Untitled
Untitled
Untitled
S u p e r c o n d u c t iv it y
Revised Printing
T h eo ry of
S u perco n d u ctiv ity
R e v is e d P r i n t i n g
J. R . ScHRIEFFER
Florida State University
Tallahassee, Florida
C R C Press
Taylor & Francis G ro u p
Boca Raton London New York
ABP
A dvanced B o o k P rogram
The publisher is pleased to acknowledge the assistance of Paul Orban, who produced
the illustrations for the original edition. The revised edition (198.3) includes an
Appendix containing Nobel Lectures, December 11, 1972, by J. R. Schrieffer, Leon N.
Cooper, and John Bardeen, reprinted with permission of the Nobel Foundation.
Copyright © 1973 by the Nobel Foundation.
Preface xv
Preface to the Revised Printing xvi i
I
Ch a pter in t r o d u c t io n 1
1-1 Simple Experimental Facts 4
1-2 Phenomenological Theories 9
Ch a pter 2 t h e p a ir in g t h e o r y o f s u p e r c o n d u c t iv it y 24
2-1 Physical Nature of the Superconducting State 24
2-2 The One-Pair Problem 28
2-3 Landau’s Theory of a Fermi Liquid 34
2-4 The Pairing Approximation 36
2-5 Quasi-Particle Excitations 44
2-6 Linearized Equations of Motion 49
2-7 Concluding Remarks 57
ix
x C o n te n ts
Ch a pter 3 a p p l ic a t io n s o f t h e p a ir in g th eo ry 61
3-1 Justification of the Pairing Hypothesis 61
3-2 Acoustic Attenuation Rate 62
3-3 Nuclear-Spin Relaxation Rate 69
3-4 Electromagnetic Absorption 72
3-5 Physical Origin of the Coherence Factors 74
3-6 Electron Tunneling 78
3-7 Other Applications of the Pairing Theory 87
Ch a pter 4 -
electro n io n s y s t e m 89
4-1 The Electron-ion Hamiltonian 89
4-2 Bare Phonons 92
4-3 Bare Electrons 95
4-4 Bare Electron-Phonon Interaction 98
4-5 The Electron-Phonon Hamiltonian 102
Ch a pter 5 f ie l d -t h e o r e t ic m e t h o d s in t h e
6
Ch a p t er elem en ta r y e x c it a t io n s in norm al m eta ls 137
6-1 The Electron Gas with Coulomb Interactions 137
6-2 The Coupled Electron-Phonon System 148
C o n te n ts xi
Ch a pter 7 f ie l d -t h e o r e t ic m eth o d s a p p l ie d to
Ch apter 8 e l e c t r o m a g n e t ic p r o p e r t ie s of
SUPERCONDUCTORS 203
8-1 London Rigidity 203
8-2 Weak-Field Response 206
8-3 The Meissner-Ochsenfeld Effect 212
8-4 Electromagnetic Properties for Finite q and w 220
8-5 Gauge Invariance 224
8-6 The Vertex Function and Collective Modes 233
8-7 Flux Quantization 240
8-8 The Knight Shift 244
8-9 The Ginsburg-Landau-Gor’kov Theory 248
C o n c l u s io n 254
A p p e n d ix second q u a n t iz a t io n f o r m a l is m 257
A-l Occupation-Number Representation 257
A-2 Second Quantization for Bosons 259
A-3 Second Quantization for Fermions 265
A N
p p e n d ix L
o bel , 1972
ectu res 267
Macroscopic Quantum Phenomena from Pairing
in Superconductors
J. R. Schrieffer 267
Microscopic Quantum Interference Effects in the
Theory of Superconductivity
Leon N. Cooper 279
Electron-Phonon Interactions and Superconductivity
John Bardeen 300
xii C o n te n ts
In d ex 329
t \ b / A 4 * c e J - Q o s o k Q , Im 4 * c 4
in Chapter 3. This first portion of the book uses only the tech
niques of quantum mechanics which are covered in a standard
graduate course on quantum theory. While the notation of
second quantization is used as a convenient shorthand, this
formalism is reviewed in the appendix.
In Chapters 4 and 5 the many-body aspects of the coupled
electron-ion system are developed with a view to treating in a
more realistic manner the effective interaction between electrons
which brings about superconductivity. In addition, the basis for
treating strong quasi-particle damping effects important in strong
coupling superconductors is developed. In Chapter 6 a discussion
of elementary excitations in normal metals is given, which lays
the ground work for the field-theoretic treatm ent of the super
conducting state given in Chapter 7. There, the noninstantaneous
nature of the interaction bringing about superconductivity is
treated as well as the breakdown of the quasi-particle approxima
tion and the resolution of this difficulty. In the final chapter the
electromagnetic properties of superconductors are treated, as well
as the collective excitations of the system.
I should like to thank Drs. P. W. Anderson, J. Bardeen,
L. P. Kadanoff, D. J. Scalapino, Y. Wada, and J. W. Wilkins for
many helpful discussions during the preparation of this manuscript.
I am also indebted to Drs. F. Bassani and J. E. Robinson, who
prepared a set of notes covering a lecture series I gave at Argonne
National Laboratory during the spring of 1961. Much of the
material in Chapters 4 and 5 and in the appendix is related to their
notes. In addition, I would like to express my sincere apprecia
tion to Mrs. Dorothea Hofford for the speed and accuracy with
which she typed the manuscript. Finally, I should like to thank
my wife for her considerable help in preparing this book.
J . R . Sc h r ie f f e r
Philadelphia , Pennsylvania
J u l y 1964
PREFACE TO THE
REVISED PRINTING
S u p e r c o n d u c t iv it y
Revised Printing
CHAPTER 1
INTRODUCTION
M SIM PLE E X P E R IM E N T A L FA CT S
Electromagnetic Properties
The d-c electrical resistivity of materials in the soft super
conducting state is zero. This fact is established to better than
one part in 1015 of the resistance of the normal state at the corre
sponding temperature.161 At T = 0 the resistivity of a super
conductor ideally remains zero up to a critical frequency
Hwg ~ 3.5JcBT c (presumably the threshold for creating excitations
out of the condensate). In practice, the edge of the gap is
smeared and a precursor electromagnetic absorption is observed
below the edge of the gap in certain cases. At finite tempera
tures, there is a finite a-c resistivity for all a> > 0 (presumably
because of absorption by the thermally excited normal fluid if
cd < cDg). For cd » cDg, the resistivities of the normal and super
conducting states are essentially equal, independent of
temperature.
In 1933, Meissner and Oschenfeld6 discovered that a bulk
superconductor is a perfect diamagnet. Thus the magnetic field
B penetrates only to a depth A ~ 500 A and is excluded from the
main body of the material. If one (incorrectly) argues that the
vanishing zero-frequency electrical resistance implies that there can
be no electric field (of any frequency) in a superconductor,
Maxwell’s equation
V x E = - -c ^dt (1-1)
shows that the magnetic field present in the normal metal will be
“ frozen in ” when the metal becomes superconducting. This is
contrary to the Meissner effect, which states that the field is
expelled in the superconducting phase. The point is that the
superfluid gives rise to a purely inductive impedance which
vanishes only at zero frequency.9 It is this nonzero impedance
In tro d u ctio n 5
10
Ces 1\
yTc tin
10"1
- N .
10"2
T JT
1 1 .J---------1---------
Isotope Effect
As we discussed above, the isotope effect shows that lattice
vibrations play an essential role in bringing about superconduc
tivity. In particular, one finds that the critical field at zero
temperature H 0 and the transition temperature T c vary as
Tc ~ ^ ~ H0 (a ~ |) (1-2)
when the isotopic mass M of the material is varied. Thus, T c and
H0 are larger for lighter isotopes. If lattice vibrations were not
important in the phenomenon there is no reason why T c should
change as neutrons are added to the nuclei since their main effect
is to change the mass of the ions. While the value a = 0.45 to
0.50 is approximately correct for many superconductors there are
a number of notable exceptions, for example Ru, Mo, Nb3Sn, and
Os23 which have small or vanishing isotope effects. As Garland24
has shown, this does not preclude the phonons from causing the
transition. Although the actual mechanism in these materials
is not firmly established at present, it is not unlikely that the
electron-phonon interaction is the appropriate mechanism even
in these exceptional cases.
Energy Gap
There are several direct ways of observing the energy gap in
the elementary excitation spectrum of superconductors.16dj e As
we mentioned above, the threshold for absorbing electromagnetic
radiation gives a value for the energy gap.25 An even simpler
method26 (due to Giaever) is to measure the electron tunneling
current between two films of a superconducting material separated
by a thin (~ 20 A) oxide layer. As T -> 0, no current flows until
the applied voltage (times the electronic charge) V exceeds the
energy gap 2A. As the temperature is increased, a finite current
flows for V < 2A(T); however, a break in the curve persists for
8 T h e o ry o f Su p ercon d u ctivity
Coherence Effects
If one attempts to account for the rate of electromagnetic and
acoustic absorption as well as the rate of nuclear spin relaxation
in superconductors on the basis of a simple two-fluid energy-gap
model, one quickly discovers inconsistencies. Experimentally the
rate of acoustic absorption decreases monotonically as T decreases
below Tc,27 while the nuclear spin relaxation rate initially rises,
passing through a peak before dropping to zero at low tem
perature.28 If one takes the same matrix elements as in the
normal state for the coupling of the excitations with phonons
In tro d u ctio n 9
and with the nuclear spins, the two processes should have identical
temperature dependences. Therefore, at least some of these
matrix elements differ from those in the normal metal. As we
shall see, the matrix elements appropriate to the superconducting
state are linear combinations of those in the normal state.8 Since
the coefficients in the linear combination depend upon the nature
of the coupling (scalar, vector, spin), the square of the matrix
elements in the superconducting state differ for coupling the
excitations to acoustic (scalar) and electromagnetic (vector) or
nuclear magnetization (spin) variables.
1-2 P H E N O M E N O L O G IC A L T H E O R IE S
Gorter-Casimir Model
In 1934, Gorter and Casimir2 advanced a two-fluid model
along the lines which we discussed above. If x represents the
fraction of electrons which are in the “ normal” fluid and (1 — x)
the fraction condensed into the superfluid, they assumed the free
energy for the electrons is of the form
F(x, T) = x * f n(T) + (1 - x)f,(T ) (1-3)
This latter equation leads to the Meissner effect. One can see this
by considering the curl of one of Maxwell’s equations:
VxVxB = yVxJj (1-12)
where we have neglected the displacement current and the normal
fluid current J n since we are interested in the static Meissner
effect. On combining (1-11) and (1-12) one has
V2B = ! ^ B = A ?B
where London’s penetration depth AL is defined by
/ me2 \,3 1/2
(1-14)
\4:7Tnse2j
If (1-13) is applied to a plane boundary located at x = 0, the
magnetic field (parallel to the surface) decreases into the super
conductor according to
B(x) = B(0)e~XIAL (1-15)
Therefore the magnetic field vanishes in the bulk of the material
12 T h e o ry o f Su p ercon d u ctivity
A(T) [1 - (T/T,)4]1' 2
(1-17)
Thus, for T — T c, A = oo so that no flux is excluded at T c.
As T drops infinitesimally below T cy A decreases rapidly, thereby
establishing the Meissner effect in bulk specimens for all T < T c.
This temperature dependence is surprisingly close to that observed
experimentally although the results of the microscopic theory are
in somewhat better agreement with experiment than is (1-17).
The fact that the supercurrents are uniquely determined by
the magnetic-field configuration (according to the Meissner effect)
guarantees that one can apply reversible thermodynamics to
quasi-static processes in superconductors, an important fact.4
If we introduce the vector potential A, the second London
equation (1-11) can be written as
V x Js = Vx A (1-18)
As London pointed out, this equation can be satisfied by taking
In trod u ction 13
( 1-22 )
for the flux through the loop. If the path is taken within the
interior of the superconductor where B = 0, then V x A = 0 and
we may write A as the gradient of a scalar
A = VX (1-23)
14 T h e o ry of Su p e rco n d u ctiv ity
(1-26)
clearly vanishes. If a weak magnetic field is applied to the
system and Ws is unaffected to first order by this perturbation,
the paramagnetic current (1-26) continues to vanish, while the
diamagnetic current is given by
inner and outer cylinders, so that the flux through the hole of the
outer cylinder is also equal to 0. The inner cylinder acts only
as a shield to ensure that no magnetic field touches the physically
interesting outer cylinder. Let be the wave function for the
outer cylinder when there is no flux trapped, 0 = 0. To deter
mine the wave function xf/0 in the case 0 / 0, we note that the
vector potential in the outer ring is in the 6 direction and has the
value
>1 / x 0 1 0 /0 9 \ _ / 0 0\
( 1 -2 8 )
—
60
he
= integer (1-30)
or 0 is quantized to the London values
tf>n = (n = 0, ±1, ± 2 ,...) (1-31)
To complete the argument, suppose that the inner cylinder is
made normal so that the magnetic field fills the entire hole in the
outer cylinder. Owing to the Meissner effect, the magnetic field
will penetrate only a small distance (~ 5 .10“6 cm) into the outer
cylinder. Therefore the above argument should continue to hold
since this small perturbation should not affect the wave function
W appreciably (particularly if London’s “ rigidity” is effective).
On the basis of this argument London concluded that the flux
trapped through any hole in a massive specimen is quantized to
multiples of hcje.
In 1953 Onsager32 suggested that the actual value of the
flux quantum might be one-half this value, presumably because
In tro d u ctio n 17
in agreement with experiment. The fact that these are the only
allowed values for 0 follows from the BCS pairing theory since
other values of 0 lead to an extraordinarily high energy of the
electron system and are therefore unstable. The problem of flux
quantization is discussed further in Chapter 8.
Pippard’s Nonlocal Generalization of the London Theory
The basic equations (1-10) and (1-11) of the London theory
are “ local” in the sense that they relate the current densities and
the electromagnetic potentials at the same point in space. On
the basis of numerous experimental results, Pippard33 concluded
that these local relations must be replaced by nonlocal relations
giving the currents at a given point in space as a space average
of the field strengths taken over a region of extent £0 ~ 10" 4 cm
about the point in question. One of the most compelling argu
ments for this generalization is that the penetration depth A
increases appreciably if a sufficient amount of impurity is intro
duced into the material. This effect sets in when the mean free
path I of electrons in the normal state falls below a distance £0,
known as Pippard’s “ coherence” length. As we shall see, £0 is a
18 T h e o ry o f Su p e rcon d u ctivity
measure of the size of the pair bound state from which the super
fluid wave function is constructed. In the microscopic theory it
is related to the energy gap 2A by £0 = Hvf IttA, where vF is the
Fermi velocity. On the other hand, in the London theory A is
not expected to be appreciably affected by impurities, particularly
near T = 0, where all of the electrons are condensed. In choosing
a form for the nonlocal relations, Pippard was guided by Chamber’s
nonlocal expression,34 relating the current density and electric
field strength in the normal metal
R = r - r' (1-34)
where a is the long wavelength electrical conductivity. Chamber’s
expression is a solution of Boltzmann’s transport equation if the
scattering mechanism is characterized by a mean free path I.
For fields varying slowly over a mean free path Z, (1-34) reduces
to Ohm’s law J = oE. With Chamber’s expression in mind,
Pippard assumed that London’s equation
should be replaced by
(1-36)
The effective coherence length £ is given by
Ginsburg-Landau Theory
In 1950 Ginsburg and Landau36 proposed an extension of the
London theory which takes into account the possibility of the
superfluid density ns varying in space. They phrased the theory
in terms of an effective wave function !P(r) which we normalize
such that the local density of condensed electrons is given by
|<F(r)|2 = (1.38)
Tl
+ d3r +
J 8tt
d 3r (1-47)
with respect to W(r), one finds the constitutive equation of the
Ginsburg-Landau theory
THE PAIRING TH EO RY OF
SUPERCONDUCTIVITY
2-1 P H Y S IC A L N A T U R E O F T H E S U P E R C O N D U C T IN G
STA TE
This qualitative difference in the excitation spectra is paral
leled by a qualitative difference in the wave functions of the N-
and ^-phases of metals. In the AT-phase, the probability that
two single-particle states i and j are simultaneously occupied
24
T h e Pairing T h e o ry o f Su p ercon d u ctivity 25
is a smoothly varying function of k and k' (so long as one does not
cross the Fermi surface in varying k or k'). Here \N} represents
a typical state in the normal phase and nfcT is the operator which
measures the number of electrons in k f , etc. (see the Appendix).
In the superconducting phase,8 the corresponding probability
P kk's = O SK tT vJS) (2-2)
is also a smoothly varying function of k and k' except when k
and k' are related by the “ pairing” condition. This condition
states that for a given state k, there exists a single mate k such
that the probability P kks is larger than P kk,s by a finite amount,
for all states k' in the vicinity of k. This singular behavior of
the two-particle correlation function, which has been stressed by
Yang,39 is no doubt the sort of picture F. London had in mind
when he suggested that superconductivity is due to a condensation
of the electrons in momentum space.1 When proper account is
taken of residual interactions conventionally neglected in the
description of the normal state, these “ pairing correlations”
leading to superconductivity emerge in a natural manner. Above
the superconducting transition temperature, the pairing correla
tions are broken up by thermal fluctuations and play no important
role in the normal phase.
It is essential to realize at the outset that the lowering in
energy of the $-phase due to interactions between mates (say
k | and k j ) of a given pair depends critically on the choice of
mates (k' f and k' | ) for other pairs. In fact, the energy gap and
most of the observed properties of the superconducting phase
would be absent were it not for strong correlations between the
pairs. The reason for the simple BCS model working so well is
that in real metals these pair-pair correlations are almost entirely
due to Pauli principle restrictions rather than correlations due to
true dynamical interactions between the pairs. This fact allows
one to treat the system in lowest order as if dynamical interactions
26 T h e o ry o f Su p ercon d u ctivity
2-2 T H E O N E -P A IR P R O B L E M
To understand the origin and consequence of pairing corre
lations, it is helpful to consider the problem, first studied by
Cooper,41 of a pair of electrons interacting above a noninteracting
Fermi sea of electrons via a velocity-dependent nonretarded two-
body potential V. Thus, all but two of the electrons are assumed
to be noninteracting. The background electrons enter the total
problem only through the Pauli principle by blocking states below
the Fermi surface from participating in the remaining two-particle
problem. If we measure the kinetic energy ek relative to its
value at the Fermi surface, only states with ek > 0 are available
to the interacting pair of electrons. Since the system is assumed
to be translationally invariant and one neglects spin-dependent
forces, the center-of-mass momentum hq of the pair and the total
spin S are constants of motion. The orbital wave function of the
pair can then be written as
^(r,, r2) = <p,(p)e">-R (2-3)
where the relative and center-of-mass coordinates are defined
by p = r: - r2 and R = (rx + r2)/2, respectively. The relative
coordinate wave function is symmetric for the singlet spin state
(S = 0) and antisymmetric for the triplet states (S = 1). In
the limit q —> 0 the relative coordinate problem is spherically
symmetric so that <p(p) is an eigenfunction of angular momentum
and can be labeled by the angular momentum quantum numbers
I and ni. For q / 0, the component of angular momentum
along q and parity remain good quantum numbers but I is no
longer sharp.
For simplicity, we first consider the zero momentum states
q = 0, so that ip can be expanded as
</'(ri> r2) = <p(p) = 2k afceik'p = 2k akeik''ie ~ik'r2 (2-4)
In (2-4), the sum is restricted to the available states (ek > 0).
Since the factors eik ri and e can be thought of as single
~ i k r 2
particle states of momentum k and —k, we see that the pair wave
function is a superposition of configurations in each of which a
definite pair state (k, —k) is occupied.
The Pairing T h e o ry o f Su p e rcon d u ctivity 29
«k = a k Y,m(Qk) (2-9b)
Equation (2-9a) can be written as
(2-10a)
where the constant C is defined as
(2-10b)
By substituting (2-10a) into the definition (2-10b) one obtains
the equation
1 = A, 2 IV I 2
k VV I m ~
53 <2-11>
determining the energy eigenvalues Wlm. If we work in a large
but finite box the single-particle energies ek form a discrete set
so that when W passes from below to above 2ek, &(W) jumps
from —oo to oo. As IT moves toward the next higher value of
2ek, O(W) again approaches —oo and jumps to + oo as W passes
through this higher value. The function <P(W) is shown sche
matically in Figure 2-2. As W passes through the origin to
negative values (i.e., the region of bound states) 0 (W ) increases
from — oo to zero as shown. The eigenvalues Wim are given
by the intersections of &(W) with the constant function 1/A*,
as shown for both positive (repulsive) and negative (attractive)
A,. While the eigenvalues in the continuum are trapped between
the unperturbed energies 2tk and approach the unperturbed
energies as the size of the box goes to infinity, a state is bound
The Pairing T h e o ry o f Su p ercon d u ctivity 31
F IG U R E 2-2 A plot of the function <P(W) [see (2-1 I)] which determines
the eigenenergies in C o o p e r’s one-pair problem. For a repulsive
interaction (A, > 0), all states are trapped in the continuum, while for
an attractive interaction, a bound state is split off.
32 T h e o ry o f Su p ercon d uctivity
w q\ = \w0\ -
vFhq
2
(2-15)
where
as above. Thus, the pair energy increases linearly with the center-
of-mass momentum in the limit q -> 0, rather than as q 2, as one
might expect. As Cooper pointed out, the drift of the pair with
respect to the noninteracting Fermi sea strongly reduces the
binding energy of the pair due to the reduced density of low-
energy states available to the pair. This effect dominates the
q2 increase of kinetic energy for small q.
If | W01is imagined to be of order k T c, (2-15) shows that the
pair would have lost most of its binding energy when
so that
(2-25)
is a properly normalized state. By expanding out the infinite
product one sees that |0O> has a nonvanishing amplitude for all
even numbers of electrons, 0, 2, 4, .... However, by choosing gk
appropriately the mean number of particles described by |0O> can
be adjusted to be the required number N 0. As in the grand canon
ical ensemble, one can show that the width of the distribution is
of order N 01/2 so that particle number fluctuations cause no
difficulty in a large system.
Since we want to minimize the ground-state energy subject
to the constraint
<-£0|2\U^o> = <<Ao| I nks\4<0> = N 0 (2-26)
and
u kv k = 2 j |“ (2-30c)
where E k is defined by
E k = +[(cfc - /x)2 + ^ 2]1/2 (2-30d)
As we shall see, E k turns out to be the energy required to create a
quasi-particle of momentum k in the superconducting state. The
“ energy-gap” parameter A k satisfies the integral equation
| m~, ( 2 - 3 0 e )
2-5 Q U A S I-P A R T IC L E E X C IT A T IO N S
To find the excited states of the BCS reduced Hamiltonian
(2-17) we consider adding an electron to the system in the state
p f (its mate —p j being empty). The only effect of this process
is to block the pair state (p f , —p | ) from participating in the
pairing interaction (due to the Pauli principle). Since —p j is
assumed to be empty, the electron in p f cannot be scattered out
of this state, due to the form of the pairing interaction (2 -2 0 ).
Of course, residual interactions not explicitly included in H red will
allow this process to take place; however, these interactions appear
to have a small effect on the excitation spectrum (since they are
implicitly included in the quasi-particles of the normal state).
The quasi-particle energy is defined to be the total excitation
energy of the system when the extra electron is added to the sys
tem. From (2-28) we see that by deleting the pair state (p f ,
—p | ), the energy of the interacting pairs is increased by
(2-42)
To this we must add the single-particle energy €p of the added
electron. The total excitation energy is given by
ep[l - 2vp2] + 2 ApUpVj
p^pVp (2-43)
where we have used the gap equation (2-30e) to simplify the
interaction energy term. If we use the results (2-30) for uk and
vk (with ft = 0 ), we find the excitation energy
(2-44)
Thus, the parameter E p defined by (2-30c) is just the energy
required to create a quasi-particle in state p f . A plot of E p vs. p
is given in Figure 2-3. The minimum energy required to add an
electron to the system is A kf = A 0 ~ 1 0 " 3 — 1 0 " 4 ev. In prin
ciple the chemical potential /z should be shifted a small amount to
ensure <A> = N 0 + 1 in the excited state; however, this correc
tion has negligible effect in a large system.
The Pairing T h e o ry o f Sup e rcon d u ctiv ity 45
where we have chosen phases so that all quantities are real. The
secular equation for (2-61) and (2-63) is
Qp €p Ap
= Qp2 - € P2 - Ap2 = 0 (2-64)
Ap Qp 4- €p
where the parameter Ap is defined by
dp = “ 2 (2-65)
k
The formal similarity between these results and those of the last
section is complete if we require that | 0 , A> be the ground state
of the system, that is,
ypt| 0 , A> = 0 (2-70a)
and
y _pi | 0 , A ) = 0 (2-70b)
Thus, the y + ’s create noninteracting fermion excitations from the
“ vacuum state” 10 , A>.
By inverting (2-67a), (2-67b), and their Hermitian conjugates
to solve for the c-operators in terms of the y’s, one finds from the
definition of B k (2-62c) the relation
bh = ukvk =A (2-71)
On combining this result with che definition of Zlp (2-65) we find
an equation determining the parameter zJp:
( 2 ' 7 2 )
and
Bk = tanh p E k (2-75)
By inserting this result into (2-65) we obtain the finite-temperature
BCS gap equation
-2 v
A k i . PEk
/” 2E k U n h - r
‘
(2-76)
This finite-temperature treatm ent of the pairing theory is
entirely equivalent to the BCS treatment, which, as we mentioned
above, gives an exact account of the system described by the
reduced Hamiltonian (in the limit of large volume). If Vkp is
approximated by (2-32), A k is again of the form (2-33) where
A0(j3) satisfies
1 _ fWc de
tanh A + V ) 1' (2-77)
A0)T = Jo > 2 + ^o 2 ) U2
As T increases from zero, A 0 decreases as shown in Figure 2-5,
vanishing at the transition temperature T c. Thus, T c is given by
de
tanh 2 kRT r
(2-78)
APPLICATIONS OF THE
PAIRING TH EO RY
= SoAo^oAo - (3-12)
where the acoustic attenuation rate is given by
«?oAo = 477 2 l£jpp'A0|2n2(P> P'Km - V t)
pp'
S ( E P- - E p - w,oAo)
(3-13)
and S isthe spontaneous emission rate. When the sum is
performed, the occupation numbers can be replaced by their local
average values and one has
ato*o = 4 n pp'
2 |£pp-a„|M p, P')(/p - fp') H E P. - E p - w?oAo)
(3-14)
where p' = p -f q0 + K and K is a reciprocal lattice vector. The
spontaneous emission rate S'qoAo is given by (3-11) with N QoAo = 0.
The expression for a simplifies if we assume that g depends
only on the momentum transfer (i.e., q 0 for normal processes).
For | q01 small compared to kF only normal processes enter the
sum with appreciable weight, so that a is given by
«<oA0 = 27r |^ oAo|2 2 (X+ ~ /P‘>
x 8(EP. - E p - u>qoXo) (3-15)
where p' = p + q0, and the relation
»’(p, p') = (upup. - vpvp. f = \ (l + €p€p'E J ’ Ap) (3-16)
has been used. Since the speed of sound is small compared to
the Fermi velocity, —\0~3vF in typical cases, energy and momen
tum conservation require that q0 be essentially tangent to the
66 T h e o ry o f Su p ercon d u ctivity
x [f(E ) - f ( E ’)]
A p p lic atio n s o f the Pairing T h e o ry 67
T /T c
F IG U R E 3-2 The longitudinal acoustic-attenuation coefficient in the
superconducting state relative to that in the normal state compared
with the result of the simple pairing theory.
68 T h e o ry o f Su p e rco n d u ctiv ity
EE'
{E2 - A2)1I2(E'2 - A2)112
3-3 N U C L E A R -S P IN R E L A X A T IO N RA T E
The above calculation of the acoustic attenuation rate is
easily modified to account for the relaxation rate of oriented
nuclear spins due to their hyperfine coupling to the valence elec
trons. The interaction for a given nuclear spin I is of the form
Hi s = ^ 2 a k * a k[h(ck't X t - c^+Cfci)
k, k'
+ I +Ck'i+Ck\ + I - Ck"\+Ckl] (3-21)
where a k is proportional to the amplitude of the Bloch function
Xk(r ) the nuclear site in question, so that a _ k = afc*, and
I ± = Ix ± H y 66 To calculate the rate at which a given nucleus
decreases its 2 -component of spin we observe that the Zeeman and
hyperfine energies are in general small compared to the energy
gap so that only quasi-particle spin-flip processes enter. We
consider a typical initial state |/> excited at the temperature T ,
as in the preceding section, and notice that the final states are of
the form
7p2
t + 7pi i 1 -0 nuclear spin flips down (3-22a)
Vpi i +Vp21 1^) nuclear spin flips up (3-22b)
70 T h e o ry of Su p e rcon d u ctivity
P 1 .P 2 * \ ± j P x £ j P2 >
x / Pld ~ f P2)Z (E P2 - E Pl - w) (3-28)
if we neglect crystalline anisotropy, since l2(px, p2) is given by
the ratio ccsjaN for spin relaxation rate is its predicted rise as T
drops below T c as shown in Figure 3-4. This result is distinctly
different from the result for the acoustic attenuation rate. The
only difference in the predicted rates for the two types of processes
is that the coherence factor n2 appears in the acoustic case while
I2 appears in the spin relaxation case. As we mentioned, the
anomalously small matrix element for quasi-particles near the
Fermi surface being scattered by phonons exactly cancels out
the large density of quasi-particle states in this vicinity. On the
other hand, the quasi-particles are coupled to the nuclear spins
with essentially the -same strength as single particles in the
normal state so that the large density of quasi-particle states near
the Fermi surface leads to an increased relaxation rate. Of
course, at low enough temperatures few quasi-particles are excited
so that the relaxation rate goes to zero as T -> 0 . It is clear that
a simple energy-gap form of a two-fluid model could not account
for the sharp drop in the acoustic attenuation rate near T c and
simultaneously a rapid rise of the nuclear-spin relaxation rate.
It is interesting to note that the beautiful experiments of Hebei
and Slichter were being carried out during the period when the
BCS theory was being formulated, and that their experiments
gave one of the first substantiations of the detailed nature of the
pairing correlations which are basic to the theory.
3-4 E L E C T R O M A G N E T IC A B S O R P T IO N
Another example of resonant energy absorption is the real
part of the electrical conductivity oq in a thin film. If we de
scribe the electromagnetic field by the vector potential
A(r, t) = A0ei(«'r- w0 + c.c. (3-32)
the first-order coupling is of the form
H A{t) = - ^ Jj(r) • A(r, t) <Pr
As above, there are two terms in a given sum which lead to the
same quasi-particle transition; for example, the combination
CP'T+CPT - c_pi+c_pU = Z(p, p')(yPM+ypt - y - p i +y - p^)
- p(v> P')(yp't+y-pi + ^ + yPiy - v ’\R + ) (3-34)
enters here due to (2k -b q) -> —(2k -f q) as k —> —(k + q) and
(k + q) -» -k . The first terms on the right-hand side of (3-34)
lead to quasi-particle scattering and contribute at T ^ 0 while
the last terms lead to creation or destruction of two quasi-particles.
They contribute only if to ^ 2A.
For simplicity we consider only T = 0 so that absorption
occurs only for to ^ 2 A. By calculating the rate of photon
absorption just as we did for phonon absorption, one finds
<71S 1 E) - A2] dE
I f f (a , -
+ (” 6)
where x — oj/2 A ^ I. A plot of the theoretical ratio is shown in
Figure 3-5 and is in quite good agreement with experiment.
For general w and temperature, the integrals must be done
numerically. For to « 2 A, a shows a rise as T decreases below
T c, as in the case of nuclear spin relaxation, followed by the low-
temperature exponential drop. For to > k T c/2, the ratio no
longer shows a peak. Several cases have been worked out by
Miller.70 Note: At the time of the first printing of this volume, it
had been reported that a precursor absorption (co/2A~0.85) was
observed at low temperature.71,72 Subsequent measurements and
improved processing of the data showed that the precursor was
an artifact. Collective modes which were proposed to account for
the precursor gave too weak an absorption to agree with experi
ment (see Ch. 8).
74 T h e o ry o f S u p ercon d u ctivity
hco/2 J ( 0)
F IG U R E 3-5 The frequency dependence of o1/<jn and a2jan at T = 0
as calculated by Tinkham from the w ork of Mattis and Bardeen.
3-5 P H Y S IC A L O R IG IN O F T H E C O H E R E N C E
FACTORS
Aside from the coherence factors, one might have guessed
the results of this chapter on the basis of a simple single-particle
energy-gap model for the superconductor. The physical origin
of the coherence factors is, however, fairly simple.
Suppose we are interested in a process in which a quasi
particle is scattered from an initial state, say k f , to a final state,
say k' f , by absorbing a boson (a phonon or photon) of momentum
k' — k. For simplicity we assume there are no quasi-particles
in the states —k j , k' f , and —k' j initially. (The argument is
easily generalized to include excitations in these states.)
As we saw in Chapter 2 , a quasi-particle in k f (and none in
—k j ) corresponds to an electron definitely occupying the Bloch
state k f (i.e., with unit probability) and the mate state —k |
being definitely empty. The pair state (k' f , —k' j ), with no
quasi-particles in it, has a probability amplitude uk>of k' f and
—k '| being empty, and an amplitude vk> of these states being
A p p lic atio n s o f the Pairing T h e o ry 75
k' t
a m p l it u d e u k> a m p l it u d e vk>
k k' t
a m p l it u d e u k a m p l it u d e vk
F IG U R E 3-6 (a) and (b) The two configurations entering the wave func
tion for a state with a quasi-particle in k f . (c) and (d) The tw o config
urations entering the wave function for the state with a quasi-particle
in &f , showing how a) -> c)
and b ) - > f ) when the electrons couple
to a field which does not flip the electronic spin (acoustic or electromag
netic fields are examples), (e) and (f) The tw o configurations entering
for a state with a quasi-particle in — j , showing how a)->e) and
b) /)w h en the electrons couple to a field which flips the electronic spin
(the hyperfine coupling involved in nuclear spin relaxation is an example).
76 T h e o ry o f Su p ercon d u ctivity
a m p l it u d e 1
Falicov, and Phillips ,75 and more recently by Prange .76 In this
approach, one describes the system by an effective Hamiltonian
H = Hx + Hr + Ht (3-37)
where H x and H r are the full many-body Hamiltonians for the
left and right metals in the absence of tunneling, and H T is a
one-body operator which transfers electrons between the two
metals,
H t = 2 {r ^kk’ck'sr+cksl + H.c.} (3-38)
kk's
Bardeen has shown that T kk>is given by the matrix element
of the current density at center of the oxide taken between single
particle states which decay exponentially as one moves into the
oxide layer. Harrison 77 has evaluated T kw within the WKB
approximation and finds
I\Tmkk]219 = 1
7 2 r
>fc|<T' eX P
Pi Pi
where p± is the one-dimensional density of states for motion
r
J xl
Jc±(x) dx (3-39)
H r \Pr> = ‘ B \Pr> ^ }
and the energies ea and are measured relative to the ground-
state energies in I and r, respectively. At zero temperature
electrons cannot tunnel in the reverse direction due to energy
80 T h e o ry o f Su p ercon d u ctivity
conservation. From (3-40) one can readily see that the current
density is proportional to
rv
I(V ) oc j V N T+r(E)N T_ l(V - E )dE (3-42)
where
N T+r(E) = ^ |<|Sr|<v + |0r>|2 8(e/, - E)
k .0
s iVr(0) P dekPS +\ k , E ) (3-43a)
and
J - 00
where we have used the fact that if a given ek makes the argu
ment of the delta function vanish, so does €k> = —ek so that
uk2 + uk>2 = 1 . This simple manner in which the coherence
factors vanish in the expression for the tunneling current was first
pointed out by Cohen, Falicov, and Phillips . 75 It is interesting to
note that N T+ is just the density of quasi-particle states which
we used earlier in this chapter, as one would have guessed on the
basis of a simple energy-gap model without coherence effects.
In a similar manner one finds for a nonretarded pairing
potential model
(3-46)
where the coherence factors vk2 vanishes in the result just as u 2
did in N T+.
As we shall see in Chapter 7, the expressions (3-45) and (3-46)
are incorrect in real metals due to the strong retardation effects
associated with the phonon interaction between electrons. There
one finds the simple result
(3-47)
as opposed to N(0)(E — \ dA2jdE)l[E2 — A2(E)]112, which follows
from (3-46).
Returning to expression (3-42) for I(V ) we find for tunneling
between a normal and a superconducting metal that
(3-48)
if we use (3-47), where I s and I N are the currents flowing
when the superconductor is in the S- or A-state, respectively.
Therefore, the tunneling experiment can give detailed information
about the energy dependence of the gap parameter.
The finite-temperature tunneling current can be treated in an
analogous manner if one includes a thermodynamic average over
82 T h e o ry o f Su p ercon d u ctivity
initial states rather than using | 0 f >|0 r>, and further one includes
currents from I to r and r to Z.78, 79 The tunneling current can
then be expressed in terms of the spectral weight functions for the
thermodynamic Green’s functions. In the simple case of a non
retarded two-body pairing potential, the expression reduces to the
golden rule result for a simple energy-gap model without coherence
effects, as for T = 0 . A typical I - V characteristic for tunneling
between a normal and a superconducting metal is shown in Figure
3-8 for several temperatures for an energy independent A . Experi
mental curves are in general agreement with theory although small
deviations exist, some of which we shall discuss below.
A pictorial view of the one-particle tunneling process between
a normal and a superconducting metal is illustrated in Figure
3-9a and b. In 3-9a, an electron in k f , beneath the Fermi
surface in the normal metal tunnels through the oxide to state
k ' f above the Fermi surface in the superconductor. In the pro
cess a hole is left behind in Zgiving an excitation energy ea = |efc|
for this metal. In addition, a quasi-particle is placed in k ' f
giving an excitation energy e# = E w = (ek 2 + Ak 2)112 for the
Figure 3-9 (a) and (b) Tw o final states for a given initially occupied
state k \ . These processes enter the expression for the single-particle
tunneling rate between a normal and a superconducting metal.
84 T h e o ry o f Su p ercon d u ctivity
Figure 3-10 (a) and (b) Processes analogous to those in Figure 3-9,
however the tunneling is between superconductors here.
— (W \W ) — 2 (3-50)
where H T{2) is the second-order tunneling Hamiltonian given
by
= Ht e ^ it 0Ht (3’51)
and
h J y = 4|<0v+1|^ r<2)|0v>| (3_52)
To find the current, note that the rate of transfer of pairs is
d<» IdEA J. ,
~ w " \ a f ) ■t <sm “> <3-53*>
where the average is taken in a wave-packet state formed from
the s,incomplete analogy with the tight-binding approach
to the one-electron theory of metals. In the absence of an
applied bias, the momentum ha (canonically conjugate to the
pair number v) is a constant of motion; however, for V ^ 0 one
has
d(hay
-V =2F (3-53b>
86 T h e o ry o f S u p ercon d u ctivity
From (3-53a) and (3-53b) it follows that the rate at which electrons
flow across the barrier is
J (t) = = J 1 sin + «0 (3-54)
so that an alternating current of frequency 2V/h = 483.6 Mc/sec/
/ivolt is expected to flow for V ^ 0 , while for V = 0 , a steady
current is expected, according to (3-53a). The d-c effect has been
observed by Rowell and Anderson .85 As Josephson pointed out ,83
the current is sharply reduced when a magnetic field is applied
to the junction of such a strength that a multiple of the flux
quantum occurs in the junction. This effect has been observed
by Rowell. 86 The a-c effect has been observed by Shapiro .860
Tunneling experiments by Burstein and Taylor 81 show that
in many cases an excess current between two superconductors at
low reduced temperature begins at an applied bias Ax or Ar, that
3-7 O T H E R A P P L IC A T IO N S O F T H E P A IR IN G
THEORY
In this chapter only a few of the simplest applications of the
pairing theory have been discussed. In Chapter 8 we shall discuss
the electromagnetic properties of superconductors, i.e., the Meiss
ner effect, the persistence of supercurrents, magnetic flux quanti
zation, etc., as well as the paramagnetic spin susceptibility and the
resultant Knight shift. Further support for the pairing theory
comes from the distinctly different effects magnetic and nonmag
netic impurities have on the energy gap. As Anderson showed,
nonmagnetic impurities do not smear the gap edge as one might
intuitively expect; on the contrary, nonmagnetic impurities remove
88 T h e o ry o f Su p ercon d u ctivity
ELECTRON-ION SYSTEM
5 .2 ,1 ^ + ? &
+ \ 2 W (Rv, Rv,) + 2 U(r„ Rv) (4-1)
V*V ' I, V
4-2 BARE P H O N O N S
We introduce bare-phonon coordinates Qq x of wavevector q
and polarization A which describe the deviations §RVof the ions
from their equilibrium positions Rv°. This is done by performing
the canonical transformation
2 v i r + \ 2 w* - ^ \ l
~ q , A
+
+ const- W Q 'S Q J
(4-10)
V
(4-14)
then the states Xk are n°t in general orthogonal to the core states.
Even if we arrange U0 so that the “ conduction band” solutions of
96 T h e o ry of Sup ercon d uctivity
(4-14) are orthogonal to the core states when the ions are in their
equilibrium positions, the orthogonality is not maintained when
the ions vibrate. A partial solution to this difficulty has been
given by Wilkins,94 using the pseudo-potential method of Kleinman
and Phillips. This work was generalized to the many-body
problem in the work of Bassani, Robinson, Goodman, and the
author , 95 who treat the case of rigid cores described within the one-
electron approximation. We shall not discuss this treatment here
because of the mathematical complications necessary to carry
through the analysis. It suffices to say that an auxiliary wave
field describing the conduction electrons can be introduced in
such a manner that the conduction and core states are properly
orthogonal even if the cores vibrate. The equations of motion
of this auxiliary wave field are the same as for the original wave
field except for a redefinition of the potentials involved. Since
these potentials are difficult to estimate from first principles at
present, we shall simply disregard the above complication and
proceed using the one-electron states (4-14) as the bare conduction
electron states.
To make the states Xk precise, we must define U0. In order
that the electrons are not scattered by the lattice when no phonons
are present, UQshould include the electron-ion interaction with the
ions fixed on their equilibrium positions. Since the ion system
has a large positive charge, this leads to a very large negative
potential acting on a conduction electron. Since the Coulomb
interactions with the remaining conduction electrons cancel most
of this interaction, we include in U0 the potential due to the re
maining conduction electrons occupying a standard configuration.
This configuration could be a uniform distribution of electronic
charge or the distribution given by treating the conduction elec
trons within the Hartree-Fock approximation. Of course the
better one does in choosing U0, the less there is to take into account
as coupling between the bare particles. In any event, U0 should
be chosen to have the periodicity of the lattice (although it may not
be diagonal in the coordinate representation) so that Bloch’s
theorem 96 holds
Xk(r + a) = eikaxte(r) (4-15)
E le c tro n -lo n System 97
4-4 B A R E E L E C T R O N - P H O N O N IN T E R A C T IO N
We included in H 0 the interaction of the electrons with the
ions in their equilibrium positions. The difference between this
potential and the full electron-ion potential remains as a pertur
bation (along with several other terms). It turns out that it is
sufficient for most purposes to expand this difference in powers of
the ionic displacements 8 RV, and retain only the leading term.
Thus, the bare electron-phonon interaction is of the conventional
form for a boson-fermion coupling, that is, linear in the boson
field, bilinear in the fermion field.
A reliable first-principles calculation of the coupling is not
possible at present for most superconductors, since one requires
accurate one-electron wave functions as well as reliable ionic
E le c tro n -io n System 99
where we have used (4-2) and have denoted U(rt, Rv) by UltV as
well as U (rp Rv°) by U%tv°. For given values of q and Athe matrix
element of this potential between bare electron states k and k' is
- ( N M
Q c) v 2 2 <*'l V . t g * ) • e » e * - * . 0 (4-21)
where Zce is the total ionic charge per unit cell and Qp =
(477-ArcZc2e2/M c)112 is the ionic plasma frequency. In the long
wavelength limit (q —> 0 ), where the model is presumably reason
able, the coupling of the electrons to bare longitudinal phonons is
singular. The singularity is clearly due to the long-range Cou
lomb force. When screening is taken into account, the dressed
interaction vanishes as q -> 0 , again showing the importance of
screening in metals. We note the relation
2g 2 4:7re2
(4-29)
hOP = ~Y~
holds for longitudinal phonons in jellium. This relation has
significance in superconductivity since the left-hand side turns
out to be related to the bare electron-electron interaction resulting
from the exchange of virtual phonons. The equality (4-29) states
that the phonon attraction is exactly cancelled by the Coulomb re
pulsion if the electrons scatter without changing their energy (i.e.,
the net interaction vanishes in the static limit). This result is
peculiar to jellium; however, the relative scale of the phonon and
Coulomb interactions in real metals is roughly set by (4-29).
The jellium model has the added simplicity that the electrons
and transverse phonons are uncoupled since their interaction is
proportional to q • e qAi which vanishes in this case.
Actual metals are considerably more complicated than
jellium since transverse phonons play a strong role in umklapp
processes96 (transitions in which the momentum transfer k' — k
lies outside of the first Brillouin zone). Transverse phonons can
also enter normal (non-umklapp) processes if the electronic energy
contours in k-space are not spherical or if q is not in a symmetry
direction. Also, there is an electromagnetic coupling between
the transverse phonons and the electrons .67 In addition, the
bare longitudinal matrix elements are certainly more complicated
than (4-28) since they will reflect crystalline anisotropy as well as
details of the core potential and the behavior of the Bloch functions
near the cores. The reader is referred to Ziman’s book for further
details about the electron-phonon interaction in real metals. It
102 T h e o ry of Sup ercon d u ctivity
FIELD-THEORETIC
METHODS IN THE
MANY-BODY PROBLEM
5-1 T H E S C H R O D IN G E R , H E ISE N B E R G , A N D
IN T E R A C T IO N P IC T U R E S
In elementary discussions of quantum mechanics one usually
works in the “ Schrodinger picture” in which the dynamical
variables are taken to be time independent so that the wave
function contains the time dependence of the problem. In this
picture, the wave function ¥$(£) satisfies
= H{t^ {t) {5'l)
While the dynamical variables are independent of time, the
Hamiltonian may contain an explicit time dependence because
103
104 T h e o ry o f Sup ercon d uctivity
ih d^ = [0h(0, H] +
ir i h ( 5 ' 5 )
5-2 T H E G R E E N ’S F U N C T IO N A P P R O A C H
In the many-body problem we are ultimately interested in pre
dicting such quantities as the thermodynamic and mechanical
properties of the system as well as nonequilibrium properties such
as electrical and thermal conductivities, and absorption of quanta
of external fields. It is clear that we cannot determine these
quantities by solving for the exact many-body eigenfunctions
except for extremely simple systems; even if these functions were
available they would be hopelessly complicated unless expressed
in a form suitable for calculating a particular property of the system.
One would prefer to work with dynamical quantities which are
more closely related to experiment and contain less information
than the full wave functions. One would then approximate
these dynamical quantities directly rather than working with the
Y 9*. Quantities which satisfy these conditions are the Green’s
functions of quantum field theory .9 1 , 99 The one-electron Green’s
function gives information about the spin and charge densities
106 T h e o ry of Sup ercon d uctivity
2?t
-P (5-24)
0
J~< Po irjp
The integral is evaluated by using Cauchy’s theorem. For r ^ 0
the contour can be closed around an infinite semicircle in the
upper half of the complex p 0-plane, as shown in Figure 5-1,
since e ~ ipo(l~d) vanishes on this added piece of contour. For
r > 0 the contour may be closed in the lower half-plane for similar
reasons. Because of the factor irjp, the pole is in the upper half
plane for p < p F and in the lower half-plane for p > p F. Thus,
for r > 0 we obtain
G (p . t ) =
0 P < Pf
-ie ~ i€pl p > pF
while for t ^ Owe find
l€pT p < p Fte
0
G (p , ) =
P > Pf
t
5-4 S P E C T R A L R E P R E S E N T A T IO N O F G(p, r)
The one-particle Green’s function G for the interacting Fermi
gas is closely related to G for the noninteracting system. To see
this, consider the Green’s function for spin-up particles:
<?(P, t ) = -t'< 0 |T { Cp(r)cp + (0)}|0>
—t<0|cp(0)e ~ iH,c„ + (0) 10>ei£°nt r > 0 (5-31)
i<0|cp + (0)eiHtcp(0 )|0 > e-i£o"' T sj 0
where E 0n is the ground-state energy of the interacting n-particle
system. By inserting between c and c+ a complete set of eigen
states |!Fmn±1> of H for the n ± 1 particle system, we find
H 2 l(«P +)m.o|2« -i(£"n+1- £»")l T> 0
0 ( P, r) = m (5-32)
i ' 2 \ ( c p ) m . o \ 2 e i < E m" ~ 1 ~ E ° n ) l T<°
The matrix elements are defined by
(cP+L.o = < ^ mn +1 |Cp+l^on> (5-33a)
K ) m.o = <Vm" - 1 |cp|'F0") (5-33b)
that is, matrix elements of the operators for bare electrons taken
between exact eigenstates of the full Hamiltonian. For con
venience we introduce the energies o>mn ± 1 defined by
£ m" +1 - E 0n = +1 (5-34a)
and
E J - E 0* = —o>mn_ 1 - ^
- 1 (5-34b)
where /zn = E 0n+ — E 0n is the chemical potential of the n-
1
m m
and using the completeness of the states |^ mn±1) to write this as
— 00
A(p, w) dw = <0 |cpcp + + cp +cp| 0 > = 1
5-5 A N A L Y T IC P R O P ER T IES O F G
C . In our case the contour C cuts the plane into two regions so
that two different functions / : and f u can be defined. As shown
in Figure 5-2, f^Po - p) is analytic in region I and f u ( p - p) 0
5-6 P H Y S IC A L IN T E R P R E T A T IO N O F G(p, p 0)
Gs(r, t ; r, t) = »'<0 |</<s + (r, <)<As(r, <)|0> = *'<0|ps(r, <)|0 > (5-46)
particle in the bare state p. The wave function for the system is
then
cP+ | ^ iO(0 )>
if ¥ 5 o represents the ground state. At a later time r the wave
function has evolved into
e~iHzcp +\ 0(0)>
If the particle were created in state p at t — r rather than t = 0 ,
the wave function would be
cp+e - ‘"' | ^ , o(0 )>
Thus, the probability amplitude for finding the system at t = r > 0
in the same state as it was in at t = 0 is the scalar product of the
two Schrodinger states at time r:
< ^ o ( 0 ) | c ,HV ’ ' % + l ^ o ( 0 ) >
which in the Heisenberg picture becomes
<0| cp ( t )cp + (0)|0> = -t<?(p,T) (r > 0)
On physical grounds we might expect this probability amplitude
to oscillate with some frequency (Ep + /x) and decrease in magni
tude with a decay rate |-Tp|:
<?(p, t) e~ i(Ep +u)re~ |rp|T
oc
As is apparent from (5-22), this form is exact for the free Fermi
gas with E p + /x = ep and r p = 0 .
Let us now investigate the behavior of G(p, r) for large
positive times. By taking the Fourier inverse of the spectral
representation (5-37) one finds for r > 0
/•00
<?(P> T) = - i J A(p, w) e- i(a)+li)z doj (5-48)
While A( p, co) is a sum of delta functions for a finite system, it is
a continuous function in the limit of a large system. We carry
out this integral by pushing the contour into the lower half-plane
as shown in Figure 5-3 and we use the analytic continuation of
A in this region. If the continuation of A is analytic in the lower
118 T h e o ry o f Su p e rco n d u ctiv ity
co-plane
= G (p ,p 0) I m p , < 0 ( ' b)
Since (5(p, p 0) has poles only along the real axis, the poles of A
and G in the upper half-plane coincide. Therefore, the poles of
the analytic continuation of 6r(p, p 0) (for real p < p) into the 0
(5-37b)
122 T h e o ry of Sup ercon d u ctivity
This weight function is more complicated than that for the free
Fermi gas, 8 (o> - [ep - /*]), for three reasons.
1 . The quasi-particle energy E p(±) is not in general equal to
± |cp — fx| but includes “ self-energy” effects arising from inter
actions of the injected particle (or hole) with the medium.
2 . The delta function has been smeared out into a finite-
width Lorentzian function, owing to |0 P> not being an eigenstate
of the interacting system in general.
3. There are two peaks in the weight function (5-52), one for
positive and one for negative o>, both of which have finite weights
up 2 and vp2, respectively. In accordance with our discussion
above, the two peaks reflect the possibility of either adding an
electron to momentum state p (leading to the positive energy
peak) or removing an electron from this state (giving the negative
energy peak). Both processes are possible in the interacting
system for a given state p since the probability that this state is
occupied is neither one nor zero.
Field-T heoretic M e th o d s in the M a n y -B o d y Problem 123
< - 1 (■ + t t ) (5-55a)
^ - 5 (* ” V ) (5' “ b)
and give the probability that the bare particle state p is unoccu
pied or occupied, respectively. For Ap -> 0 , the BCS spectral
function goes over to that of the free Fermi gas. If one goes
beyond the BCS approximation, finite level widths appear (as well
as more complicated “ continuum” contributions, which will be
discussed later). While ^4(p, o>) is in general an extremely com
plicated function, its form for small ep — /z and small oj is expected
to be simply the sum of two Lorentzian functions as in (5-52);
however, the sum of the residues may be less than unity. This
simplicity is exploited in Landau’s theory of the Fermi liquid,
discussed in Chapter 2 , although a correspondingly simple picture
holds for the superconducting state. Since most transport prop
erties emphasize this low-energy part of the spectral function, a
quasi-particle approximation is often sufficient for calculating the
electrical and thermal conductivities, etc.
124 T h e o ry o f Su p e rco n d u ctiv ity
5-8 T H E O N E - P H O N O N G R E E N ’S F U N C T IO N
In a completely analogous manner we introduce the one-
phonon Green’s function D defined by
, t 1; r2, t2) = —^ 0 | T{<pA(ri, tx)(pA+ (r2, G)}|0/ (5-56)
where the T-product is given by
n r, - M r,"
(r 2 J ^)(?)/l(ri’ *2!
IVa ^l) > <!
< ^2
D depends on time only through the difference r = tx —t2, and
fora translationally invariant system depends on the spatial
variables only through r = r — r2. With the aid of the
1
expansion
<pAr, 0 = 2
Q
we see that the propagator of phonons of wave vector q and
polarization A is
A,(q, r) = —i(0| T{<pqA(r)q>qA+ (0)}|0> (5-57)
for a translationally invariant system. We recall that the
phonon-field amplitude is related to the creation and destruction
operators by
*PqA ® - q?.
5-9 P E R T U R B A T IO N SERIES
We shall not go into the details of deriving the perturbation-
series expansions for the electron and phonon Green’s functions
for two reasons: there exist several good discussions of this
derivation in print103; also one need not keep the derivation in
mind when using the simple rules that follow from this derivation.
The perturbation expansion is based on two main assump
tions. If the zero-order Hamiltonian which describes the bare
particles (i.e., our Bloch electrons and bare phonons) is designated
by H q, then the ground state of H is assumed to go over adia-
0
function
G (P) = --------------------------------------- (5-23a)
Po - «p +
0
Term (b) describes the interaction of the added electron (or hole)
with the average charge distribution of the bare conduction elec
trons of both spin orientations. While this term is formally
infinite in the limit of a system with infinite volume, this infinity
is cancelled by an infinite contribution of opposite sign arising
from the one-body potential U as shown in Figure 5-4c (see
Sections 4 - 4 and 4-5 for the definition of U). By thinking through
the definition of U, it becomes clear that the cancellation simply
reflects the over-all electrical neutrality of the electron-ion system.
A similar cancellation occurs whenever a diagonal matrix element
of the Coulomb interaction appears, so that we need not worry
Field-Theoretic M e th o d s in the M a n y -B o d y P roblem 129
(d) (e)
W p) = G j f ) ~ I{P )
(a) (b)
-L
(c)
F IG U R E 5-5 Higher-order corrections to iG(p, p 0). (a) This graph is
automatically included if one includes the contribution shown in
Figure 3-4b in the irreducible self-energy E. (b) This graph is also
included automatically if Figures 3-4d and 3-4e are included in E. (c)
This graph is not automatically included since it has an irreducible self
energy part. Its contribution must be included explicitly in E.
132 T h e o ry o f Su p e rcon d u ctivity
ptPn'5
dpo ________
27T p 0' - f p. + ^ p'
- 1 IP'I < P f
= - u = 0 |P'| > P f
(5-72)
The first two terms on the right-hand side of (5-71) give the direct
and exchange Coulomb contributions; the former is largely
cancelled by the third term (p\U\p}, as we mentioned above.
The last term in (5-71) is most easily evaluated by writing
i /%■ D° ^ p ~ p ')G°{p,)
_ i [ dJ E l ( 1______________ I______ )
J 2n \p0' - Po - QqA + 1 8 Po - Po + Qqx - *8/
x Po
—;------ -------------- (5-73)
- ep. + ir)p.
and noticing that the first factor in D does not contribute for
0
IP'\ > P f since neither this factor nor G (p') has poles in the 0
factor nor GQ( p ‘) has poles in the lower half-plane, and we obtain
zero by closing in the lower half-plane in this case. The remaining
contributions are
r l
J 2 tt p 0' - p 0 - QqA + p0 €p> i8
Po - V +1 |P'|
1
< Pf
(5_74)
_i r ^pp_ _________ i 1
Thus, the last term in (5-71), which arises from the lowest-order
phonon process, is
* -0» - ^pp' a|
1 - fp - , fp- (5-75)
Po - - Q qA + i8 Po - «p. + Qqx - ih
The two terms in this expression can be easily understood
within the framework of time-independent (Brillouin-Wigner)
perturbation theory. Consider adding to the system an electron
in the state p above the Fermi surface. The familiar second-order
process shown in Figure 5-6, in which a phonon of energy Qq A
is virtually emitted and reabsorbed by the added electron, leads
to the energy shift
dzP' IsWa | 2 ( 1 ~
AE == Z^ P
r \
TTTTS
( 2
r £iLlU
po - v -
fy ,
7T
QqX1 (5-76)
q> A
ELEMENTARY
EXCITATIONS IN NORMAL
METALS
6-1 T H E E L E C T R O N G A S W IT H C O U L O M B
IN T E R A C T IO N S
This singularity, because of the long-range nature of the
Coulomb potential, can be circumvented by summing a set of
graphs which physically represents the screening of this potential
by the valence electrons. We neglect the phonons for the moment.
From the pioneering work of Bohm and Pines 105 and from the
more recent work of Gell-Mann and Brueckner , 106 one knows that
the most important screening effects in the limit q -> 0 are given
137
138 T h e o ry o f Su p e rco n d u ctiv ity
»■— - + - 0 - + - - o - o —
+ — c ^ ) — — O ) — +,tc
F IG U R E 6-1 The random phase approximation (RPA ) for the effective
interaction in the electron gas.
/ q0) = 1 +
Re #c0 (q, 2 e 2 m kF i _ + j_ y
24f/
h
Trq 2 q
/mqo _q_\
1 +
x In \qkF 2 k FJ
(6 -6 a)
i _ M * +
\?&F 2k F)
/Lp + M go 3_\
_ M go <7_\
1
2 kFJ
2? \qkF 2 kF) _ im o g\
\qkF 2 kFJ
and
f0 for 2 m|g0| > q2 -j- 2 qkF
0 for q > 2kF and 2m\q0\< q2 — 2qkF
*o(q. 0) = 1 + ^ (6-8)
as q -> 0 , where the screening wave number is given by
fan e2
Ep
2
Re K0 (q>?0) = 1 - ^ (6 - 1 1 )
HQ
The RPA result is given by retaining only the first term in the
series for P(q).
To obtain a compact expression for /c(g), we consider the
quantit}' (evaluated in the Heisenberg picture)
<0 |T { p _ ,(-r)p ,( 0 )}|0 > (6-13a)
Here pq is the gth Fourier component of the electron density
operator,
where
F(q, <o) = V(q) 2 |<«|p_ 9 | 0 ) | 2 S(o> - con0) for «, > 0 (6-14b)
n
and
F(q,o>) = - F ( - q , |w|) for < 0 (6-14c)
For a system with inversion symmetry (6-14c) is equivalent to
F(q,<v) = - F ( q, —co) so that (6-14a) becomes
— — - 1 = f °° F( q, co) (--------------------- r-g) da>
*(q> 0o) Jo Wo - ^ + Wo° 0o + w + Wo<7
(6-15)
The right-hand side of (6-15) is of the same form as the spectral
representation of the phonon (boson) Green’s function. The
essential difference between the two is that the weight function
F(q, co) involves matrix elements of the electronic density fluctua
tion operator, while the longitudinal phonon weight function
0 (q, co) involves matrix elements of the ionic density fluctuation
operator. Thus, in analogy with the phonon Green’s function,
the poles of 1 //c(q, g0) glve boson-like excitations of the
electron gas which are excited by a longitudinal field.
144 T h e o ry o f S u p ercon d u ctivity
(6-16a)
I 7T *(q, q0)
for q o c 0
or
= 1# *2(9)sgn 9o (6 - 16b)
7T Kl (q) + 22
2 K (q)
// r(-q) \
H * A
GQ{p + q)
F IG U R E 6-6 The electron self-energy evaluated within the screened
exchange approximation. This graph presumably gives the most
important contribution to E(p) from small momentum transfer pro
cesses (|q| « kF).
Elem entary Excitations in N o rm a l M etals 147
k — q,
6-2 T H E C O U P L E D E L E C T R O N - P H O N O N SY STEM
Screening also plays an important role in determining the
dressed phonon frequencies and the dressed electron-phonon
interaction, since the long-range Coulomb forces between ions
and between ions and electrons are screened out by the conduction
electrons. For long wavelength effects we can again use the
random phase approximation since it gives the leading corrections
in this limit. For shorter wavelengths, processes neglected by the
RPA begin to play a role. However, the RPA gives roughly
correct answers for all wavelengths, since in the region where it is
least accurate it gives small corrections to the bare quantities so
that errors in these small corrections can often be neglected.
We begin by treating the entire electron-phonon system
within the RPA and, for simplicity, neglect umklapp processes
in both the Coulomb interaction and the electron-phonon inter
action. The (irreducible) longitudinal phonon self-energy n x(q)
is given by the series of graphs shown in Figure 6 -8 . The string
of bubbles, characteristic of the RPA, simply represents the
screening of the ion-ion potential. From the series for r c(q)
shown in Figure 6 - 1 we see that i^ c(q) and TJ^q) differ in two
respects: (a) the incoming and outgoing bare Coulomb lines in
y ' c[q) have been replaced by electron-phonon matrix elements in
n x{q) and (b) the leading term V(q) in the series for y c{q) is
missing in n x(q). Therefore, we have the relation
where we have used (6-3) in the left equality and have assumed
gv p. A is a function only of the momentum transfer for longi
tudinal phonons. By rearranging (6 -2 1 ) we have
n,(q) =
ball2 1
1 (6 - 2 lb)
_«(?)
-
V(q)
or
(6-25b)
From this expression we see that the dressed longitudinal phonons
have a sound-wave type of dispersion law with sound velocity
(mZl3M)ll vF, where vF is the Fermi velocity. This result for
2
i,,= 4 . 5 0 ^ - " + . 0 — O — * + e te
F IG U R E 6-9 The random phase approximation for the screened inter
action between electrons and longitudinal phonons.
where {gQtJ 2 = gqig _ qi. The total self-energy within this approxi
mation is
Z(p) = i j o 0(p + q ) \ r c{q) + { g ^ D ^ - q ) ] (6-28)
£>i(-q)
= * J^ o r(?)___________ (6-29b)
1 + 7(g)P(g) - ^ + iS (277)4
vo
The denominator in (6-29b) is just the total dynamical dielectric
constant of the system including electronic and ionic polariza-
bilities, since the ionic polarizability is given by the high-frequency
= JQ (6-32)
m CP o
dq 0 q dq{gqlf D,(q)
where q (p 0) is the wave number such that ojql = \p0\. For the
jellium model with K(q) given by the static long wavelength limit
k /q [see (6 - 8 )], we find the damping rate of electrons, because of
2 2
the jellium model) for all p by using (6-34) for |_p0| less than the
0
(6-35)
where {a, b} means the smaller quantity, a or 6 . Since Qv ~
1 0~ E F phonon processes dominate the damping rate up to
3
For jellium this leads to the effective mass correction near the
Fermi surface,
(6-37)
From (6-37) we see that the phonon cloud surrounding the
electron increases its effective mass. The numerical value of
Smph given by this expression is not to be trusted in real metals
[nor is the damping rate (6-34) to be trusted] since umklapp pro
cesses have been treated as if they were normal processes in our
jellium model and transverse phonons have been neglected al
together. In addition, the bare electron-phonon matrix elements
in real metals no doubt differ widely from the jellium results for
large momentum transfer, as discussed in Chapter 4. Neverthe
less, these expressions suggest that damping effects and mass
corrections due to phonon processes are quite important, a result
supported by experiment.
We turn now to an important observation made by Migdal. 15
Consider for the moment the class of graphs for E, in which no
Coulomb interactions appear, other than those implicitly included
Elem entary Excitations in N o r m a l M etals 157
P + ?
functionr(p,
F IG U R E 6-13 The perturbation series for the electron-phonon vertex
for clarity.
q). The external electron and phonon lines are included
X
(a)
(b)
F IG U R E 6-14 The one-electron spectral weight function ^4(p, w) for
electrons Interacting with dressed phonons of a constant frequency cj0.
The Bloch energy €p is measured relative to the Fermi surface, and
A is expressed in the form A(€pt xw0). The coupling constant was
chosen such that g2N(0)lwo = The plots are given for (a) ep = 0,
(b) €p = OJS wq, (c) €p = 2w0, and (d) cp = 5ca0. The vertical heavy line
represents a delta function in each plot. (See p. 163.)
Elem entary Excitations in N o r m a l M etals 159
(d )
FIGURE 6-14 (continued )
x
(a)
».75a>£>* x w d )
(b)
F IG U R E 6-15 The one-electron spectral weight function ^4(p, w) for
electrons interacting with dressed phonons having a D eb/e spectrum
(oq oc q with a maximum frequency toD. The Bloch energy ep is measured
relative to the Fermi surface, (a) Spectral weight function
for €p = 0. There is a delta-function contribution at x = 0. (b)
Spectral weight function for cp = 0.75o>D. (c) Spectral weight function
for €p = 2o)D. (d) Spectral weight function for ep = 5wD. (see p. 163.)
Elementary Excitations in Normal Metals 161
(c)
(d)
FIELD-THEORETIC
METHODS APPLIED TO
SUPERCOND UCTIVITY
—j— e tc .
+ + + + -*■
- ( - e tc .
\
+ + Q e tc .
/
F IG U R E 7-3 A class of ladder graphs which contributes to the electron-
phonon vertex function T and leads to the superconductor instability.
< £(7 o )
appear. For A0 < 0 and very small, again no bound states appear.
However as A0 becomes a large negative number (i.e., stronger
attractive interaction) the 1 /A0 line slides up toward the g0-axis
until it reaches the critical value such that it is tangent to the
peak in 0 ( q Q) near q0 = 0 . Thus far, no bound states have
appeared. Beyond this point, however, two pure imaginary roots
appear corresponding to the two roots which disappeared for A0
greater than the critical value. These pure imaginary roots 10 1
can be obtained by converting the sums in (7-6) to integrals again
and assuming the density of Bloch states in energy near the Fermi
surface is a constant N ( 0 ). Then if we set q0 = i a , we have
(7-7)
For iV(0)A0 < 0 (and larger than the critical value mentioned
above, which is vanishingly small in the limit of a large system)
one has the pair of pure imaginary roots
a =
where the last equality is valid for weak coupling, that is,
|iV(0 )A0| < 14 ’
There are several important differences between the results
of Cooper’s two-particle problem and the ^-matrix problem dis-
cussed above. The solutions of the two-particle problem gave
a real energy corresponding to a stable bound state. When inter
actions are included within the Fermi sea, as in t, there is no longer
a stable bound state formed from the normal phase, but there is a
rapid unstable growth in amplitude of correlated pair states.
Unfortunately, one cannot trace the time evolution of the un
stable system long enough to discover directly what state it is
tending toward, since the ^-approximation is not accurate enough
to treat the correlations in the superconducting phase.
Another difference is that the binding energy of the one-pair
Field-Theoretic M e th o d s A p p lie d to S u p ercon d u ctivity 169
= # 0 + # in t (7-9)
In the Hartree-Fock approximation one linearizes the inter
action term with respect to a given state | 0 > so that the two-
body operator H lnt is replaced by a one-body operator. In the
linearization one replaces a typical operator product cx+c2 +c3c4 by
c1 +c2 +c3c4 => <(0 |c1 +c4 | 0 )>c2 +c3 — +c3 | 0 )>c2 +c4
+ (Olca +CalO^ +c* - <0|c2 +c4 |0>c1 +c3 (7-10)
170 T h e o ry o f Sup ercon d u ctivity
determining yp.
While this appears to be a complicated way of phrasing the
H F approximation, the scheme is easily extended to include the
pairing correlations. The idea is simply to generalize the lineari
zation (7-10) to include Hartree-like terms involving (0\c1+c2 + 10>
and <0 |c3c4 | 0 >. The state | 0 > is then to be determined self-
consistently as in the standard Hartree-Fock approach.
At this point one might argue that since the full Hamiltonian
commutes with the total number of particles operator Aop, the
exact eigenstates of H are eigenfunctions of N op. One might then
argue that our approximate state | 0 > should also be chosen to be
an eigenfunction of N op and therefore the Hartree-like terms
<0 |c1 +c2 +| 0 ) and <0 |c3c4 | 0 > vanish identically. We can make
two counter arguments. First, in the limit of a large system the
ground states of the N 0 and the N 0 + v particle system are
degenerate for |v| « N 0, if the origin of energy is shifted by
Field-T heoretic M e th o d s A p p lie d to Su p e rco n d u ctiv ity 173
Ho = 2k Pk +[*kT3
X + 4>kiTi + j>k2 T2 W k
k
+ 2 *k(7-33)
where <f>kl and <f>k2 are the real and imaginary parts, respectively,
of <f>k and €k = €k + Xk ~~ P as before. The last term in (7-33)
is included to compensate for the —1 occurring in (7-32c).We
can easily remove this annoying (infinite) term once we go over
to the Green’s function scheme. The form (7-33) of H 0' is just
that required to make the Feynman-Dyson perturbation series
rules work if H int takes the form of a one- and/or two-body
Field-Theoretic M e th o d s A p p lie d to S u p ercon d u ctivity 175
~^p 0
G022(P) = Po2 „(Po2 -- E «P)e
p2+ iS
(7-43b)
<t>p*
= Po2
V -- VUp E 2 + ii88 - ,7 -4 3 o )
(7-50b)
(7-50c)
where the pairing matrix Vpp>is given by
VPV = <p'. - P 'I F Ip> - P> = <p'. p |r|p , p'> (7-51)
[see (7-34b)]. Since the total Hamiltonian is invariant to rota
tions in T-space about the r3-axis (rx and r 2 never enter H ), we
are free to choose phases so that </>2 = 0 and (7-50b) reduces to the
BCS energy-gap equation, where <f>p is identified as the energy-gap
parameter Ap. It is interesting to note that the Hartree-Fock
potential yP is °f the expected form since vp>2 gives the average
occupancy of states p', s. Thus, this generalized HF scheme is
equivalent to the BCS theory discussed in Chapter 2 if yP is
lumped with the energy ep of that chapter.
To make connection with the spectral weight function ^4(p, co)
of Chapter 5 , we note that G011(p) can be written
where
Field-Theoretic M e th o d s A p p lie d to Su p e rc o n d u c tiv ity 179
= V ( , - E p) + v 8 (0 , + E p)
8 0 p 2 (7-54)
The physical interpretation of the two peaks in A(p, co) was given
in connection with (5-53) and will not be repeated here.
Another way of viewing the generalized HF scheme, which is
easily extended to include retardation and damping effects, etc.,
is that of self-consistent perturbation theory. In this approach
one calculates Z(^p) as a perturbation series in which one uses one-
particle propagators which themselves include the self-energy
being calculated. Therefore, one obtains an approximate integral
equation determining Z. In carrying out the procedure one must
be careful not to double count graphs. If we continue to use the
Nambu notation, the most general form for 'L(p) is
H(p) = [1 - Z (p )]p 01 + X(P)rs + HV)?! + f c p ) r (7-55) 2
(7-56)
(7-57)
180 T h e o ry o f S u p ercon d u ctivity
7-3 Z E R O -T E M P E R A T U R E E X C IT A T IO N S P E C T R U M
We are now in the position of being able to handle damping
and retardation effects in determining the quasi-particle spectrum
of a superconductor at zero temperature. N am bu114 and Eliash-
berg122 have treated the problem by self-consistent perturbation
theory and have retained the lowest-order dressed-phonon and
dressed-Coulomb contributions to Z, as shown in Figure 7-5.
Within this approximation one finds the (matrix) equation
E(p) = i JT 3G(fOT3|^2 f a p - P') + ^ c ( P - p')
(7-61)
determining 27, where G is given by (7-58). From the above
equation one can determine the unknown complex functions
Field-Theoretic M e th o d s A p p lie d to Sup e rcon d u ctiv ity 181
s,h()>» - ( w w r . iv ° r . *■"
[Z(p0')p0'l -
[Z(Po')Po'? ~ V2- 4>2(Po') + i 8
(*2k
X2 X JO
F 1 d q iV q ifD M ’ Po - Po') (7-64)
where we have used the same approximations as in (6-31) and for
simplicity have assumed particle-hole symmetry to be valid near
the Fermi surface in order to make the r 3-term in Z ph vanish.
We use the integration procedure of Eliashberg122 and break up
D (p - p ) into two parts, Du which is analytic in the upper
half of the g?0'-plane and D l analytic in the lower half-plane [see
the spectral representation of D (5-59)]. For the term Du, the
portion of the ^ 0'-contour originally along the positive real axis
Field-Theoretic M e th o d s A p p lie d to Sup ercon d u ctivity 183
co m p le x to — p lan e
—Aq Aq
F IG U R E 7-6 Folded contour for evaluating 27uph.
184 T h e o ry o f Sup ercon d uctivity
■ M M iiiiiw £ w m m m m^m--------------------
m m m mm,
—
F IG U R E 7-7 Folded contour for evaluating Z'/ph.
By sending p 0' -> —Vo in (7-67) and using the fact that Z ( p )
and </>(p) are even functions of p 0, a fact which follows from (7-61),
one finds for |p | ~ p F,
& * ( p ) = X fH p ) + Xuph(p)
= m f dp.' Re {[ ( Z P .)
(7-69)
The interaction kernels K +vh and A _ph are defined by
where /z is the cosine of the angle between p and p'. The pseudo
potential equation (7-75) now becomes
0x >
(7-78b)
0x <
The physical interpretation of the pseudo-potential is clear
from the form of this equation since it is just the equation satisfied
by the 5 -wave part of the ^-matrix for two-particles scattering in
the region outside of —coc —> cuc. It is reasonable that the
effective potential to use between particles near the Fermi surface
is given by the sum of all multiple scatterings of the particles in
the region away from the Fermi surface, plus the Born term for
scattering near the Fermi surface, in agreement with (7-78a).
The pseudo-potential can be obtained explicitly if Vc(p, p ') is
approximated by a factorizable potential:
otherwise (7-79)
where the maximum energy a>m is of order the Fermi energy E F.
While this is a rough approximation, it should give the correct
order of magnitude for C7c, and one finds
U c(p ,p ') = (7-80)
1+ N (0 )V c In
if the density of Bloch states is taken constant for \ep\ < ajm.
This result, first given by Bogoliubov, Tolmachev, and Shirkov,52
shows that the effective Coulomb repulsion to be used near the
Fermi surface is weaker than the true screened Coulomb inter-
action due to the factor [1 + N ( 0) Vc In (wm/wc)], which is typically
of order 2 or 3. Physically, this reduction is associated with
the fact that scatterings far from the Fermi surface lead to a
smaller probability for two electrons being within the range of the
188 T h e o ry o f S u p ercon d u ctivity
which determine A and Z where A0, the value of the gap parameter
at the edge of the gap, is defined by
A o = A(A0) (7-84c)
Numerous approximate calculations have been made in an attem pt
to solve these equations. The first numerical calculations were
performed by Swihart,125 who set Z = 1 and in effect took K +
to be a square-well potential in the variable p 0 — p 0', attempting
to approximate the Bardeen-Pines potential. From his results he
concluded that A (p 0) decreases monotonically and changes sign
as p 0 increases from zero. Morel and Anderson124 and Culler
et al.126 used the correct Eliashberg potential K appearing in
(7-84) but took Z = 1, as before. The former group treated
analytically the case of constant-frequency phonons (an Einstein
spectrum), while the latter group treated the Debye spectrum
(wQoc q). In both cases it was found that A increased initially
with increasing p 0 before changing sign near the average phonon
frequency, the peak being due to the resonant nature of the
Eliashberg interaction.
The most complete calculations at present are those of
Scalapino, Wilkins, and the author.78 In attempting to explain
the tunnel-current anomalies observed in superconducting lead
by Rowell, Anderson, and Thomas,88 and earlier by Giaever,
Hart, and Megerle,87 they assumed that the weight function
determining K ±ph (7-61) is of the form
F IG U R E 7-9 (a) The same as Figure 7-8 except the Coulom b pseudo
potential is set equal to a value roughly appropriate for lead, N(0)Uc =
0.11, (b) The real and imaginary parts of the renormalization function
Z(w) for this case, (c) Tunneling density of states.
Field-Theoretic M e th o d s A p p lie d to Su p ercon d u ctivity 193
J -» p 0 - O ) + l S J - „ ^0 + a>-t8
where the spectral functions are given by
P<+>(P, co) = 2 Mn|<m |cp+ |w>|2 - E n - W) (7-95a)
n, m
It follows from (7-95) and (7-96) that p(+) and p<_) are related by
p<->(p,co) = e « V +)(P. - « ) (7‘97)
Field-T heoretic M e th o d s A p p lie d to Su p e rco n d u ctiv ity 195
/•oo /•00
A(p, t o ) d w =[/>< + ) ( p , o i ) + / ( _ ) ( p , a> )] ( i ai
3
J—
00 J— 00
= 2 w„{<n|cpcp+|n> + <n|cp+cp|w»
n, m
= Tr u = 1 (7-99b)
where we have used the anticommutation relations for the c’s.
As pointed out in Chapter 5, the functions A, p( +), and p(~ )
are related to the functions A BK, G >, and G <, discussed by
Kadanoff and Baym,91c by
A bk{]>, a>) = 2ttA( p , w) (7-100a)
G > (p , to) = 2 n p ( + )( p , w ) (7-100b)
( ^ ( p , to) = 277/)( - ) ( p , - t o ) (7-100c)
196 T h e o ry o f Sup e rcon d u ctiv ity
that is, an even multiple of 7n//? (while the electronic energies are
replaced by odd multiples of 7ri/j3). The difference between the
boson and fermion rules comes from the definition of the T-product.
In analogy with (7-102), g0-integrals are replaced by
r. i H . l « » •*>
The zero-order propagators to be used in the perturbation series
Field-Theoretic M e th o d s A p p lie d to S u p ercon d u ctivity 197
for the sum (7-111). In order that ^(p, z) be bounded for large
z we must set
f(icon ± Qq)= eiPa}ne±pn(i + j = i _ e ± w q (7“113)
before replacing ia>n by 2 . Therefore, the properly continued
self-energy is given by
4>* (7-116)
„ ..4 „ (*«v)2 - V 2 - K 2
The Poisson summation formula (7-113) allows us to convert the
w'-sum to the form
Field-Theoretic M e th o d s A p p lie d to S u p ercon d u ctivity 201
where Vpp> = <p\ p| F|p, p'>, in agreement with the result of BCS
and of the linearized equation of motion treatment (2-76). The
finite-temperature generalization of the retarded interaction
problem discussed in Section 7-3 is also straightforward to carry
out by these techniques.78
Although we have emphasized the calculation of G at finite
temperature, the phonon Green’s function follows in a similar
manner.
As a general conclusion of this chapter, one can say that, on
refining the BCS theory to include noninstantaneous interactions
and damping effects, only details of the quasi-particle spectrum
are altered. In particular, the energy gap in the elementary
(*}
E v. E v.
ELECTROMAGNETIC
PROPERTIES OF
SUPERCONDUCTORS
8-2 W E A K -F IE L D R E S P O N S E
In the beginning of Chapter 2 we argued that transverse
electromagnetic fields need not be included directly in calculating
the detailed pairing interactions which bring about superconduc
tivity. Their effect can be taken into account in terms of a
space- and time-dependent average field which is calculated self-
consistentlv from the external field and the currents flowing in the
material. While the externally applied magnetic field generally
represents a large perturbation on the system, the induced field
arising from the supercurrents cancels the external field over most
of the material, as we know from the Meissner effect. Therefore,
the net field acts only very near the surface and can often be
treated as a weak perturbation on the system as a whole. Thus,
we shall formally treat the total transverse electromagnetic field
as an externally applied field and solve for self-consistency as a
separate problem.
As we have seen above, the Coulomb potential plays an es
sential role in the pairing theory. It cannot be treated by the
self-consistent field scheme we use for the transverse field and
therefore we include the total Coulomb interaction in the zero-
order Hamiltonian.
We begin by considering a simply connected bulk super
conductor of unit volume in the presence of a weak externally
applied electromagnetic field described by the vector and scalar
Electrom agnetic P roperties of Su p e rco n d u cto rs 207
jup(x) (p = i = 1, 2, 3) (8-5)
Pe(x) = - e ^ <Ps+(x )>Ps(x ) = ~ep{x) {p. = 0)
j uix ) = j A x ) + j A x ) (8-6)
where the diamagnetic current density j d is given by
• d(~\
J u ( x ) = <mc
Pe(x )^i(x ) (n = i = 1, 2, 3)
(8-7)
I 0 (/x = 0)
The full coupling of the electrons to the perturbing electromagnetic
field is then
H' = H> + H d
where the diamagnetic coupling is defined by
nd = -
J
f p»{x) i=i
2 A >2(x ) d3r (8-8)
208 T h e o ry of Su p e rcon d u ctivity
TYIC
(#* = * = 1, 2, 3) (8-16)
This relation is clearly not gauge-invariant since the predicted
current depends upon the choice of gauge. In London’s equation,
only the transverse part of A is to be used1 and therefore J is
properly gauge-invariant. Since the longitudinal part of A
couples to longitudinal excitations, the wave function is not
“ rigid” with respect to this type of perturbation and the para
magnetic term does not vanish in this case. In fact, if A is purely
a gauge potential, the paramagnetic and diamagnetic terms
exactly cancel as required by gauge invariance. In carrying out
210 T h e o ry o f Su p e rcon d u ctivity
8-3 T H E M E IS S N E R -O C H S E N F E L D EFFECT
As Schafroth has shown,14 the Meissner effect requires that the
transverse part of the kernel K uv remain finite in the long wave
length limit (q —^ 0) for zero frequency (q0 = 0). Now gauge
invariance and charge conservation require that
3
2 Kitfj = 0 (gauge invariance) (8-28a)
j= i
and
^ Qi^ij = 0 (charge conservation) (8-28b)
i
X < 0 |T { ¥ V (r)in + 1
< ( T ) ^ k. +/ ( 0 ) iq 'k.(0)}|0> (8-36)
214 T h e o ry of Su p e rcon d u ctivity
PM = “5 J ( 0 ( k + l)t(k + + tjTr^k W
where q = (q, q0). If the pairing potential is nonretarded, we
saw in Chapter 7 that within the pairing approximation G(A;) is
given by
G(*) = * *0
°,V ~ + Io = (^ 2 + V ) 1'2] (8-39)
[see (7-41)]. Since we are interested in the static Meissner effect,
we set ^o = 0 and P XJ reduces to
r
J ^ (k , q) = ]--------------- (8-46)
\€k+q €kI
if the k and k + q are on opposite sides of the Fermi surface, and
zero otherwise, as required by the Pauli principle. As we argued
above, a finite value of P lj(q, 0) arises in the normal metal as
q2 —>■0 despite the fact that most of the matrix elements vanish,
because of the vanishingly small-energy denominator in this
case. If one calculates the magnitude of P tJ one finds that it
almost exactly cancels the diamagnetic term, leaving the weak
Landau diamagnetism of the normal state.
In “ gapless” superconductors172 both the matrix elements
and energy denominators vanish, but the density of states near
the Fermi surface is small enough to ensure that P XJ does not
cancel the diamagnetic term as q -> 0.
To extend this calculation to finite temperature, we use the
prescription discussed in Chapter 7 to convert the zero tempera
ture form (8-38) to one involving the discrete frequency sums.
The only change is that L ( k, q) becomes
+ ( 8 ' 4 9 )
1 ( M l,
( m) ) ( (8_53)
Ps(0) =n = ^ (8-54b)
(8-62)
to that in the normal state and express the ratio in terms of the
complex surface conductivities cr in the two states. Thus,
ct, = a 1 + ia 2 = 1(oj, 0, T ) _g g
Gn crn inhaj
The expression for G1/crn was given in Chapter 3 while the expres
sion for o 2 is
cr 1 [ '
2 [1 - 2f ( E + hoj)][E2 + hwE + A2] dE
^ ^ L-HO.-A {[^2 - E 2I(E + M 2 - ^2]}1/2
(8-70)
the lower limit being the larger of the two quantities A — %a>
and —A. At zero temperature, the ratio cr2/crn is
<7n
= ( l + p \ E (k) ~ ^ - K ( k )
\ f id ) } fid)
(8-72)
In these expressions E and K are the complete elliptic integrals
and
Jc' = (1 — k 2)112 where k = 22AA — heo
+ hd) (8-73)
The functions o x and cr2 have been calculated by Tinkham 129 for
T = 0. For T ^ 0 numerical calculations are necessary to
determine the surface conductivity; however, a simple low-
frequency limit is
CTo 7tA . A
^ = az t a n h < 8 ' 7 4 )
8-5 G A U G E IN V A R IA N C E
While the simple pairing approximation gives an accurate
account of the response of the system to transverse electromagnetic
fields, it does not in general give the correct response to longi
tudinal fields. In particular, we saw in a previous section that
it predicts unphysical longitudinal currents which depend on the
choice of gauge of the electromagnetic potentials. The physical
origin of this difficulty was first recognized by Bardeen,131 who
pointed out that a (longitudinal) gauge potential couples primarily
to the collective density fluctuation mode of the electron system
(i.e., the plasmons of the charged electron gas). He argued that
if one generalizes the pairing scheme to include this mode in a
consistent way, a gauge-invariant theory would be obtained.
While a number of authors have contributed to the detailed
resolution of this problem, the pioneering work of Anderson47
followed by that of Rickayzen132 gave the essentials of a general
ized pairing scheme which includes these effects. In essence,
their approach is to extend the random phase approximation to
include pairing correlations.
It is well known that a gauge-invariant response is a conse
quence of local charge conservation in the system. By local charge
conservation we mean that the electronic current and charge
density operators satisfy the continuity equation at each point in
space and time,
V -j(M ) + = 0 (8-75)
In Fourier transform variables this becomes
q • j(q. %) - goMq. 9o) = 0 (8-76a)
Electrom agn etic P roperties o f Su p e rcon d u cto rs 225
With the definitions (8-5) and (8-6) plus the metric (1, 1, 1, —1)
used previously, the continuity equation for the four-current
becomes
2 = 0 (8-76b)
u= o
where q = (q, q0) as usual. From these relations, it follows that
the expected current
•J/iM ) = <j*(r»0> (8‘77)
satisfies the continuity equation
V • J(r, t) + at/° ^ ’ ° = 0 (8-78a)
or in Fourier transform space,
2 qJM ) = 0 (8-78b)
U=0
If we concentrate on the linear response of the system to the
potential A u, we define [see (8-25)]
= ~ h 2 K u M A v(q)
77 v= 0
(8-79)
It follows from the continuity equation (8-78b) that the response
kernel K uv(q) must satisfy the equation
3
uI =o
[For qQ = 0, this condition reduces to the condition (8-28b),
used in discussing the static Meissner effect.]
Turning now to the restrictions imposed on K by gauge
invariance, we note that the most general gauge transformation
is of the form
A(r, t) => A(r, t) + Wl(r, t)
226 T h e o ry o f Su p ercon d u ctivity
v= 0
2 K uM qv =0 (8-82)
To show the equivalence between the restrictions of local
charge conservation (8-80) and of gauge invariance (8-82), we
note that K uv(q) satisfies the symmetry relations
Re K uv(q) = Re K vu( - q ) (8-83a)
Im K uv(q) = - I m K vu( - q ) (8-83b)
as a consequence of the definition (8-26) and the spectral repre
sentation (8-18) for the retarded commutator E uv. Therefore by
changing dummy indices, the real part of the charge-conservation
restriction (8-80) can be written as
2 qv Re K vu(q) = 2 Re K uA = 0 (8-84a)
V V
or sending qu —qu
2 Re K liv(q)qv= 0 (8-84b)
V
or
2 Im K uv( q ) q v = 0 (8-84d)
Electrom agn etic Pro pe rties o f Su p e rco n d u cto rs 227
i t +1 - ^ { [ i W +%rrH),f" (!')}|0>
+ <0|T{[jo(2), ^ ) ] ^ +(y)}|0> S(z0 - x0)
+ < 0|T {ne)[jo(*), ^ +(2/)]}|0> 8(z0 - Vo) (»-90)
The last two terms on the right-hand side arise from differentiating
the time dependence due to the time-ordering symbol T. Now
the first term on the right-hand side of this expression vanishes
by virtue of the continuity equation (8-75). If we use the equal
time anticommutation relations of the W s (7-21), the commutators
in (8-90) can be reduced to
[jo(2)> ¥V )] s(2o - x o) = eT3W(z) 84(z - x) (8-91a)
and
[j0(z), Xf/+ (y)} S(20 - Vo) = - e f /+ (y)x3 s4(z - y)
(8-9 lb)
230 T h e o ry o f Su p e rcon d u ctivity
- I
G(a: — x') 3 s r, +, dT0
dz,o 2
G(y' - y) d*x' d*y' (8-92)
p +q r v(p,p + q)
y»(p + q>p)
P uv (q ) = — i e 2 x -----
* H <S?
= ie2 j*Tr [y„( 2>,2> + q)G(p + q ^ G ' ^ p + q)r 3
to
2 p uv(q)qv = i<? |Tr[{Y w(l> + q ,p ) ~ y u( p , P ~ ?)}
U J
x r 3G ( y )(Z7
] |^T) (8-101)
Electrom agn etic Pro pe rties o f Su p e rco n d u cto rs 233
[
4,jtw62 1 1
mc?~ + W " [1 " **-°] = ° ( 8 ' 1 0 6 )
8-6 T H E V E R T E X F U N C T IO N A N D C O L L E C T IV E
M ODES
In seeking a gauge-invariant generalization of the pairing
scheme, Nambu used a well-known prescription of quantum-
field theory for constructing approximations which satisfy the
generalized W ard’s identity (GWI) (8-89).137 If G is described
234 T h e o ry of Su p e rcon d u ctivity
f / S
U______L- -(- -JL .u____ L—i__ |___ Li----Li----U- + -J-1-L—L___\ 1 .1
. \ / \ "“ v \
p + q
+ r.
s
F IG U R E 8-5 An equation for the vertex function Tu which leads to
manifestly gauge-invariant results for the electromagnetic kernel
within the pairing scheme.
Ele ctrom agn etic Pro pe rties o f Su p e rco n d u cto rs 235
where the G’s are the Nambu functions evaluated within the
pairing approximation (8-108). To check that the solution of
this vertex equation is in fact consistent with the GWI, we form
the quantity
2 + 9>P) = 2 A p + 9 ,p )
+ i f r 3G{k + q) 2 quTu(k + q, k)
Jj *■rH " "'
x—
(2tt)«* (8-110)
d'k
G(k)r3-T(p - k)
= T ^ ~ 1(P) - G - ^ p + q)r3
3 (8-114)
which is the required GWI. Therefore, if G and Tu are given by
solutions of Eqs. (8-107), (8-108), and (8-109), the electromagnetic
kernel K uv determined through (8-99) will be manifestly gauge-
invariant.
To understand the mechanism by which gauge invariance has
been restored in this rather formal scheme, we again look at the
236 T h e o ry o f S u p ercon d u ctivity
By inserting the relation (8-118) into (8-119) and solving for CQi
one finds the explicit solution
r u(p + q,p) = Yu{p + q,p) + [1 - A<J>(g)]~1x w(?)w*(p) (8- 120)
where <|>(g) is a generalization of the function &(q) [see (5-21)] used
in discussing the instability of the normal state. The matrix
function <j>(g) is given by
4(9) = i J t 3‘Gl(k + 9)T3'G'(*)|W(k)|a (8-121)
where p and P are the mass density and pressure of the free-
electron gas. Therefore within this approximation pairing corre
lations play no direct role in determining the velocity of this mode,
their main function being to remove low-lying single-particle
states which would otherwise lead to damping of the wave.
Returning to the solution (8-120) for the vertex function, if
the two-body potential is purely 5-wave, the transverse part of
%u vanishes, as one can easily see on symmetry grounds from
(8-122). Therefore vertex corrections do not affect the Meissner
kernel in this case. If there is a finite attractive d-wave part of the
potential, d-state excitons exist and will contribute to the vertex
function. Calculations by Rickayzen show that these collective
corrections to the Meissner kernel are in general small.
If there is a strongly attractive d -wave potential one should
see a precursor for infrared absorption below the gap edge, owing
to creation of d-state excitons.138,139 Such anomalies were first
observed by Ginsberg, Richards, and Tinkham,71, 72 although as
Tsuneto138 has shown, the predicted absorption is an order of
P + q P + q
p + q
P P p k P
To make connection with the RPA for the electron gas, we note
that P RPA (4-2) is proportional to P 00 in (8-99) when we (1)
include only the polarization term (8-127) in the equation for
(2) replace all G’s by Gq and (3) set gq
s , = It is easily seen
0.
8-7 F L U X Q U A N T IZ A T IO N
A qualitatively new effect arises when we investigate the
electromagnetic behavior of a multiply connected superconducting
system, e.g., a long, hollow cylinder. In this case, flux can be
trapped in the hole and persist in the absence of an externally
applied field. On the basis of London’s “ rigidity” concept he
concluded that for a cylinder with walls thick compared to the
penetration depth A, flux could be trapped only in multiples of
hc/e = 4 x 10 ~7 gauss cm2.1 This value of the flux quantum
follows if one assumes that the only low-lying current-carrying
states of the superfluid are those given by multiplying the super
fluid ground state by a single-valued phase factor, as we saw in
Chapter 1. We shall see below that there are two distinct sets
of low-lying states, one being the set considered by London, the
other arising from phase factors multiplying a basic state which
is not included in London’s set. Owing to these two sets of
states, the flux quantum in superconductors is actually hcj2e,
i.e., one-half the London unit. The even multiples are associated
with London-type states, while the odd multiples are due to the
Electrom agnetic P ro pe rties o f Su p e rco n d u cto rs 241
8-8 T H E K N IG H T SHIFT
On the basis of our earlier discussion it would appear that
the electronic spin susceptibility should vanish in a superconductor
as T —> 0 if one uses s-state pairing. In this case the Pauli
principle ensures that the spins are paired in the singlet state
so that a finite spin magnetization can result only if the spin
Zeeman energy 2fxBH is greater than 2 J 0, the minimum energy to
break up a pair. A means of checking this prediction is the
Knight shift142 (the change in nuclear magnetic resonance fre
quency due to the coupling of the nuclear spins with the polariza
tion of the electrons). If only the electronic spin polarization
(as opposed to orbital effects) is important in the shift, the Knight
shift gives a measure of the electronic spin susceptibility. Reif143
found that the shift in superconducting mercury, when extrap
olated to 0°K was about two-thirds of the value in the normal
state, contrary to the simple pairing theory. Androes and
Knight144 found similar results in tin, while the shift in vanadium
is nearly the same in the N - and S-phases.
Ele ctrom agn e tic Pro pe rties o f Su p e rc o n d u c to rs 245
Collective Magnetization
In the original BCS paper8 it was suggested that there might
be low-lying collective spin wave states in a superconductor
which give rise to the observed Knight shift. Bardasis and
Schrieffer139 found that by retaining the p-wave part of the two-
body potential, spin wave states can exist in the energy gap.
However, to obtain a nonzero long wavelength susceptibility the
spectrum must go down to zero energy in this limit. These authors
found that two situations can exist. If the p-wave part of the
two-body potential is weaker than the 5-wave part, the spin
waves possess a finite energy as their momentum goes to zero.
If the jo-wave potential is stronger than the 5-wave potential, the
spin wave states are unstable and the ground state is then formed
by p-state pairing. One is then led back to the difficulties of the
first proposal.
8-9 T H E G IN S B U R G - L A N D A U - G O R ’K O V T H E O R Y
Thus far we have concentrated on the response of a super
conductor to weak electromagnetic fields. There are many im
portant problems, e.g., N-S phase boundary, the intermediate
and mixed states, etc., in which the magnetic field enters in a
nonperturbative manner. These problems typically involve the
energy-gap parameter A varying as a function of position in the
materials. As we discussed in Chapter 1, the phenomenological
theory of Ginsburg and Landau36 (proposed in 1950) gives in
many instances a good account of these strong field situations.
An important advance in the microscopic theory was made by
Gor’kov,37 who showed how the GL equations follow from the
pairing theory when T is near T c and the magnetic field varies
slowly in space over a coherence length. Gor’kov found that the
GL effective wave function ^(r) is proportional to the local
value of the gap parameter A(r) and the effective charge e* of the
GL theory is equal to 2e, the charge of a pair of electrons. It is
interesting to note that these results were guessed prior to
Gor’kov’s work, the identification of ^(r) andd(r) being suggested
by Bardeen8 and the effective charge e* = 2e being suggested
by Ginsburg153 prior to the BCS theory in order to fit the GL
theory with experiment.
We give a brief summary of Gor’kov’s derivation below.
To familiarize the reader with Gor’kov’s scheme, we use his
notation. For simplicity Gor’kov used a nonretarded zero-range
attractive potential to describe the pairing interactions. Since
this singular potential leads to divergences, it is cut off in momen
tum space at the appropriate point in the derivation. The vector
Electrom agn etic Prope rties o f Su p e rco n d u cto rs 249
(8-144)
As we discussed in Chapter 7, the pure imaginary time
Green’s functions can be expressed in the Fourier series variable
ajn = (2n -f l)nlp (n = integer) and one finds the Fourier com
ponents ^ w(r, r') and S' J r , r') satisfy
and
252 T h e o ry o f Su p e rco n d u ctiv ity
(8-150e)
and
** ^ %P
^ w(r, r') = - r') exp - (r - r') • A (r)
(8-150f)
The singularity of K 0 as r r' arises from the zero-range
two-body potential. If one cuts off the potential Vkk> outside
the energy range —cu0 oj0 (centered about the Fermi surface),
one has
J K 0(R) d R = N { 0) J “° i tanh ^ de
(8-151)
In the reduction we have used the equation determining k BT c =
1l(3c. By expanding the normal metal Green’s functions in powers
of the small quantity (e/c)(r — r') • A(r) and assuming that A(r)
varies slowly over a coherence length, Gor’kov obtains the
equation
lG 12{-nkBT cf (8-153)
and i(x) is the Riemann zeta function.
Ele ctrom agn e tic P ro pe rties o f S u p e rc o n d u c to rs 253
as in the GL theory.
Recently, the derivation of Gor’kov has been extended to all
temperatures by W ertham er154 and by Tewordt,155 who continue
to assume the system is such that A and A vary slowly over a
coherence length. Their equations are somewhat more compli
cated than the GLG form, as one might expect. Gor’kov has
extended his treatm ent to include finite mean-free-path effects.
He finds the equations have the same form as above, except that
the “mass” m is increased relative to that of the pure material.
CONCLUSION
SECOND-QUANTIZATION
FORMALISM
A-1 O C C U P A T IO N - N U M B E R R E P R E S E N T A T IO N
Let us consider a system of n identical particles described
in the Schrodinger representation by the Hamiltonian
H {x1 • • -xn) = 2 ^ + 2 CiOCj) + \ y v 2(x{, X j ) (A-l)
The coordinate xt labels the position and spin of particle i.
Three-body potentials and higher interactions can be included in
a straightforward manner, but for the moment we shall confine
ourselves to two-body interactions.
The many-body Schrodinger equation is
H(x, •• xnm x , •••*„,«) = ih 8W(Xldt 'Xn’ t] (A-2)
257
258 T h e o ry o f Su p e rco n d u ctiv ity
0 • -ukn(xn) (A-5)
where i f = (1/n!) 2 R in Bose statistics and i f = (1/n!) 2 (~ 1)pP
in Fermi statistics and the summation is over all n! possible
permutations of the coordinates • -xn and p is the order of
the permutation. Rather than labeling 0 by the quantum
numbers k x, k2 - k n, we may specify the state by stating how
many times each single-particle state enters the product. Let
this occupation number be n k for state k. Then the set of numbers
n l9 n 2 - • -nk uniquely specifies the symmetrized state 0 nitn2• -nk-
If we describe a system of zi-particles, we have clearly n k = n.
For Fermi statistics the occupation numbers n k are restricted to
the values 0 and 1, whereas for Bose statistics they can have all
possible positive integer values (as well as 0). The functions
form a complete orthonormal set of n-particle
^ r t ! ■■■ nk - - ( x i ‘ * ’ x n)
A-2 S E C O N D Q U A N T IZ A T IO N F O R B O S O N S
For Bose statistics we introduce a set of operators a k and a k +
defined by
“ k + t f V - . n * . •■(*!• ' - * n ) = K + - - ■ -Xn + 1)
(A-8)
•(*!• ' 'xn) = (nk)m ^ ni---nk - l - - i x l - ' ' - 1)
The operator a k + (creation operator) creates an additional particle
in state k , and a k (annihilation operator) destroys a particle in
state k. If nk = 0 in 0, the operator a k gives 0.
We can see that a k + is the Hermitian conjugate of a k by noting
that the only nonvanishing matrix element of a k is
(a ) n * Uk ~ 1
which is equal to (nk)112, and its Hermitian conjugate operator will
have as its only nonvanishing matrix element
* = K ) 1'2
The operator which has only this nonvanishing matrix element is
indeed a k + by definition. If we define a new operator N k = a k +a k,
it follows from definition (A-8) that its eigenvalue equation is
Nk®ni...nir..(xi- ■•*„) = nk0 ni...nk...(x1- • -xn) (A-9)
Therefore, N k may be interpreted as the operator which measures
the number of particles of state k (the number operator). We
may now construct the operator N which measures the total
number of particles in the system:
N = 2 = 2 a k +Clk (A-10)
260 T h e o ry o f Su p e rco n d u ctiv ity
For instance,
(aka k + - rii •••n k ••• 10
= [(% + 1) - nk]0,rix •••n k •• . = , •••n k
rii
fci
+ \ ^ Uk1 {n k1 ~
(A-13)
2. We now illustrate off-diagonal elements between wave
functions which differ in the occupation numbers of two states
i and j. Let the wave functions be
• rt,rty • • • and 0
with
n i + nj = m i + ra;
because the number of particles is fixed. The only matrix
elements different from zero are those for which either
(1) n} = rrij + 1 ft* = m i ± 1
or
(2) +2 = ra* + 2
On choosing the upper signs, the matrix elements for the first
case are
l \k,y[(ml + 1)mj]112 + + 1)m,]112
x V\k,k,} + ( k lk i\V\kjk l')) (A-14)
and for the second case
- l)(m, + !)("h + 2)]ll2( k jkj \V2\kik i')
On choosing the lower signs, we have the same expressions with
i and j interchanged.
3. We show here off-diagonal elements between wave
functions which differ in the occupation numbers of only three
states i, j , and I. Those different from zero obey conditions of the
type
ni — mx ± 1 7ij = mj ± 1 nx = m x + 2
262 T h e o ry o f Su p e rco n d u ctiv ity
When the uk(x) are given simply by plane waves eik xand we
normalize in a box of unit volume, we have k' = k —q and,
consequently,
p = 2 a k, + a k' + q (A- 2 1 )
k'
We may express the Hamiltonian operator in terms of the
field variable i(j (x ); from Eq. (A-12) and the definition of the iff we
obtain
H = j i/j +(x )H1(x )i/j (x ) dx
The order of the operators ensures that the term i = j has been
omitted in the two-body potential. If V2{x, x ) = V2(x — #'),
that is to say, if our two-body operator is translationally invariant,
the Hamiltonian can be written as
H = [ 4 , + {x)H1{x)>li(x)dx + V 2 ( 9 ) v ( P q + Pq) ( A- 2 3 )
J q
where rj is the normally ordered product such that all the i/j + are
placed to the left and all ifj to the right in the product. The
proof is obtained by expanding V2(x — x ) as its Fourier trans
form and using the definition of p q.
The prescription for expressing an n-body interaction in the
occupation-number representation is now clear:
Vv = t J V + (*i)- • - f W F v f e - • - xv)4i(xv)- ■-^(x^ dx1- ■■dxv
( A- 2 4 )
I. I ntroduction
It gives me great pleasure to have the opportunity to join my colleagues John
Bardeen and Leon Cooper in discussing with you the theory of superconduct
ivity. Since the discovery of superconductivity by H. Kamerlingh Onnes in
1911, an enormous effort has been devoted by a spectrum of outstanding scien
tists to understanding this phenomenon. As in most developments in our branch
of science, the accomplishments honored by this Nobel prize were made
possible by a large number of developments preceding them. A general under
standing of these developments is important as a backdrop for our own contri
bution.
On December 11, 1913, Kamerlingh Onnes discussed in his Nobel lecture (1)
his striking discovery that on cooling mercury to near the absolute zero of tem
perature, the electrical resistance became vanishingly small, but this dis
appearance “did not take place gradually but abruptly.” His Fig. 17 is re
produced as Fig. 1. He said, “Thus, mercury at 4.2 K has entered a new state
Fig. 1
267
which owing to its particular electrical properties can be called the state of
superconductivity.55 H e found this state could be destroyed by applying a
sufficiently strong magnetic field, now called the critical field Hc. In April —
June, 1914, Onnes discovered that a current, once induced in a closed loop of
superconducting wire, persists for long periods without decay, as he later graphi
cally demonstrated by carrying a loop of superconducting wire containing a
persistent current from Leiden to Cambridge.
In 1933, W. Meissner and R. Ochsenfeld (2) discovered that a superconductor
is a perfect diamagnet as well as a perfect conductor. The magnetic field van
ishes in the interior of a bulk specimen, even when cooled down below the
transition temperature in the presence of a magnetic field. The diamagnetic
currents which flow in a thin penetration layer near the surface of a simply
connected body to shield the interior from an externally applied field are stable
rather than metastable. On the other hand, persistent currents flowing in a
multiply connected body, e.g., a loop, are metastable.
An important advance in the understanding of superconductivity occurred
in 1934, when C. J. Gorter and H. B. G. Casimir (3) advanced a two fluid
model to account for the observed second order phase transition at Tc and
other thermodynamic properties. They proposed that the total density of
electrons g could be divided into two components
e= (i)
where a fraction Qs/Qn of the electrons can be viewed as being condensed into a
“superfluid,55 which is primarily responsible for the remarkable properties of
superconductors, while the remaining electrons form an interpenetrating
fluid of “normal” electrons. The fraction QslQn grows steadily from zero at Tc
to unity at T = 0, where “all of the electrons55 are in the superfluid condensate.
A second important theoretical advance came in the following year, when
Fritz and Hans London set down their phenomenological theory of the electro
magnetic properties of superconductors, in which the diamagnetic rather than
electric aspects are assumed to be basic. They proposed that the electrical
current density j s carried by the superfluid is related to the magnetic vector
potential A at each point in space by
j8= - LAc a (2)
where A is a constant dependent on the material in question, which for a free
electron gas model is given by A — m lg se 2, m and e being the electronic mass
and charge, respectively. A is to be chosen such that p • A = 0 to ensure cur
rent conservation. From (2) it follows that a magnetic field is excluded from a
superconductor except within a distance
— J Ac214:71
which is of order 10-6 cm in typical superconductors for T well below Tc.
Observed values of A are generally several times the London value.
In the same year (1935) Fritz London (4) suggested how the diamagnetic
268
property (2) might follow from quantum mechanics, if there was a “rigidity”
or stiffness of the wavefunction ip of the superconducting state such that ip was
essentially unchanged by the presence of an externally applied magnetic field.
This concept is basic to much of the theoretical development since that time,
in that it sets the stage for the gap in the excitation spectrum of a supercon
ductor which separates the energy of superfluid electrons from the energy of
electrons in the normal fluid. As Leon Cooper will discuss, this gap plays a
central role in the properties of superconductors.
In his book published in 1950, F. London extended his theoretical conjec
tures by suggesting that a superconductor is a “quantum structure on a macro
scopic scale [which is a] kind of solidification or condensation of the average
momentum distribution” of the electrons. This momentum space condensation
locks the average momentum of each electron to a common value which ex
tends over appreciable distance in space. A specific type of condensation in
momentum space is central to the work Bardeen, Cooper and I did together.
It is a great tribute to the insight of the early workers in this field that many
of the important general concepts were correctly conceived before the micro
scopic theory was developed. Their insight was of significant aid in our own
work.
The phenomenological London theory was extended in 1950 by Ginzburg
and Landau (5) to include a spatial variation of q8. They suggested that
Qs/ q be written in terms of a phenomenological condensate wavefunction ip(r)
as g«(r)/p = |v;(r)|2 and that the free energy difference AF between the
superconducting and normal states at temperature T be given by
AF . , _ (F +
. . h*
“ A(r) ) V>(r) - a ( T ) \v (r)\* + bS Il ip(r) •d*r (3)
' - / { i2m
where e, m, a and b are phenomenological constants, with a(Tc) = 0.
They applied this approach to the calculation of boundary energies between
normal and superconducting phases and to other problems.
As John Bardeen will discuss, a significant step in understanding which forces
cause the condensation into the superfluid came with the experimental discov
ery of the isotope effect by E. M axwell and, independently, by Reynolds, et al.
(6). Their work indicated that superconductivity arises from the interaction
of electrons with lattice vibrations, or phonons. Quite independently, Herbert
Frohlich (7) developed a theory based on electron-phonon interactions which
yielded the isotope effect but failed to predict other superconducting properties.
A somewhat similar approach by Bardeen (8) stimulated by the isotope effect
experiments also ran into difficulties. N. Bohr, W. Heisenberg and other
distinguished theorists had continuing interest in the general problem, but met
with similar difficulties.
An important concept was introduced by A. B. Pippard (9) in 1953. On the
basis of a broad range of experimental facts he concluded that a coherence
length £ is associated with the superconducting state such that a perturbation
of the superconductor at a point necessarily influences the superfluid within a
distance £ of that point. For pure metals, £ ~ 10~4 cm. for T Tc. H egener-
269
alized the London equation (3) to a non-local form and accounted for the
fact that the experimental value of the penetration depth is several times
larger than the London value. Subsequently, Bardeen (10) showed that
Pippard’s non-local relation would likely follow from an energy gap model.
A major problem in constructing a first principles theory was the fact that
the physically important condensation energy AF amounts typically to only
10-8 electron volts (e.V.) per electron, while the uncertainty in calculating
the total energy of the electron-phonon system in even the normal state
amounted to of order 1 e.V. per electron. Clearly, one had to isolate those
correlations peculiar to the superconducting phase and treat them accurately,
the remaining large effects presumably being the same in the two phases and
therefore cancelling. Landau’s theory of a Fermi liquid (11), developed to
account for the properties of liquid H e3, formed a good starting point for such a
scheme. Landau argued that as long as the interactions between the particles
(H e3 atoms in his case, electrons in our case) do not lead to discontinuous
changes in the microscopic properties of the system, a “quasi-particle” de
scription of the low energy excitations is legitim ate; that is, excitations of the
fully interacting normal phase are in one-to-one correspondence with the
excitations of a non-interacting fermi gas. The effective mass m and the Fermi
velocity vp of the quasi-particles differ from their free electron values, but aside
from a weak decay rate which vanishes for states at the Fermi surface there is
no essential change. It is the residual interaction between the quasi-particles
which is responsible for the special correlations characterizing superconductivi
ty. The ground state wavefunction of the superconductor \p0is then represented
by a particular superposition of these normal state configurations, &n.
A clue to the nature of the states @n entering strongly in yi0 is given by com
bining Pippard’s coherence length £ with Heisenberg’s uncertainty principle
Ap ~ H/g ~ 10-*pF (4)
where pp is the Fermi momentum. Thus, W0 is made up of states with quasi
particles (electrons) being excited above the normal ground state by a
momentum of order Ap. Since electrons can only be excited to states which are
initially empty, it is plausible that only electronic states within a momentum
10-4 pp of the Fermi surface are involved significantly in the condensation,
i.e., about 10-4 of the electrons are significantly affected. This view fits nicely
with the fact that the condensation energy is observed to be of order 10_4@-
ksTc. Thus, electrons within an energy ~ vpAp ~ kT c of the Fermi surface
have their energies lowered by of order kT c in the condensation. In summary,
the problem was how to account for the phase transition in which a condensa
tion of electrons occurs in momentum space for electrons very near the Fermi
surface. A proper theory should automatically account for the perfect conduc
tivity and diamagnetism, as well as for the energy gap in the excitation
spectrum.
II. T h e P a ir in g C o n c e p t
In 1955, stimulated by writing a review article on the status of the theory
of superconductivity, John Bardeen decided to renew the attack on the problem.
270
H e invited Leon Cooper, whose background was in elementary particle physics
and who was at that time working with C. N. Yang at the Institute for Advanced
Study to join in the effort starting in the fall of 1955. I had the good fortune
to be a graduate student of Bardeen at that time, and, having finished my
graduate preliminary work, I was delighted to accept an invitation to join them.
W e focused on trying to understand how to construct a ground state W0
formed as a coherent superposition of normal state configurations 0 n,
W0 = Z a n0 n (5)
n
such that the energy would be as low as possible. Since the energy is given in
terms of the Hamiltonian H by
E0 = (W0, HWo) = 2 an'*On H 0 n)
n,n
(0n'9 (6)
we attempted to make E0minimum by restricting the coefficients an so that only
states which gave negative off-diagonal matrix elements would enter (6). In
this case all terms would add in phase and E0 would be low.
By studying the eigenvalue spectrum of a class of matrices with off-diagonal
elements all of one sign (negative), Cooper discovered that frequently a single
eigenvalue is split off from the bottom of the spectrum. H e worked out the
problem of two electrons interacting via an attractive potential- V above a
quiescent Fermi sea, i.e., the electrons in the sea were not influenced by V and
the extra pair was restricted to states within an energy Hcdd above the Fermi
surface, as illustrated in Fig. 2. As a consequence of the non-zero density of
quasi-particle states JV(0) at the Fermi surface, he found the energy eigenvalue
spectrum for two electrons having zero total momentum had a bound state
split off from the continuum of scattering states, the binding energy being
2
E b = Hcdbc --------- (7)
JV( 0 ) v K '
if the matrix elements of the potential are constant equal to V in the region of
interaction. This important result, published in 1956 (12), showed that, re
gardless of how weak the residual interaction between quasi-particles is, if the
interaction is attractive the system is unstable with respect to the formation of
bound pairs of electrons. Further, if E b is taken to be of order ksTc, the un
certainty principle shows the average separation between electrons in the bound
state is of order 10-4 cm.
W hile Cooper’s result was highly suggestive, a major problem arose. If,
as we discussed above, a fraction 10-4 of the electrons is significantly involved
in the condensation, the average spacing between these condensed electrons
271
is roughly 10-# cm. Therefore, within the volume occupied by the bound state
o f a given pair, the centers o f approximately (lCM/lO-6)8 £ 10® other pairs will
be found, on the average. Thus, rather than a picture of a dilute gas of strongly
bound pairs, quite the opposite picture is true. The pairs overlap so strongly
in space that the mechanism of condensation would appear to be destroyed
due to the numerous pair-pair collisions interrupting the binding process of
a given pair.
Returning to the variational approach, we noted that the matrix elements
(&n'i H<Pn) in (6) alternate randomly in sign as one randomly varies n and ri
over the normal state configurations. Clearly this cannot be corrected to obtain
a low value of E0 by adjusting the sign of the an’s since there are JV2 matrix
elements to be corrected with only j\f parameters an. We noticed that if the
sum in (6) is restricted to include only configurations in which, if any quasi
particle state, say k , s} is occupied (s = j or j is the spin index), its “m ate”
state k , s is also occupied, then the matrix elements of H between such states
would have a unique sign and a coherent lowering of the energy would be
obtained. This correlated occupancy of pairs of states in momentum space is
consonant with London’s concept of a condensation in momentum.
In choosing the state k , s to be paired with a given state kf s , it is important
to note that in a perfect crystal lattice, the interaction between quasi-particles
conserves total (crystal) momentum. Thus, as a given pair of quasi-particles
interact, their center of mass momentum is conserved. To obtain the largest
number of non-zero matrix elements, and hence the lowest energy, one must
choose the total momentum o f each pair to be the same, that is
k-{-k = q. (8)
States with q # 0 represent states with net current flow. The lowest energy
state is for q = 0, that is, the pairing is such that if any state A:j is occupied in
an admissible 0 n , so is—A;j occupied. The choice of jj spin pairing is not
restrictive since it encompasses triplet and singlet paired states.
Through this reasoning, the problem was reduced to finding the ground state
of the reduced Hamiltonian
tfred = 2 e* 71*. - 2 V f t bk'+bk. (9)
ks kk'
The first term in this equation gives the unperturbed energy of the quasi
particles forming the pairs, while the second term is the pairing interaction
in which a pair of quasi-particles in (A;j, —A:j) scatter to (£'J, —k’J). The
operators b\ = ckJ C- b e i n g a product of two fermion (quasi-particle)
creation operators, do not satisfy Bose statistics, since bk+2 = 0. This point is
essential to the theory and leads to the energy gap being present not only for
dissociating a pair but also for making a pair move with a totalmomentum
different from the common momentum of the rest of the pairs. It is this feature
which enforces long range order in the superfluid over macroscopic distances.
III. T h e G ro u n d S t a t e
In constructing the ground state wavefunction, it seemed clear that the average
occupancy of a pair state (Arf, —£[) should be unity for k far below the Fermi
272
surface and 0 for k far above it, the fall off occurring symmetrically about kp
over a range of momenta of order
~ 104 cm-1.
One could not use a trial W0 as one in which each pair state is definitely oc
cupied or definitely empty since the pairs could not scatter and lower the
energy in this case. Rather there had to be an amplitude, say vjc, that (£j,
—A;j) is occupied in and consequently an amplitude ujc = J\ —VJ* that the
pair state is empty. After we had made a number of unsuccessful attempts
to construct a wavefunction sufficiently simple to allow calculations to be
carried out, it occurred to me that since an enormous number ( ~ 1019) of pair
states (A;'|, —A;'j) are involved in scattering into and out of a given pair state
(yfcj, —&[), the “instantaneous” occupancy of this pair state should be essen
tially uncorrelated with the occupancy of the other pair states at that “instant”.
Rather, only the average occupancies of these pair states are related.
On this basis, I wrote down the trial ground state as a product of operators
—one for each pair state—acting on the vacuum (state of no electrons),
= n {uk+vkbk) |0>, (10)
where m* = J i _ ^ 2 Since the pair creation operators bk+ commute for different
k?s, it is clear that represents uncorrelated occupancy of the various pair
states. I recall being quite concerned at tfye time that was an admixture of
states with different numbers of electrons, a wholly new concept to me, and as
I later learned to others as well. Since by varying vjc the mean number of
electrons varied, I used a Lagrange multiplier (the chemical potential) to
make sure that the mean number of electrons (JVop) represented by was the
desired number N. Thus by minimizing
E0—/j,N = (!P0, [//r e d -^ o p ]^ o )
with respect to vjc, I found that vjc was given by
z>* —1
where
Ek —J (ek—fiy+Ajc2 (12)
and the parameter Ak satisfied what is now called the energy gap equation:
(13)
From this expression, it followed that for the simple model
{
F, |e* —//| CncoB and |e*'—/J < Hcojy
0 ..
Vk'K u .
, otherwise
273
and the condensation energy at zero temperature is
AF = \N{0)A* (15)
The idea occurred to me while I was in New York at the end of January,
1957, and I returned to Urbana a few days later where John Bardeen quickly
recognized what he believed to be the essential validity of the scheme, much to
my pleasure and amazement. Leon Cooper will pick up the story from here to
describe our excitement in the weeks that followed, and our pleasure in un
folding the properties of the excited states.
IV. Q u an tu m P h en om en a o n a M a c r o sc o p ic S c a le
Superconductors are remarkable in that they exhibit quantum effects on a
broad range of scales. The persistence of current flow in a loop of wire many
meters in diameter illustrates that the pairing condensation makes the super
fluid wavefunction coherent over macroscopic distances. On the other hand,
the absorption of short wavelength sound and light by a superconductor is
sharply reduced from the normal state value, as Leon Cooper will discuss.
I will concentrate on the large scale quantum effects here.
The stability of persistent currents is best understood by considering a cir
cular loop of superconducting wire as shown in Fig. 3. For an ideal small
diameter wire, one would use the eigenstates e<m0, (m = 0 ,± 1 ,± 2 , . . .), of the
angular momentum Lz about the symmetry axis to form the pairing. In the
ground state no net current flows and one pairs m f with — mj, instead of
with —£[ as in a bulk superconductor. In both cases, the paired states are
time reversed conjugates, a general feature of the ground state. In a current
carrying state, one pairs (m-fv)J with (— (v — 0 ,=b l , i 2 . . .), so
that the total angular momentum of each pair is identical, 2 hv. It is this com
m onality of the center of mass angular momentum of each pair which preserves
the condensation energy and long range order even in states with current flow.
Another set of flow states which interweave with these states is formed by
pairing (m +v)J with (—m + r + l)J, (v = 0 ,± 1 ,± 2 . . .), with the pair
angular momentum being (2v-{-l)H. The totality of states forms a set with all
integer multiples n of h for allowed total angular momentum of pairs. Thus,
even though the pairs greatly overlap in space, the system exhibits quantiza
tion effects as if the pairs were well defined.
There are two important consequences of the above discussion. First, the
fact that the coherent condensate continues to exist in flow states shows that
to scatter a pair out of the (rotating) condensate requires an increase of energy.
274
Crudely speaking, slowing down a given pair requires it ot give up its binding
energy and hence this process will occur only as a fluctuation. These fluctua
tions average out to zero. The only way in which the flow can stop is if all pairs
simultaneously change their pairing condition from, say, v to v — 1. In this
process the system must fluctate to the normal state, at least in a section of the
wire, in order to change the pairing. This requires an energy of order the
condensation energy AF. A thermal fluctuation of this size is an exceedingly
rare event and therefore the current persists.
The second striking consequence of the pair angular momentum quantization
is that the magnetic flux 0 trapped within the loop is also quantized,
<Pn= n ~
2e (* = 0,±1,±2...). (16)
This result follows from the fact that if the wire diameter d is large compared
to the penetration depth A, the electric current in the center of the wire is
essentially zero, so that the canonical angular momentum of a pair is
2e
•f-pair == —C fpair X A (17)
where rpair is the center of mass coordinate of a pair and A is the magnetic
vector potential. If one integrates Lpair, around the loop along a path in the
center of the wire, the integral is nh, while the integral of the right hand side of
2e ^
(17) is — 0 .
c
A similar argument was given by F. London (4b) except that he considered
only states in which the superfluid flows as a whole without a change in its
internal strucutre, i.e., states analogous to the (iw +v)|, (— w + r) J set. H e found
0 z = n-hcje. The pairing (m + r)j, (m -fv + l)j cannot be obtained by adding
v to each state, yet this type of pairing gives an energy as low as the more
conventional flow states and these states enter experimentally on the same basis
as those considered by London. Experiments by Deaver and Fairbank (13),
and independently by Doll and Nabauer (14) confirmed the flux quantization
phenomenon and provided support for the pairing concept by showing that
2e rather than e enters the flux quantum. Following these experiments a clear
discussion of flux quantization in the pairing scheme was given by Beyers and
Yang (15).
The idea that electron pairs were somehow important in superconductivity
has been considered for some time (16, 17). Since the superfluidity of liquid
H e4 is qualitatively accounted for by Bose condensation, and since pairs of
electrons behave in some respects as a boson, the idea is attractive. The
essential point is that while a dilute gas of tightly bound pairs of electrons might
behave like a Bose gas (18) this is not the case when the mean spacing between
pairs is very small compared to the size of a given pair. In this case the inner
structure of the pair, i.e., the fact that it is made o f fermions, is essential;
it is this which distinguishes the pairing condensation, with its energy gap for
single pair translation as well as dissociation, from the spectrum of a Bose con-
275
OXIDE BARRIER = 0
densate, in which the low energy exictations are Bose-like rather than Fermi-
like as occurs in acutal superconductors. As London emphasized, the con
densation is an ordering in occupying momentum space, and not a space-like
condensation of clusters which then undergo Bose condensation.
In 1960, Ivar Giaever (19) carried out pioneering experiments in which elec
trons in one superconductor (Sx) tunneled through a thin oxide layer ( ~ 20 —
30 A) to a second superconductor (S2) as shown in Fig. 4. Giaever’s experi
ments were dramatic evidence of the energy gap for quasi-particle excitations.
Subsequently, Brian Josephson made a highly significant contribution by
showing theoretically that a superfluid current could flow between Sx and S2
with zero applied bias. Thus, the superfluid wavefunction is coherent not only
in St and S2 separately, but throughout the entire system, Sx—0 —S2, under
suitable circumstances. W hile the condensate amplitude is small in the oxide,
it is sufficient to lock the phases of Sx and S2 together, as has been discussed in
detail by Josephson (20) and by P. W. Anderson (21).
To understand the meaning of phase in this context, it is useful to go back
to the ground state wavefunction Wo, (10). Suppose we write the parameter v*
as |z>*| exp i<p and choose m* to be real. If we expand out the k-product in
we note that the terms containing jV pairs will have a phase factor exp (i N<p),
that is, each occupied pair state contributes a phase q>to W0. Let this wavefunc
tion, say IfV1* represent Sl3 and have phase <px. Similarly, let !?V2) represent St
and have phase angle q>2. If we write the state of the combined system as a
product
(18)
then by expanding out the double product we see that the phase of that part of
!F0<1»2) which has N x pairs in Sx and N 2 pairs in S2 is N x <px+ N 2<p2. For a truly
isolated system, 2(AiH-A2) = 2JV is a fixed number of electrons; however N x
and JV2 are not separately fixed and, as Josephson showed, the energy of the
combined system is minimized when gol = (p2 due to tunneling of electrons
between the superconductors. Furthermore, if (px = cp2, a current flows between
Sj and S2
j = j x sin(<px—<p2) (19)
If <px—<p2 = <p is constant in time, a constant current flows with no voltage
applied across the junction. If a bias voltage is V applied between St and S2,
then, according to quantum mechanics, the phase changes as
276
2 eV d q>
h ~ dt
(20)
Hence a constant voltage applied across such a junction produces an
alternating current of frequency
v = ------= 483 TH z/V . (21)
h
These effects predicted by Josephson were observed experimentally in a
series of beautiful experiments (22) by many scientists, which I cannot discuss
in detail here for lack of time. I would mention, as an example, the work of
Langenberg and his collaborators (23) at the University of Pennsylvania on
the precision determination of the fundamental constant ejh using the fre
quency-voltage relation obeyed by the alternating Josephson supercurrent.
These experiments have decreased the uncertainty in our experimental knowl
edge of this constant by several orders of magnitude and provide, in combina
tion with other experiments, the most accurate available value of the Sommerfeld
fine structure constant. They have resulted in the resolution of several dis
crepancies between theory and experiment in quantum electrodynamics and
in the development of an “atomic” voltage standard which is now being used
by the United States National Bureau of Standards to maintain the U .S. legal
volt.
V. C o n clu sio n
As I have attempted to sketch, the development of the theory of superconduct
ivity was truly a collaborative effort, involving not only John Bardeen, Leon
Cooper and myself, but also a host of outstanding scientists working over a
period of half a century. As my colleagues will discuss, the theory opened up
the field for many exciting new developments, both scientific and technological,
many of which no doubt lie in the future. I feel highly honored to have played
a role in this work and I deeply appreciate the honor you have bestowed on me
in awarding us the Nobel prize.
R eferences
1. Kamerlingh Onnes, H., Nobel Lectures, Vol. 1, pp. 306 —336.
2. Meissner, W. and Ochsenfeld, R., Naturwiss. 21, 787 (1933).
3. Gorter, C. J. and Gasimir, H. B. G., Phys. Z. 35, 963 (1934); Z. Techn. Phys. 15,
539 (1934).
4. London, F., [a] Phys. Rev. 24, 562 (1948); [b] Superfluids, Vol. 1 (John Wiley & Sons,
New York, 1950).
5. Ginzburg, V. L. and Landau, L. D., J. Exp. Theor. Phys. (U.S.S.R.) 20, 1064 (1950).
6. Maxwell, E., Phys. Rev. 78, M l (1950); Reynolds, C. A., Serin, B., Wright W. H. and
Nesbitt, L. B., Phys. Rev. 78, 487 (1950).
7. Frohlich, H., Phys. Rev. 79, 845 (1950).
277
8. Bardeen, J., Rev. Mod. Phys. 23, 261 (1951).
9. Pippard, A. B., Proc. Royal Soc. (London) A 216, 547 (1953).
10. Bardeen, J., [a] Phys. Rev. 97, 1724 (1955); [b] Encyclopedia of Physics, Vol. 15
(Springer-Verlag, Berlin, 1956), p. 274.
11. Landau, L. D., J. Exp. Theor. Phys. (U.S.S.R.) 30 (3), 1058 (920) (1956); 32 (5),
59 (101) (1957).
12. Cooper, L. N ., Phys. Rev. 104, 1189 (1956).
13. Deaver, B. S. Jr., and Fairbank, W. M., Phys. Rev. Letters 7, 43 (1961).
14. Doll, R. and Nabauer, M ., Phys. Rev. Letters 7, 51 (1961).
15. Beyers, N. and Yang, C. N ., Phys. Rev. Letters 7, 46 (1961).
16. Ginzburg, V. L., Usp. Fiz. Nauk 48, 25 (1952); transl. Fortsch. d. Phys. 1, 101 (1953).
17. Schafroth, M. R., Phys. Rev. 96, 1442 (1954); 100, 463 (1955).
18. Schafroth, M. R., Blatt, J. M. and Butler, S. T., Helv. Phys. Acta 30, 93 (1957).
19. Giaever, I., Phys. Rev. Letters 5, 147 (1960).
20. Josephson, B. D., Phys. Letters 1, 251 (1962); Advan. Phys. 14, 419 (1965).
21. Anderson, P. W ., in Lectures on the Many-body Problem, edited by E. R. Caianiello
(Academic Press, Inc. New York, 1964), Vol. II.
22. See Superconductivity, Parks, R. D., ed. (Dekker New York, 1969).
23. See, for example, Parker, W. H. Taylor B. N. and Langenberg, D. N. Phys. Rev.
Letters 18, 287 (1967); Finnegan, T. F. Denenstein A. and Langenberg, D. N. Phys.
Rev. B4, 1487 (1971).
278
M ICROSCOPIC QUANTUM INTERFERENCE
EFFECTS IN THE THEORY OF
SUPERCONDUCTIVITY
Nobel Lecture, December 11, 1972
by
L eon N C oo per
It is an honor and a pleasure to speak to you today about the theory of super
conductivity. In a short lecture one can no more than touch on the long history
of experimental and theoretical work on this subject before 1957. Nor can one
hope to give an adequate account of how our understanding of superconductivi
ty has evolved since that time. The theory (1) we presented in 1957, applied
to uniform materials in the weak coupling limit so defining an ideal supercon
ductor, has been extended in almost every imaginable direction. To these
developments so many authors have contributed (2) that we can make no
pretense of doing them justice. I will confine myself here to an outline of some
of the main features of our 1957 theory, an indication of directions taken since
and a discussion of quantum interference effects due to the singlet-spin pairing
in superconductors which might be considered the microscopic analogue of the
effects discussed by Professor Schrieffer.
N ormal M eta l
where F0(r ) is the periodic potential and in general might be a linear operator
to include exchange terms.
The Pauli exclusion principle requires that the many electron wave function
be antisymmetric in all of its coordinates. As a result no two electrons can be
279
The normal ground state wavefunction, An excitation of the normal system.
0 O> is a filled Fermi sphere for both spin
directions.
in the same single particle Bloch state. The energy of the entire system is
2JV
W = E £{
1= 1
where £* is the Bloch energy of the ith single electron state. The ground state
of the system is obtained when the lowest JV Bloch states of each spin are
occupied by single electrons; this can be pictured in momentum space as the
filling in of a Fermi sphere, Fig. 1. In the ground-state wave function there is
no correlation between electrons of opposite spin and only a statistical correla
tion (through the general anti-symmetry requirement on the total wave func
tion) of electrons of the same spin.
Single particle excitations are given by wave functions identical to the ground
state except that one electron states k i < k p are replaced by others k j < kp.
This may be pictured in momentum space as opening vacancies below the
Fermi surface and placing excited electrons above, Fig. 2. The energy difference
between the ground state and the excited state with the particle excitation kj
and the hole excitation ki is
S j — Si = £j — £ p — (£i — £ p ) = £ j — Si = |ty| + |£*|
where we define e as the energy measured relative to the Fermi energy
£i = £i — £ p .
W hen Coulomb, lattice-electron and other interactions, which have been
omitted in constructing the independent particle Bloch model are taken into
account, various modifications which have been discussed by Professor Schrief
fer are introduced into both the ground state wave function and the excitations.
These may be summarized as follows: The normal metal is described by a
ground state 0 O and by an excitation spectrum which, in addition to the
various collective excitations, consists of quasi-fermions which satisfy the usual
anticommutation relations. It is defined by the sharpness of the Fermi surface,
the finite density of excitations, and the continuous decline of the single particle
excitation energy to zero as the Fermi surface is approached.
280
E lectron C orrelations that P roduce S u perco nductivity
For a description of the superconducting phase we expect to include correla
tions that are not present in the normal metal. Professor Schrieffer has discussed
the correlations introduced by an attractive electron-electron interaction and
Professor Bardeen will discuss the role of the electron-phonon interaction in
producing the electron-electron interaction which is responsible for supercon
ductivity. It seems to be the case that any attractive interaction between the
fermions in a many-fermion system can produce a superconducting-like state.
This is believed at present to be the case in nuclei, in the interior of neutron
stars and has possibly been observed (4) very recently in H e3. We will therefore
develop the consequences of an attractive two-body interaction in a degenerate
many-fermion system without enquiring further about its source.
The fundamental qualitative difference between the superconducting and
normal ground state wave function is produced when the large degeneracy of
the single particle electron levels in the normal state is removed. If we visualize
the Hamiltonian matrix which results from an attractive two-body interaction
in the basis of normal metal configurations, we find in this enormous matrix,
sub-matrices in which all single-particle states except for one pair of electrons
remain unchanged. These two electrons can scatter via the electron-electron
interaction to all states of the same total momentum. We may envisage the
pair wending its way (so to speak) over all states unoccupied by other electrons.
[The electron-electron interaction in which we are interested is both weak
and slowly varying over the Fermi surface. This and the fact that the energy
involved in the transition into the superconducting state is small leads us to
guess that only single particle excitations in a small shell near the Fermi
surface play a role. It turns out, further, that due to exchange terms in the
electron-electron matrix element, the effective interaction in metals between
electrons of singlet spin is much stronger than that between electrons of triplet
spin— thus our preoccupation with singlet spin correlations near the Fermi
surface.] Since every such state is connected to every other, if the interaction
is attractive and does not vary rapidly, we are presented with submatrices of
the entire Hamiltonian of the form shown in Fig. 3. For purposes of illustration
we have set all off diagonal matrix elements equal to the constant— V and
the diagonal terms equal to zero (the single particle excitation energy at the
Fermi surface) as though all the initial electron levels were completely degener
ate. Needless to say, these simplifications are not essential to the qualitative
result.
Diagonalizing this matrix results in an energy level structure with M — 1
levels raised in energy to E = -f V while one level (which is a superposition
of all of the original levels and quite different in character) is lowered in energy
to
E = — (M — \)V.
Since M , the number of unoccupied levels, is proportional to the volume
of the container while V, the scattering matrix element, is proportional to
1/volume, the product is independent of the volume. Thus the removal of
281
M COLUMNS
0 -V -V -V
-V 0 -V •
-V-V 0 -V
M ROWS
-V 0 -V
-V -V -V-V 0
-V
- . . . 0
• • ■ • -(M-I)V
For V = 0, For V > 0,M — 1 levels at
M levels E = V and one level
at E = 0 at E = — (M—1) V
Fig. 3.
the degeneracy produces a single level separated from the others by a volume
independent energy gap.
To incorporate this into a solution of the full Hamiltonian, one must devise
a technique by which all of the electrons pairs can scatter while obeying the
exclusion principle. The wave function which accomplishes this has been dis
cussed by Professor Schrieffer. Each pair gains an energy due to the removal of
the degeneracy as above and one obtains the maximum correlation of the entire
wave function if the pairs all have the same total momentum. This gives a
coherence to the wave function in which forva combination of dynamical and
statistical reasons there is a strong preference for momentum zero, singlet spin
correlations, while for statistical reasons alone there is an equally strong
preference that all of the correlations have the same total momentum.
In what follows I shall present an outline of our 1957 theory modified by
introducing the quasi-particles of Bogoliubov and Valatin. (5) This leads to
a formulation which is generally applicable to a wide range of calculations
282
kt
Fig. 4.
The ground state of the superconductor is
a linear superposition of states in which pairs
(Atf —k l ) are occupied or unoccupied.
IS»T
“t
in a manner analogous to similar calculations in the theory o f normal metals.
We limit the interactions to terms which scatter (and thus correlate) singlet
zero-momentum pairs. To do this, it is convenient to introduce the pair
operators:
bk = C-gCK
, * * *
^k — CK C-K
and using these we extract from the full Hamiltonian the so-called reduced
Hamiltonian
•^reduced = E 2 |e | bk bk -j- X <2.ebk bk -f- 2J Vk,kbk,bk
k<kf k>kf kk'
where is the scattering matrix element between the pair states k and k'.
G round S tate
As Professor Schrieffer has explained, the ground state of the superconductor
is a linear superposition of pair states in which the pairs (k T , — k I ) are occupied
or unoccupied as indicated in Fig. 4. It can be decomposed into two disjoint
vectors—one in which the pair state k is occupied, 0 k and one in which it is
unoccupied, 0(ky
V»o =
The probability amplitude that the pair state k is (is not) occupied in the
ground state is then vk(uk). Normalization requires that |m|2 -(-|p |2 = 1 . The
phase of the ground state wave function may be chosen so that with no loss o
generality uk is real. We can then write
u = (1 - h ) 11*
v = h11* eiv
where
0 < h < 1.
A further decomposition of the ground state wave function of the supercon
ductor in which the pair states k and k ' are either occupied or unoccupied
Fig. 5 is:
ip0 = UkUV®{k),(V) + UkVV®(k), k' + VkUk’0k, (k') + VkVk,^ k tV •
This is a Hartree-like approximation in the probability amplitudes for the
occupation of pair states. It can be shown that for a fermion system the wave
283
Fig. 5.
A decomposition of the ground state of the superconductor into states in which the pair
states k and k' are either occupied or unoccupied.
function cannot have this property unless there are a variable number of
particles. To terms of order 1/JV, however, this decomposition is possible for
a fixed number of particles; the errors introduced go to zero as the number of
particles become infinite. (6 )
The correlation energy, WC) is the expectation value of HTe&for the state y)0
Wc = (yj0, Hre^xpo) = WQ [h,(p].
Setting the variation of Wc with respect to h and q>equal to zero in order to
minimize the energy gives
h = 1/2 (1 -e /E )
E = (£2+ |Zl|2) 1/2
where
A = |zl|ei,p
satisfies the integral equation
^ 1 + exp(|e|/A:BT)'
If we make our simplifications of 1957, (defining in this way an ‘ideal*
superconductor)
FjYk == F jfij < H(d$v
= 0 otherwise
and replace the energy dependent density of states by its value at the Fermi
surface, JV(0), the integral equation for A becomes
fcwav de / /ca 4 . a U\
1 = jV(0) V J - = = = = = tanh l f - ± £ L
oJ + \ 2kBT )
The solution of this equation, Fig. 7, gives A (T ) and with this f and h.
We can then calculate the free energy of the superconducting state and obtain
the thermodynamic properties of the system.
287
0 0.2 0.4 0.6 0.8 1.0
T / Tc
Fig. 7.
Variation of the energy gap with temperature for the ideal superconductor.
In particular one finds that at Tc (in the absence of a magnetic field) there
is a second-order transition (no latent heat: Wc = 0 at T c) and a discontinuity
in the specific heat. At very low temperatures the specific heat goes to zero
exponentially. For this ideal superconductor one also obtains a law of cor
responding states in which the ratio
where
y = 2 / 37r2jV(0 )£B2.
The experimental data scatter about the number 0.170. The ratio of A
to kBT c is given as a universal constant
AlkBTc = 1.75.
There are no arbitrary parameters in the idealized theory. In the region
o f empirical interest all thermodynamic properties are determined by the quanti
ties y and hwav e The first, y, is found by observation of the normal
specific heat, while the second is found from the critical temperature, given by
kBTc = 1.14 t o w e - i'W ,
At the absolute zero
zl =foJav/Sin h ^ J _ ) .
Further, defining a weak coupling limit [JV(0)F 1] which is one region
of interest empirically, we obtain
A ~ 2fc»ave-1/*<°>F.
The energy difference between the normal and superconducting states be
comes (again in the weak coupling limit)
W s-W n = Wc = —2N(0)(ficosw) 2 e- 2 /*<0)F
288
The dependence of the correlation energy on (&oav) 2 gives the isotope effect,
while the exponential factor reduces the correlation energy from the dimen-
sionally expected 7V(0)(^coav) 2 to the much smaller observed value. This,
however, is more a demonstration that the isotope effect is consistent with our
model rather than a consequence of it, as will be discussed further by Professor
Bardeen.
The thermodynamic properties calculated for the ideal superconductor are
in qualitative agreement with experiment for weakly coupled superconductors.
Very detailed comparison between experiment and theory has been made by
many authors. A summary of the recent status may be found in reference (2).
W hen one considers that in the theory of the ideal superconductor the existence
of an actual metal is no more than hinted at (We have in fact done all the
calculations considering weakly interacting fermions in a container.) so that
in principle (with appropriate modifications) the calculations apply to neutron
stars as well as metals, we must regard detailed quantitative agreement as a
gift from above. We should be content if there is a single metal for which such
agreement exists. [Pure single crystals of tin or vanadium are possible candi
dates.]
To make comparison between theory and experiments on actual metals, a
plethora of detailed considerations must be made. Professor Bardeen will
discuss developments in the theory of the electron-phonon interaction and the
resulting dependence of the electron-electron interaction and superconducting
properties on the phonon spectrum and the range of the Coulomb repulsion.
Crystal symmetry, Brillouin zone structure and the actual wave function (S,
P or D states) of the conduction electrons all play a role in determining real
metal behavior. There is a fundamental distinction between superconduc
tors w ich always show a Meissner effect and those (type II) which allow mag
netic field penetration in units of the flux quantum.
When one considers, in addition, specimens with impurities (magnetic and
otherwise) superimposed films, small samples, and so on, one obtains a variety
of situations, developed in the years since 1957 by many authors, whose rich
ness and detail takes volumes to discuss. The theory of the ideal superconductor
has so far allowed the addition of those extensions and modifications necessary
to describe, in what must be considered remarkable detail, all of the experience
actually encountered.
M icroscopic I nterference E ffects
In its interaction with external perturbations the superconductor displays
remarkable interference effects which result from the paired nature of the
wave function and are not at all present in similar normal metal interactions.
Neither would they be present in any ordinary two-fluid model. These “co
herence effects” are in a sense manifestations of interference in spin and
momentum space on a microscopic scale, analogous to the macroscopic
quantum effects due to interference in ordinary space which Professor Schrieffer
discussed. They depend on the behavior under time reversal of the perturbing
fields. (8 ) It is intriguing to speculate that if one could somehow amplify them
289
Fig. 8.
Ultrasonic attenuation as a function of temperature across the superconducting transition
as measured by Morse and Bohm.
properly, the time reversal symmetry of a fundamental interaction might be
tested. Further, if helium 3 does in fact display a phase transition analogous to
the superconducting transition in metals as may be indicated by recent experi
ments (4) and this is a spin triplet state, the coherences effects would be
greatly altered.
Near the transition temperature these coherence effects produce quite dra
matic contrasts in the behavior of coefficients which measure interactions with
the conduction electrons. Historically, the comparison with theory of the be
havior of the relaxation rate of nuclear spins (9) and the attenuation of longi
tudinal ultrasonic waves in clean samples ( 1 0 ) as the temperature is decreased
through 7"c provided an early test of the detailed structure of the theory.
The attenuation of longitudinal acoustic waves due to their interaction
with the conduction electrons in a metal undergoes a very rapid drop ( 1 0 a)
as the temperature drops below T c. Since the scattering of phonons from
“normal” electrons is responsible for most of the acoustic attenuation, a drop
was to be expected; but the rapidity o f the decrease measured by Morse and
Bohm (1 Ob) Fig. 8 was difficult to reconcile with estimates of the decrease in
the normal electron component of a two-fluid model.
The rate of relaxation of nuclear spins was measured by Hebei and Slichter
(9a) in zero magnetic field in superconducting aluminum from 0.94 K to
4.2 K just at the time of the development of our 1957 theory. Redfield and
Anderson (9b) confirmed and extended their results. The dominant relaxation
mechanism is provided by interaction with the conduction electrons so that
one would expect, on the basis of a two-fluid model, that this rate should
290
decrease below the transition temperature due to the diminishing density of
“normal” electrons. The experimental results however show just the reverse.
The relaxation rate does not drop but increases by a factor of more than two
just below the transition temperature. Fig 13. This observed increase in the
nuclear spin relaxation rate and the very sharp drop in the acoustic attenuation
coefficient as the temperature is decreased through T c impose contradictory
requirements on a conventional two-fluid model.
To illustrate how such effects come about in our theory, we consider the transi
tion probability per unit time of a process involving electronic transitions from
the excited state k to the state k' with the emission to or absorption of energy from
the interacting field. W hat is to be calculated is the rate of transition between
an initial state |i > and a final state |/ > with the absorption or emission of
the energy Ha)\k,-k\ (a phonon for example in the interaction of sound waves
with the superconductor). All of this properly summed over final states and
averaged with statistical factors over initial states may be w ritten:
2n I e x p (-W ,lk BT)\ < f \ H m |f > |* d(Wf - W , )
co = —h —--------------------------------------------------------------------
E exp (— WilksT)
W e focus our attention on the matrix element i > . This typically
contains as one of its factors matrix elements between excited states of the
superconductor of the operator
B = E BK'KCK'CK
KK'
where c*K, and cK are the creation and annihilation operators for electrons in
the states K' and K , and BK,K is the matrix element between the states K' and
K of the configuration space operator B(r)
BK,K = < K '\B ( r ) \K > .
The operator B is the electronic part of the matrix element between the full
final and initial state
< /|M n t|* '> = mft< f \ B \ i > .
In the normal system scattering from single-particle electron states K to K'
is independent of scattering from — K' to — K. But the superconducting states
are linear superpositions of (K,—K) occupied and unoccupied. Because of this
states with excitations k f and k' j are connected not only by but also
by c*kic_k,i ; if the state |f > contains the single-particle excitation k' | while
the state |z > contains k t, as a result of the superposition of occupied and
unoccupied pair states in the coherent part of the wave function, these are
connected not only by BK,K c*K,cK but also by B-K_K, c*Kc_K,.
For operators which do not flip spins we therefore write:
B = E (B k ,k ck,ck -j- B -k-.k, c_kc_k,).
k k'
Many of the operators, B , we encounter (e.g., the electric current, or the
charge density operator) have a well-defined behavior under the operation
of time reversal so that
B r 'k = i B_k _k, = BWk.
291
1
I
I Js ll> *
P P
Ufc. | v
P
k#
I
~ I
P P
Fig. 9.
The two states |t > and < f\ shown are connected by with the amplitude uvuk.
Then B becomes
B — Z Bk,k (cxfCkf i c_k^c_k,i)
kk'
where the upper (lower) sign results for operators even (odd) under time
reversal.
The matrix element of B between the initial state, tp . . . . . ., and the
final state y) . . . *,t . . . contains contributions from c*k,^ck^ Fig. 9 and un
expectedly from Fig. 10. As a result the matrix element squared
i > |2 contains terms of the form
|s *'k|2 |K-“*T ivO |a>
where the sign is determined by the behavior of B under time reversal:
upper sign B even under time reversal
lower sign B odd under time reversal.
Applied to processes involving the emission or absorption of boson quanta
such as phonons or photons, the squared matrix element above is averaged
with the appropriate statistical factors over initial and summed over final
states; substracting emission from absorption probability per unit time, we
obtain typically
292
P\M
Fig. 10.
The two states |t > and < / 1 are also connected by c*u^C-vi with the amplitude —vyvi •
a = — |w |2 E (/»■-/») H Ev - Ek—h(o
n kw
where f k is the occupation probability in the superconductor for the excitation
let or lej. [In the expression above we have considered only quasiparticle or
quasi-hole scattering processes (not including processes in which a pair of
excitations is created or annihilated from the coherent part of the wave
function) since < A, is the usual region of interest for the ultrasonic
attenuation and nuclear spin relaxation we shall contrast.]
For the ideal superconductor, there is isotropy around the Fermi surface
and symmetry between particles and holes; therefore sums of the form E can
k
be converted to integrals over the superconducting excitation energy, E:
E
E - • 2JV(0) dE
k E 2— A 2
E E
where jV(0 ) - y = = = = jV(0 ) —is the density of excitations in the super
conductor, Fig. 11.
293
Fig. 11.
i Ratio of superconducting to normal den-
3 E/A sity of excitations as a function of EfA.
294
Fig. 12.
Comparison of observed ultrasonic attenuation with the ideal theory. The data are due to
Morse and Bohtn.
. l/e
Since (m2—y2)s-> -l —\2
I , the coherence factors cancel the density of states
giving
^ = 2f { A ( T ) ) = ---------- -
Qn 1 + exp
Morse and Bohm (10b) used this result to obtain a direct experimental
determination of the variation of A with T. Comparison of their attenuation
data with the theoretical curve is shown in Figure 12.
295
In contrast the relaxation of nuclear spins which have been aligned in a
magnetic field proceeds through their interaction with the magnetic moment
o f the conduction electrons. In an isotropic superconductor this can be shown
to depend upon the z component of the electron spin operator
^ K 'K — B ( ck'\ck} c-klrc -k 'l)
so that
&KK — —B-k-k'-
This follows in general from the property of the spin operator under time
reversal
Oz(t) = — Gz( — t).
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Fig. 13.
Comparison of observed nuclear spin relaxation rate with theory. The circles represent
experimental data of Hebei and Slichter, the crosses data by Redfield and Anderson.
296
Interference effects manifest themselves in a similar manner in the interac
tion of electromagnetic radiation with the superconductor. Near T c the absorp
tion is dominated by quasi-particle scattering matrix elements of the type we
have described. Near T — 0, the number of quasi-particle excitations goes
to zero and the matrix elements that contribute are those in which quasi
particle pairs are created from y)0. For absorption these latter occur only
when H > 2A. For the linear response of the superconductor to a static
cd
magnetic field, the interference occurs in such a manner that the paramagnetic
contribution goes to zero leaving the diamagnetic part which gives the Meiss
ner effect.
The theory developed in 1957 and applied to the equilibrium properties
of uniform materials in the weak coupling region has been extended in numer
ous directions by many authors. Professor Schrieffer has spoken of Josephson
junctions and macroscopic quantum interference effects; Professor Bardeen
will discuss the modifications of the theory when the electron-phonon inter
actions are strong. The treatment of ultrasonic attenuation, generalized to
include situations in uniform superconductors in which ql < 1 , gives a sur
prisingly similar result to that above. (12) There have been extensive de
velopments using Green’s function methods (13) appropriate for type II super
conductors, materials with magnetic impurities and non-uniform materials or
boundary regions where the order parameter is a function of the spatial co
ordinates. (14) With these methods formal problems of gauge invariance and/or
current conservation have been resolved in a very elegant manner. (15) In
addition, many calculations (16) of great complexity and detail for type II
superconductors have treated ultrasonic attenuation, nuclear spin relaxation
and other phenomena in the clean and dirty limits (few or large numbers of
impurities). The results cited above are modified in various ways. For example,
the average density of excitation levels is less sharply peaked at T c in a type II
superconductor; the coherence effects also change somewhat in these altered
circumstances but nevertheless play an important role. Overall one can say
that the theory has been amenable to these generalizations and that agreement
with experiment is good.
It is now believed that the finite many-nucleon system that is the atomic
nucleus enters a correlated state analogous to that of a superconductor. (17)
Similar considerations have been applied to many-fermion systems as diverse
as neutron stars, (18) liquid H e3, (19) and to elementary fermions. (20) In
addition the idea of spontaneously broken symmetry of a degenerate vacuum
has been applied widely in elementary particle theory and recently in the
theory of weak interactions. (21) W hat the electron-phonon interaction has
produced between electrons in metals may be produced by the van der Waals
interaction between atoms in H e3, the nuclear interaction in nuclei and neutron
stars, and the fundamental interactions in elementary fermions. W hatever the
success of these attempts, for the theoretician the possible existence of this
correlated paired state must in the future be considered for any degenerate
many-fermion system where there is some kind of effective attraction between
fermions for transitions near the Fermi surface.
297
In the past few weeks my colleagues and I have been asked many tim es:
“W hat are the practical uses of your theory?” Although even a summary in
spection o f the proceedings of conferences on superconductivity and its appli
cations would give an immediate sense of the experimental, theoretical and
developmental work in this field as well as expectations, hopes and anticipa
tions—from applications in heavy electrical machinery to measuring devices
of extraordinary sensitivity and new elements with very rapid switching speeds
for computers — I, personally, feel somewhat uneasy responding. The discovery
of the phenomena and the development of the theory is a vast work to which
many scientists have contributed. In addition there are numerous practical
uses of the phenomena for which theory rightly should not take credit. A
theory (though it may guide us in reaching them) does not produce the trea
sures the world holds. And the treasures themselves occasionally dazzle our
attention; for we are not so wealthy that we may regard them as irrelevant.
But a theory is more. It is an ordering of experience that both makes ex
perience meaningful and is a pleasure to regard in its own right. Henri Poin-
car£ wrote (2 2 ):
Le savant doit ordonner; on fait la science
avec des faits comme une maison avec des
pierres; mais une accumulation de faits
n’est pas plus une science qu’un tas de
pierres n ’est une maison.
One can build from ordinary stone a humble house or the finest chateau.
Either is constructed to enclose a space, to keep out the rain and the cold.
T hey differ in the ambition and resources of their builder and the art by which
he has achieved his end. A theory, built of ordinary materials, also may serve
many a humble function. But when we enter and regard the relations in the
space of ideas, we see columns of remarkable height and arches of daring
breadth. They vault the fine structure constant, from the magnetic moment
of the electron to the behavior of metallic junctions near the absolute zero;
they span the distance from materials at the lowest temperatures to those in
the interior of stars, from the properties of operators under time reversal to the
behavior of attenuation coefficients just beyond the transition temperature.
I believe that I speak for my colleagues in theoretical science as well as
myself when I say that our ultimate, our warmest pleasure in the midst of one
o f these incredible structures comes with the realization that what we have
made is not only useful but is indeed a beautiful way to enclose a space.
298
5. Bogoliubov, N. N ., Nuovo Cimento 7, 794 (1958); Usp. Fiz. Nauk 67, 549 (1959);
Valatin, J. G., Nuovo Cimento 7, 843 (1958).
6. Bardeen, J. and Rickayzen, G., Phys. Rev. 118, 936 (1960); Mattis, D. C. and Lieb, E.,
J. Math. Phys. 2, 602 (1961); Bogoliubov, N. N ., Zubarev D. N. and Tserkovnikov, Yu.
A., Zh. Eksperim. i Teor. Fiz. 39, 120 (1960) translated: Soviet Phys. JETP 12, 88
(1961).
7. For example, Glover, R. E. I ll and Tinkham, M ., Phys. Rev. 108, 243 (1957); Biondi,
M. A. and Garfunkel, M. P., Phys. Rev. 116, 853 (1959).
8. The importance of the coupling of time reversed states in constructing electron pairs
was emphasized by P. W. Anderson,; for example, Anderson, P. W ., J. Phys. Chem.
Solids 11, 26 (1959)
9. a) Hebei, L. G. and Slichter, C. P., Phys. Rev. 113, 1504 (1959).
b) Redfield, A. G. and Anderson, A. G., Phys. Rev. 116,583 (1959).
10. a) Bommel, H. E., Phys. Rev. 96, 220 (1954).
b) Morse, R. W. and Bohm, H. V ., Phys. Rev. 108, 1094 (1957).
11. Fibich, M ., Phys. Rev. Letters 14, 561 (1965).
12. For example, Tsuneto T., Phys. Rev. 121, 402 (1961).
13. Gor’kov, L. P., Zh. Eksperim. i Teor. Fiz. 34, 735 (1958) translated: Soviet Physics
JETP 7, 505 (1958); also Martin P. G. and Schwinger J., Phys. Rev. 115, 1342 (1959);
Kadanoff L. P. and Martin P. C., Phys. Rev. 124, 670 (1961).
14. Abrikosov, A. A. and Gor’kov, L. P., Zh. Eksperim. i Teor. Fiz. 39, 1781 (1960) trans
lated: Soviet Physics JETP 12, 1243 (1961); de Gennes P. G., Superconductivity of Me
tals and Alloys, Benjamin, New York (1966).
15. For example, Ambegaokar V. and Kadanoff L. P., Nuovo Cimento 22, 914 (1961).
16. For example, Caroli C. and Matricon J., Physik Kondensierten Materie 3, 380 (1965);
Maki K., Phys. Rev. 141, 331 (1966); 156, 437 (1967); Groupe de Supraconductivit6
d’Orsay, Physik Kondensierten Materie 5, 141 (1966); Eppel D., Pesch W. and Te-
wordt L., Z. Physik 197; 46 (1966); McLean F. B. and Houghton A., Annals of Physics
48, 43 (1968).
17. Bohr, A., Mottelson, B. R. and Pines, D., Phys. Rev. 110, 936 (1958); Migdal, A. B.,
Nuclear Phys. 13, 655 (1959).
18. Ginzburg, V. L. and Kirzhnits, D. A., Zh. Eksperim. i Theor. Fiz. 47, 2006 (1964)
translated: Soviet Physics JETP 20, 1346 (1965); Pines D., Baym G. and Pethick C.,
Nature 224, 673 (1969).
19. Many authors have explored the possibility of a superconducting-like transition in He*.
Among the most recent contributions see reference 4.
20. For example. Nambu Y. and Jona-Lasinio G., Phys. Rev. 122, 345 (1961).
21. Goldstone, J., Nuovo Cimento 19, 154 (1961); Weinberg, S., Phys. Rev. Letters 19,
1264 (1967).
22. Poincar^, H., La Science et l’Hypoth£se, Flammarion, Paris, pg. 168 (1902). “The
scientist must order; science is made with facts as a house with stones; but an accumula
tion of facts is no more a science than a heap of stones is a house.”
299
ELECTRON-PHONON INTERACTIONS AND
SUPERCONDUCTIVITY
Nobel Lecture, December 11, 1972
By J ohn B a r deen
Departments of Physics and of Electrical Engineering
University of Illinois
Urbana, Illinois
1
I ntroductio n
Our present understanding of superconductivity has arisen from a close
interplay of theory and experiment. It would have been very difficult to have
arrived at the theory by purely deductive reasoning from the basic equations
of quantum mechanics. Even if someone had done so, no one would have be
lieved that such remarkable properties would really occur in nature. But, as
you well know, that is not the way it happened, a great deal had been learned
about the experimental properties of superconductors and phenomenological
equations had been given to describe many aspects before the microscopic
theory was developed. Some of these have been discussed by Schrieffer and
by Cooper in their talks.
M y first introduction to superconductivity came in the 1930’s and I greatly
profited from reading David Shoenberg’s little book on superconductivity, [ 1 ]
which gave an excellent summary of the experimental findings and of the
phenomenological theories that had been developed. At that time it was
known that superconductivity results from a phase change of the electronic
structure and the Meissner effect showed that thermodynamics could be
applied successfully to the superconductive equilibrium state. The two fluid
Gorter— Casimir model was used to describe the thermal properties and the
London brothers had given their famous phenomenological theory of the
electrodynamic properties. Most impressive were Fritz London’s speculations,
given in 1935 at a meeting of the Royal Society in London, [2] that super
conductivity is a quantum phenomenon on a macroscopic scale. He also gave
what may be the first indication of an energy gap when he stated that “the
electrons be coupled by some form of interaction in such a way that the
lowest state may be separated by a finite interval from the excited ones.”
H e strongly urged that, based on the Meissner effect, the diamagnetic aspects
of superconductivity are the really basic property.
M y first abortive attempt to construct a theory, [3] in 1940, was strongly
influenced by London’s ideas and the key idea was small energy gaps at the
Fermi surface arising from small lattice displacements. However, this work
was interrupted by several years of wartime research, and then after the war
I joined the group at the Bell Telephone Laboratories where my work turned
to semiconductors. It was not until 1950, as a result of the discovery of the
300
isotope effect, that I again began to become interested in superconductivity,
and shortly after moved to the University of Illinois.
The year 1950 was notable in several respects for superconductivity theory.
The experimental discovery of the isotope effect [4, 5] and the independent
prediction of H. Frohlich [6 ] that superconductivity arises from interaction
between the electrons and phonons (the quanta of the lattice vibrations) gave
the first clear indication of the directions along which a microscopic theory
might be, sought. Also in the same year appeared the phenomenological
Ginzburg— Landau equations which give an excellent description of super
conductivity near T c in terms of a complex order parameter, as mentioned
by Schrieffer in his talk. Finally, it was in 1950 that Fritz London’s book [7]
on superconductivity appeared. This book included very perceptive comments
about the nature of the microscopic theory that have turned out to be re
markably accurate. He suggested that superconductivity requires “a kind of
solidification or condensation of the average momentum distribution.” He
also predicted the phenomenon of flux quantization, which was not observed
for another dozen years.
The field of superconductivity is a vast one with many ramifications. Even
in a series of three talks, it is possible to touch on only a few highlights. In
this talk, I thought that it might be interesting to trace the development of
the role of electron-phonon interactions in superconductivity from its begin
nings in 1950 up to the present day, both before and after the development
of the microscopic theory in 1957. By concentrating on this one area, I hope
to give some impression of the great progress that has been made in depth
of understanding of the phenomena of superconductivity. Through develop
ments by many people, [8 ] electron-phonon interactions have grown from a
qualitative concept to such an extent that measurements on superconductors
are now used to derive detailed quantitative information about the interaction
and its energy dependence. Further, for many of the simpler metals and alloys,
it is possible to derive the interaction from first principles and calculate the
transition temperature and other superconducting properties.
The theoretical methods used make use of the methods of quantum field
theory as adopted to the many-body problem, including Green’s functions,
Feynman diagrams, Dyson equations and renormalization concepts. Following
Matsubara, temperature plays the role of an imaginary time. Even if you are
not familiar with diagrammatic methods, I hope that you will be able to
follow the physical arguments involved.
In 1950, diagrammatic methods were just being introduced into quantum
field theory to account for the interaction of electrons with the field of photons.
It was several years before they were developed with full power for application
to the quantum statistical mechanics of many interacting particles. Following
Matsubara, those prominent in the development of the theoretical methods
include Kubo, Martin and Schwinger, and particularly the Soviet physicists,
M igdal, Galitski, Abrikosov, Dzyaloshinski, and Gor'kov. The methods were
first introduced to superconductivity theory by Gor'kov [9] and a little later
in a somewhat different form by Kadanoff and Martin. [10] Problems of
301
superconductivity have provided many applications for the powerful Green’s
function methods of m any-body theory and these applications have helped to
further develop the theory.
Diagrammatic methods were first applied to discuss electron-phonon
interactions in normal metals by M igdal [11] and his method was extended
to superconductors by Eliashberg. [12] A similar approach was given by
Nambu. [13] The theories are accurate to terms of order (m /M )1/2, where m
is the mass of the electron and M the mass of the ion, and so give quite accurate
quantitative accounts of the properties of both normal metals and super
conductors.
W e will first give a brief discussion of the electron-phonon interactions as
applied to superconductivity theory from 1950 to 1957, when the pairing theory
was introduced, then discuss the M igdal theory as applied to normal metals,
and finally discuss Eliashberg’s extension to superconductors and subsequent
developments. W e will close by saying a few words about applications of the
pairing theory to systems other than those involving electron-phonon inter
actions in metals.
2
D evelopm ents from 1950— 1957
The isotope effect was discovered in the spring of 1950 by Reynolds, Serin,
et al, [4] at Rutgers University and by E. M axwell [5] at the U . S. National
Bureau of Standards. Both groups measured the transition temperatures of
separated mercury isotopes and found a positive result that could be interpreted
as T qM 112 ~ constant, where M is the isotopic mass. If the mass of the ions
is important, their motion and thus the lattice vibrations must be involved.
Independently, Frohlich, [ 6 ] who was then spending the spring term at
Purdue University, attempted to develop a theory of superconductivity based
on the self-energy of the electrons in the field of phonons. He heard about
the isotope effect in mid-M ay, shortly before he submitted his paper for
publication and was delighted to find very strong experimental confirmation
of his ideas. He used a Hamiltonian, now called the Frohlich Hamiltonian,
in which interactions between electrons and phonons are included but Cou
lomb interactions are omitted except as they can be included in the energies
of the individual electrons and phonons. Frohlich used a perturbation theory
approach and found an instability of the Fermi surface if the electron-phonon
interaction were sufficiently strong.
W hen I heard about the isotope effect in early M ay in a telephone call from
Serin, I attempted to revive my earlier theory of energy gaps at the Fermi
surface, with the gaps now arising from dynamic interactions with the phonons
rather than from small static lattice displacements. [14] I used a variational
method rather than a perturbation approach but the theory was also based on
the electron self-energy in the field of phonons. W hile we were very hopeful
at the time, it soon was found that both theories had grave difficulties, not
easy to overcome. [15] It became evident that nearly all of the self-energy is
included in the normal state and is little changed in the transition. A theory
302
involving a true many-body interaction between the electrons seemed to be
required to account for superconductivity. Schafroth [16] showed that starting
with the Frohlich Hamiltonian, one cannot derive the Meissner effect in any
order of perturbation theory. M igdal’s theory, [11] supposedly correct to
terms of order (m/M ) 1/2, gave no gap or instability at the Fermi surface and
no indication of superconductivity.
O f course Coulomb interactions really are present. The effective direct
Coulomb interaction between electrons is shielded by the other electrons and
the electrons also shield the ions involved in the vibrational motion. Pines and
I derived an effective electron-electron interaction starting from a Hamiltonian
in which phonon and Coulomb terms are included from the start. [1 7] As is the
case for the Frohlich Hamiltonian, the matrix element for scattering of a pair
of electrons near the Fermi surface from exchange of virtual phonons is
negative (attractive) if the energy difference between the electron states in
volved is less than the phonon energy. As discussed by Schrieffer, the attractive
nature of the interaction was a key factor in the development of the micro
scopic theory. In addition to the phonon induced interaction, there is the
repulsive screened Coulomb interaction, and the criterion for superconductivity
is that the attractive phonon interaction dominate the Coulomb interaction
for states near the Fermi surface. [18]
During the early 1950’s there was increasing evidence for an energy gap at
the Fermi surface. [19] Also very important was Pippard’s proposed non-local
modification [20] of the London electrodynamics which introduced a new length
the coherence distance, £0, into the theory. In 1955 I wrote a review article [17]
on the theory of superconductivity for the Handbuch der Physik, which was
published in 1956. The central theme of the article was the energy gap, and
it was shown that Pippard’s version of the electrodynamics would likely follow
from an energy gap model. Also included was a review of electron-phonon
interactions. It was pointed out that the evidence suggested that all phonons
are involved in the transition, not just the long wave length phonons, and
that their frequencies are changed very little in the normal-superconducting
transition. Thus one should be able to use the effective interaction between
electrons as a basis for a true many-body theory of the superconducting state.
Schrieffer and Cooper described in their talks how we were eventually able
to accomplish this goal.
3
G r e en ’s F unction M ethod for N ormal M etals
By use of Green’s function methods, Migdal [ 1 1 ] derived a solution of Frohlich’s
Hamiltonian, H = //ei+ #p h + 7Z ei-ph, for normal metals valid for abritrarily
strong coupling and which involves errors only of order (m\M )1/2. The Green’s
functions aredefined by thermal average of time ordered operators for the
electrons and phonons, respectively
G = —i< 7 y (l)y > + (2 )> (la)
D = —i< T ' 0 ( l ) 0 +(2 ) > (lb)
303
Here y)(r,t) is the wave field operator for electron quasi-particles and
0 (r,f) for the phonons, the symbols 1 and 2 represent the space-time points
(rn*i) and (r2,**) and the brackets represent thermal averages over an ensemble.
Fourier transforms of the Green’s functions for H0 = He\-\-Hv^ for non
interacting electrons and phonons are
°o(P) = ---------------------------------------------------------(
(On— Eo{k) + ldk
2 a)
= (-------— r v n
[V» — (Oo{q) + 1 0 Vn~\-(Oo{q) — lOj (2b)
where P = (k,con) and Q, = (q,vn) are four vectors, e0(k) is the bare electron
quasiparticle energy referred to the Fermi surface, co0(q) the bare phonon
frequency and a)n and vn the Matsubara frequencies
con = (2nJr \)n ik ^ T ; vn — 2nn\k^T (3)
for Fermi and Bose particles, respectively.
As a result of the electron-phonon interaction, i / ei-ph, both electron and
phonon energies are renormalized. The renormalized propagators, G and D ,
can be given by a sum over Feynman diagrams, each of which represents a
term in the perturbation expansion. W e shall use light lines to represent the
bare propagators, G0 and D0i heavy lines for the renormalized propagators,
G and D , straight lines for the electrons and curly lines for the phonons.
The electron-phonon interaction is described by the vertex
G(P+Q)
G(P)
which represents scattering of an electron or hole by emission or absorption
of a phonon or creation of an electron and hole by absorption of a phonon
by an electron in the Fermi sea. M igdal showed that renormalization of the
vertex represents only a small correction, of order (m /M )1/2, a result in accord
with the Born-Oppenheimer adiabatic-approximation. If terms of this order
are neglected, the electron and phonon self-energy corrections are given by
the lowest order diagrams provided that fully renormalized propagators are
used in these diagrams.
The electron self-energy E{P) in the Dyson equation:
„ — + — @------
G{P) = G0(P) +GJP)E(P)G{P) (4)
is given by the diagram
2. (5)
304
is given by
GIP+Q)
S>
G(P)
Since to order(m/M)1!* one can use an unrenormalized vertex function
a = ao, the Dyson equations form a closed system such that both Z(P) and
n (d ) can be determined. The phonon self-energy, n (QJ, gives only a small
renormalization of the phonon frequencies. As to the electrons, M igdal noted
that we are interested in states k very close to AF, s o that to a close approxima
tion Z(k,w) depends only on the frequency. For an isotropic system,
Z(k,c o) ~ Z(kv,(o) = E(oj) (7)
The renormalized electron quasi-particle energy, co*, is then given by a root
of
e(k) = co* = £oW +^(w jt) (8 )
In the thermal Green’s function formalism, one may make an analytic
continuation from the imaginary frequencies, cow, to the real co axis to determine
I{w ).
Although E((o) is small compared with the Fermi energy, E y , it changes
rapidly with energy and so can affect the density of states at the Fermi surface
and thus the low temperature electronic specific heat. The mass renormal
ization factor m*/m, at the Fermi surface may be expressed in terms of a par
ameter X :
m*lm = Z(kv) = 1 + X = (d£0/d*)F/(d£/d*)F (9)
In modern notation, the experession for X is
A= (10)
where F(co) is the density of phonon states in energy and a 2(co) is the square
of the electron-phonon coupling constant averaged over polarization directions
of the phonons. Note that X is always positive so that the Fermi surface is
stable if the lattice is stable. Values of X for various metals range from about
0.5 to 1.5. The parameter X corresponds roughly to the N(0) Fphonon of the
BCS theory.
4 N a m b u -E l ia s h b e r g T h e o r y f o r S u p e r c o n d u c t o r s
M igdal’s theory has important consequences that have been verified experi
mentally for normal metals, but gave no clue as to the origin of supercon
ductivity. Following the introduction of the BCS theory, Gor'kov showed
that pairing could be introduced through the anomalous Green’s function
F(P) = i < Ty)1y)i > , (1 1 )
Nambu showed that both types of Green’s functions can be conveniently
included with use of a spinor notation
305
M (r>0\
V= (12)
Vvl (*•»<)/
where y)^ and \pi are wave field operators for up and down spin electrons
and a matrix Green’s function with components
= - i < 7 > av » > (13)
Thus Gn and G22 are the single particle Green’s functions for up and down
spin particles and G12 ==G 21= F(P) is the anomalous Green’sfunction of
Gor'kov.
There are two self-energies, Z x and Z 2, defined by the matrix
(L\ EA
2 = 1 (14)
V* zJ
Eliashberg noted that one can describe superconductors to the same accuracy
as normal metals if one calculates the self-energies with the same diagrams that
M igdal used, but with N am bu matrix propagators in place of the usual
normal state Green’s functions. The matrix equation for G is
G = Go+GoEG (15)
The matrix equation for Z yields a pair of coupled integral equations for Z x
and Z 2. Again Z x and Z 2 depend mainly on the frequency and are essentially
independent of the momentum variables. Following Nam bu, [13] one may
define a renormalization factor £ s(g>) and a pair potential, A (co), for isotropic
systems through the equations:
oj^ s (oj) = co+ZAco) (16)
A(co) = Z 2(a>)IZ(o>)- (17)
Both Zs and A can be complex and include quasi-particle life-time effects.
Eliashberg derived coupled non-linear integral equations for Zs(^) and
A (cd) which involve the electron-phonon interaction in the function a2(co)F(co).
The Eliashberg equations have been used with great success to calculate the
properties of strongly coupled superconductors for which the frequency
dependence of Z and A is important. They reduce to the BCS theory and
to the nearly equivalent theory of Bogoliubov [21 ] based on the principle of
“compensation of dangerous diagrams” when the coupling is weak. By weak
coupling is meant that the significant phonon frequencies are very large
compared with A;b7*c, so that A{a>) can be regarded as a constant independent
of frequency in the important range of energies extending to at most a few
In weak coupling one may also neglect the difference in quasi-particle
energy renormalization and assume that Zs — Zn-
The first solutions of the Eliashberg equations were obtained by Morel and
Anderson [22] for an Einstein frequency spectrum. Coulomb interactions were
included, following Bogoliubov, by introducing a parameter /li* which re
normalizes the screened Coulomb interaction to the same energy range as the
phonon interaction, In weak coupling, N(0)V = %—ju*. They estimated A
from electronic specific heat data and /u* from the electron density and thus
the transition temperatures, 7"c, for a number of metals. Order-of-magnitude
306
agreement with experiment was found. Later work, based in large part on
tunneling data, has yielded precise information on the electron-phonon
interaction for both weak and strongly-coupled superconductors.
4
A nalysis of T unneling D ata
From the voltage dependence of the tunneling current between a normal
metal and a superconductor one can derive A(a>) and thus get direct infor
mation about the Green’s function for electrons in the superconductor. It
is possible to go further and derive empirically from tunneling data the
electron-phonon coupling, cl2((d)F(cjo), as a function of energy. That electron
tunneling should provide a powerful method for investigating the energy gap
in superconductors was suggested by I. Giaever, [23] and he first observed
the effect in the spring of 1960.
The principle of the method is illustrated in Fig. 1. At very low temperatures,
the derivative of the tunneling current with respect to voltage is proportional
to the density of states in energy in the superconductor. Thus the ratio of the
density of states in the metal in the superconducting phase, JVs, to that of the
same metal in the normal phase, JVn, at an energy eV above the Fermi surface
is given by
N s(eV) (d //d 7 )ns
An ~ (d //d F )nn
ftiv)
' 'n s
~ N SSM ~ V^ ^LA =Z
Tunneling from a normal metal into a superconductor
Fig. 1.
Schematic diagram illustrating tunneling from a normal metal into a superconductor near
T = 0°K . Shown in the lower part of the diagram is the uniform density of states in energy
of electrons in the normal metal, with the occupied states shifted by an energy eV from an
applied voltage V across the junction. The upper part of the diagram shows the density of
states in energy in the superconductor, with an energy gap 2 A . The effect of an increment of
voltage <5V giving an energy change dco is to allow tunneling from states in the range dco. Since
the tunneling probability is proportional to density of states Ng (to), the increment in current
<51 is proportional to Ns (co)<5V.
307
ENERGY (IN UNITS OF € )
Fig. 2.
Conductance of a Pb-Mg junction as a function of applied voltage (from reference 24).
E N E R G Y (m e V )
Fig. 4.
Real and imaginary parts of A (to) = ^i(w) -fLd2(a>) versus energy for Pb. (After McMillan
& Rowell).
310
i— i------ 1— i-----1— r— r
E N E R G Y ( M I L L IE L E C T R O N V O L T S )
O
C o m p a riso n of a F ( a ) and F(cu) for
P b (a ft e r M c M illa n and R o w e ll)
Fig. 5.
Comparison of azF for Pb derived from tunneling data with phonon density of states from
neutron scattering data of Stedman et al. [8]
In Figs. 5, 6 , 7, and 8 are shown other examples of a 2(co)F(co) derived from
tunneling data for Pb, In, [31] La, [32] and N b 3Sn. [33] In all cases the
results are completely consistent with the phonon mechanism. Coulomb
interactions play only a minor role, with /i* varying only slowly from one metal
to another, and generally in the range 0 . 1 — 0 2 .
311
a F ( oj) for indium.
Fig. 6.
a*F for In (after M cMillan and Rowell).
312
T
Fig. 7.
a*F for La (after Lou and Tomasch).
transition metals, a rare earth, and various alloys and compounds. Except
possibly for the metallic form of hydrogen, [35] which is presumed to exist
at very high pressures, it is unlikely that the phonon mechanism will yield
substantially higher transition temperatures than the present maximum of
about 21 K for a compound of Nb, Al and Ge.
Otjier mechanisms have been suggested for obtaining higher transition
temperatures. One of these is to get an effective attractive interaction between
electrons from exchange of virtual excitons, or electron-hole pairs. This re
quires a semiconductor in close proximity to the metal in a layer or sandwich
structure. At present, one can not say whether or not such structures are
feasible and in no case has the exciton mechanism been shown to exist. As
Ginzburg has emphasized, this problem (as well as other proposed mechanisms)
deserves study until a definite answer can be found. [36]
The pairing theory has had wide application to Fermi systems other than
electrons in metals. For example, the theory has been used to account for
313
0.6
0.5
0.4 o
II
CM CM
0.2
0.1
0 10 20 30
E (mV)
2
a F for N b 3S n ( a f t e r L.Y.L. Sh en)
Fig. 8.
arF for Nb3Sn (after Y. L. Y. Shen).
R eferences
1. Shoenberg, D. Superconductivity, Cambridge Univ. Press, Cambridge (1938). Second
edition, 1951.
2. London, F. Proc. Roy. Soc. (London) 152A, 24 (1935).
3. Bardeen, J. Phys. Rev. 59, 928A (1941).
4. Reynolds, C.A., Serin, B. Wright W. H. and Nesbitt, L. B. Phys. Rev. 78, 487 (1950).
5. Maxwell, E., Phys. Rev. 78, 477 (1950).
6. Frohlich, H., Phys. Rev. 79, 845 (1950); Proc. Roy. Soc. (London) Ser. A 213, 291 (1952).
7. London, F., Superfluids, New York, John Wiley and Sons, 1950.
8. For recent review articles with references, see the chapters by D. J. Scalapino and by
W. L. M cMillan and J. M. Rowell in Superconductivity, R. D. Parks, ed., New York,
Marcel Bekker, Inc., 1969, Vol. 1. An excellent reference for the theory and earlier
experimental work is J. R. Schrieffer, Superconductivity, New York, W. A. Benjamin,
Inc., 1964. The present lecture is based in part on a chapter by the author in Cooperative
Phenomena, H. Haken and M. Wagner, eds. to be published by Springer.
314
9. Gor'kov, L. P., Zh. Eksper i. teor. Fiz. 34, 735 (1958). (English transl. Soviet Phys. —
JETP 7, 505 (1958)).
10. Kadanoff L. P. and Martin, P. G. Phys. Rev. 124, 670 (1961).
11. Migdal, A. B., Zh. Eksper i. teor. Fiz. 34, 1438 (1958). (English transl. Soviet Phys. —
JETP 7, 996 (1958)).
12. Eliashberg, G. M., Zh. Eksper i. teor. Fiz. 38, 966 (1960). Soviet Phys. — JETP 11,
696 (1960).
13. Nambu, Y., Phys. Rev. 117, 648 (1960).
14. Bardeen, J., Phys. Rev. 79, 167 (1950); 80, 567(1950); 81 829 (1951).
15. Bardeen, J., Rev. Mod. Phys. 23, 261 (1951).
16. Schafroth, M. R., Helv. Phys. Acta 24, 645 (1951); Nuovo Cimento,9, 291 (1952).
17. For a review see Bardeen, J., Encyclopedia of Physics, S. Flugge, ed., Berlin, Springer-
Verlag, (1956) Vol. X V , p. 274.
18. Bardeen, J., L. N. Cooper and J. R. Schrieffer, Phys. Rev. 108, 1175 (1957).
19. For references, see the review article of M. A. Biondi, A. T. Forrester, M. B. Garfunkel
and C. B. Satterthwaite, Rev. Mod. Phys. 30, 1109 (1958).
20. Pippard, A. B., Proc. Roy. Soc. (London) A216, 547 (1954).
21. See N. N. Bogoliubov, V. V. Tolmachev and D. V. Shirkov, A New Method in the
Theory of Superconductivity, New York, Consultants Bureau, Inc., 1959.
22. Morel P. and Anderson, P. W., Phys. Rev. 125, 1263 (1962).
23. Giaever, I., Phys. Rev. Letters, 5, 147; 5, 464 (1960).
24. Giaever, I., Hart H. R., and Megerle K., Phys. Rev. 126, 941 (1962).
25. Rowell, J.M ., Anderson P. W. and Thomas D. E., Phys. Rev. Letters, 10, 334 (1963).
26. Culler, G .J., Fried, B. D., Huff, R. W. and Schrieffer, J. R ., Phys. Rev. Letters 8, 339
(1962).
27. Schrieffer, J. R., Scalapino, D .J. and Wilkins, J. W., Phys. Rev. Letters 10, 336 (1963);
D. J. Scalapino, J. R. Schrieffer, a n d j. W. Wilkins, Phys. Rev. 148, 263 (1966).
28. Scalapino, D. J., Wada, Y. and Swihart, J. C., Phys. Rev. Letters, 1 4,102 (1965); 1 4,106
(1965).
29. Eliashberg, G. M., Zh. Eksper i. teor. Fiz. 43, 1005 (1962). English transl. Soviet Phys. —
JETP 16, 780 (1963).
30. Bardeen, J. and Stephen, M., Phys. Rev. 136, A1485 (1964).
31. McMillan, W. L. and Rowell, J. M. in Reference 8.
32. Lou, L. F. and Tomasch, W. J., Phys. Rev. Lett. 29,858 (1972).
33. Shen, L. Y. L., Phys. Rev. Lett. 29, 1082 (1972).
34. Carbotte, J. P., Superconductivity, P. R. Wallace, ed., New York, Gordon and Breach,
1969, Vol. 1, p. 491; J. P. Carbotte and R. C. Dynes, Phys. Rev. 172, 476 (1968);
C. R. Leavens and J. P. Carbotte, Can. Journ. Phys. 49, 724 (1971).
35. Ashcroft N. W., Phys. Rev. Letters, 21, 1748 (1968).
36. See V. L. Ginzburg, “The Problem of High Temperature Superconductivity,” Annual
Review of Materials Science, Vol. 2, p. 663 (1972).
g7. Osheroff D. D., Gully W. J., Richardson R. C. and Lee, D. M. Phys. Rev. Lett. 29, 1621
(1972).
315
NOTES AND REFERENCES
157. (a) T. Kinsel et al., Rev. Mod. Phys., 36, 105 (1964).
(b) L. Dubeck et al., Rev. Mod. Phys., 36, 110 (1964).
158. (a) T. G. Berlincourt, Rev. Mod. Phys., 36, 19 (1964).
(b) P. W. Anderson and Y. B. Kim, Rev. Mod. Phys., 36,
39 (1964).
159. G. Chanin, E. A. Lynton, and B. Serin, Phys. Rev., 114,
719 (1959).
160. (a) B. T. Matthias et al., Rev. Mod. Phys., 36, 155 (1964).
(b) B. T. Matthias, T. H. Geballe, and V. B. Compton,
Rev. Mod. Phys., 35, 1 (1963).
161. (a) T. Tsuneto, Progr. Theoret. Phys. (Kyoto), 28, 857 (1962).
(b) P. G. de Gennes et al., Phys. Condensed Matter, 1, 176
(1963).
(c) D. Markowitz and L. P. Kadanoff, Phys. Rev., 131, 563
(1963).
(d) L. Gruenberg (to be published).
162. H. Suhl and B. T. Matthias, Phys. Rev., 114, 977 (1959).
163. W. Baltenspenger, Rev. Mod. Phys., 36, 157 (1964).
164. J. Bardeen, Rev. Mod. Phys., 34, 667 (1962); D. H. Douglass,
Rev. Mod. Phys., 36, 316 (1964).
165. P. G. de Gennes, Rev. Mod. Phys., 36, 225 (1964).
166. A. Bohr, B. R. Mottelson, and D. Pines, Phys. Rev., 110, 936
(1958).
167. (a) A. Bohr and B. R. Mottelson (book to be published).
(b) M. Baranger, Cargese Lectures 1962 in Theoretical
Physics, M. Levy (ed.), Benjamin, New York, 1963.
168. Y. Nambu and G. Jona-Lasinio, Phys. Rev., 122, 345
(1961); 124, 246 (1961).
169. J. C. Fisher, Phys. Rev., 129, 1414 (1963).
170. K. A. Brueckner, T. Soda, P. W. Anderson, and P. Morel,
Phys. Rev., 118, 1442 (1960).
171. W. A. Little, Rev. Mod. Phys., 36, 264 (1964).
172. (a) A. A. Abrikosov and L. P. Gorkov, J . Exptl. Theoret.
Phys. (U SSR), 39, 1781 (1960); translated in Soviet Phys.
J E T P , 12, 1243 (1961).
(b) F. Reif and M. A. Woolf, Phys. Rev. Letters, 9, 315
(1962).
(c) J. C. Phillips, Phys. Rev. Letters, 10, 96 (1963).
(d) H. Suhl and D. R. Fredkin, Phys. Rev. Letters, 10, 131,
268 (1963).
INDEX