. Pines, YBa
2
Cu
3
O
7
: A nearly antiferromagnetic Fermi liquid, Phys. Rev. B
47, 6069 (1993).
[291] P. A. Lee, N. Nagaosa, and X.-G. Wen, Doping a Mott insulator: Physics of high-temperature
superconductivity, Rev. Mod. Phys. 78, 17 (2006).
[292] P. W. Anderson, Is There Glue in Cuprate Superconductors?, Science 316, 1705 (2007).
[293] S. Caprara, C. Di Castro, M. Grilli, and D. Suppa, Charge-Fluctuation Contribution to the Raman
Response in Superconducting Cuprates, Phys. Rev. Lett. 95, 117004 (2005).
[294] S. V. Dordevic, C. C. Homes, J. J. Tu, T. Valla, M. Strongin, P. D. Johnson, G. D. Gu, and D. N.
Basov, Extracting the electron-boson spectral function
2
F() from infrared and photoemission
data using inverse theory, Phys. Rev. B 71, 104529 (2005).
[295] J. Hwang, T. Timusk, E. Schachinger, and J. P. Carbotte, Evolution of the bosonic spectral density
of the high-temperature superconductor Bi
2
Sr
2
CaCu
2
O
8+
, Phys. Rev. B 75, 144508 (2007).
[296] E. van Heumen, E. Muhlethaler, A. B. Kuzmenko, H. Eisaki, W. Meevasana, M. Greven, and D.
van der Marel, Optical determination of the relation between the electron-boson coupling function
and the critical temperature in high- T
c
cuprates, Phys. Rev. B 79, 184512 (2009).
[297] D. Poilblanc and D. J. Scalapino, Gap function (k, ) for a two-leg t J ladder and the pairing
interaction, Phys. Rev. B 71, 174403 (2005).
[298] T. A. Maier, D. Poilblanc, and D. J. Scalapino, Dynamics of the Pairing Interaction in the Hubbard
and t J Models of High-Temperature Superconductors, Phys. Rev. Lett. 100, 237001 (2008).
[299] P. Prelovsek and A. Ram sak, Spin-uctuation mechanism of superconductivity in cuprates, Phys.
Rev. B 72, 012510 (2005).
[300] N. M. Plakida, Theory of antiferromagnetic pairing in cuprate superconductors (Review article),
Low Temp. Phys. 32, 363 (2006).
[301] E. Pavarini, I. Dasgupta, T. Saha-Dasgupta, O. Jepsen, and O. K. Andersen, Band-Structure Trend
in Hole-Doped Cuprates and Correlation with T
cmax
, Phys. Rev. Lett. 87, 047003 (2001).
[302] P. W. Anderson, P. A. Lee, M. Randeria, T. M. Rice, N. Trivedi, and F. C. Zhang, The physics
behind high-temperature superconducting cuprates: the plain vanilla version of RVB, J. Phys.
Condens. Matter 16, R755 (2004).
[303] P. Monthoux, D. Pines, and G. G. Lonzarich, Superconductivity without phonons, Nature 450,
1177 (2007).
[304] G. B. Yntema, Minutes of the 1955 Annual Meeting Held at New York City, January 27-29, p.
1197, Phys. Rev. 98, 1144 (1955).
[305] J. E. Kunzler, E. Buehler, F. S. L. Hsu, and J. H. Wernick, Superconductivity in Nb
3
Sn at High
Current Density in a Magnetic Field of 88 kgauss, Phys. Rev. Lett. 6, 89 (1961).
[306] B. BioSpin, http://www.bruker-biospin.com/.
2013
140 R. HACKL AND D. EINZEL BIBLIOGRAPHY
[307] H. Hilgenkamp and J. Mannhart, Grain boundaries in high-T
c
superconductors, Rev. Mod. Phys.
74, 485 (2002).
[308] D. Koelle, R. Kleiner, F. Ludwig, E. Dantsker, and J. Clarke, High-transition-temperature super-
conducting quantum interference devices, Rev. Mod. Phys. 71, 631 (1999).
[309] A. P. Malozemoff, J. Mannhart, and D. Scalapino, High-Temperature Cuprate Superconductors
Get to Work, Physics Today 58, 41 (2005).
[310] J. Bray, Superconductors in Applications; Some Practical Aspects, Appl. Supercond. 19, 2533
(2009).
[311] Sumitomo Electric Industries Ltd., http://global-sei.com/super/index.en.html.
[312] Theva, http://www.theva.de/.
[313] American Superconductor, http://www.amsc.com/.
[314] Zenergy Power, http://www.zenergypower.com/.
[315] Superconductor Technologies Inc., http://www.suptech.com/home.htm.
[316] Siemens, http://siemens.de.
[317] D. Drung, F. Ludwig, W. Muller, U. Steinhoff, L. Trahms, H. Koch, Y. Q. Shen, M. B. Jensen, P.
Vase, T. Holst, T. Freltoft, and G. Curio, Integrated YBa
2
Cu
3
O
7x
magnetometer for biomagnetic
measurements, Appl. Phys. Lett. 68, 1421 (1996).
[318] I. Tristan Technologies, http://www.tristantech.com/.
c _ Walther-Meiner-Institut

You might also like