Sie sind auf Seite 1von 105

GO

a
w
o

SUPER-

CONDUCTIVITY
3

Ernest A. Lynton

METHUENS MONOGRAPHS ON
PHYSICAL SUBJECTS

Superconductivity

ERNEST

A.

LYNTON

Although the fascinating phenomenon of


superconductivity has been known for fifty
years, it is largely through the concentrated
experimental and theoretical work of the last
decade that a basic (though at present very
incomplete) understanding of the effect has
been reached. This monograph is a largely
descriptive introduction to superconductivity, requiring little more than an undergraduate physics background. It is written to
serve two functions ; first as a stepping stone
towards more intensive study for those who
intend to work in the field of research and
development of superconductivity and its
applications and, secondly, as a basic refer-

ence on the present state of the subject of


superconductivity.

The book

contains a description of the

principal characteristics of a superconductor,

together with a detailed discussion of the most


useful phenomenological models which have

been applied to superconductors. The second

monograph describes the fundamental microscopic properties in terms of the


theory of Bardeen, Cooper and Schrieffer. It
is shown how remarkably successful this
theory has been in explaining the behaviour
of an idealized superconductor. There is a
chapter on superconducting devices, a subject index and a bibliography of more than
330 books and articles.
part of the

SECOND EDITION

LONDON
METHUEN & CO. LTD
NEW YORK
JOHN WILEY & SONS INC

methuen's monographs

on physical subjects

General Editor: B. L.

WORSNOP,

b.sc, ph.d.

SUPERCONDUCTIVITY

Superconductivity

E. A. Lynton
Professor of Physics
Rutgers, The State University
New Brunswick, N.J., U.S.A.

Powder patterns of the intermediate stale, showing thesbrinking of the superconducting (dark) regions as /; takes on the
values (left to right, top to bottom) 0, 008, 027, 053, 0-79, and
0-90.
(After Faber, 1958. Reproduced by kind permission of the
Royal Society and the author.) Proc. Roy. Soc. A248 464,
plate 25.

LONDON: METHUEN & CO LTD


NEW YORK: JOHN WILEY & SONS INC

First published in 1962

Second edition 1964

1962 and 1964 by E. A. Lynton


Printed

in

Great Britain by

Spottiswoode, Ballantyne

London

&

& Co Ltd

Colchester

Catalogue No. Methuen 12/4081/66


2-1

CHRJS.
i

For Carla

Acknowledgements

Contents

This book has grown, beyond recognition, from a set of lecture notes
written and used during my stay at the Institut Fourier of the Univerof Grenoble in 1959-60.

sity

hosts, Professors Neel

and pleasant year.

should like once again to thank

and Weil and Dr Goodman, for a stimulating

Basic Characteristics

am very grateful to a large number of people who

have helped me with written or oral comments, with news of their unpublished work, with preprints, and with copies of graphs. In particular I thank Drs Coles, Collins, Cooper, Douglass, Faber, Garfunkel,

Goodman, Masuda,

Olsen, Pippard, Schrieffer, Shapiro,

Swihart, Tinkham, Toxen, and

Waldram.

My colleagues Lindenfeld,

McLean, and Weiss provided much helpful discussion. Above all my


gratitude is due to Bernard Serin, from whose guidance and friendship
I

all

Perfect conductivity

and the critical magnetic field

Superconducting elements and compounds

1.3

1.4

The Meissner effect


The specific heat

1.5

Theoretical treatments

9
11

Phenomcnological Thermodynamic Treatment

13

16

the time to read the entire

and suggested many improvements, not


have been wise enough to incorporate.

2.2
2.3

Interrelation

many

years.

He found

September 1961

E. A.

2.4

Preface to the Second Edition

The Gorter-Casimir

This edition contains revisions and additions which bring the monograph essentially up to date, as of the end of June 1964. The treatment

3.3

Trapped

of superconductors of the second kind has been considerably amplia discussion of the Josephson effect has been added, and a num-

3.4
3.5

The perfect conductor


The London equations

3.6

Quantized flux

incor-

Dr

Nozieres, for his valuable

lation, edited with

many

unfortunately reached

comments and

Russian transilluminating footnotes by Dr. Gor'kov,

me too

late for these

help.

comments

to be included

in the present edition.

IV

E. A.

LYNTON

22

and the intermediate

state

4.1

4.2
4.3

23

26

flux

27
for a superconductor

28

32

The Pippard Non-local Theory

4.4

August 1964

19

22

Influence of geometry

fied,

model

Static Field Description

3.2

porated in the excellent French translation of Mme Nozieres, which


was published early this year. I am very grateful to her, as well as to

thermal
17

two-fluid

Perfect diamagnetism

Many of these were also

magnetic and

between

3.1

ber of other changes have been made.

13
effects

properties

LYNTON

1.2

The phase transition


Thermodynamics of mechanical

draft of the manuscript

of which

2.1

have profited for

first

page

Introduction

my

The penetration depth, A


The dependence of A on temperature and
The range of coherence
The Pippard non-local relations

The Ginzburg-Landau Phenomcnological Theory

vi
vii

34
34
field

36

40
41

48

VI

Contents

Contents

Vlll

The Surface Energy


The
The

page 55

,.

IX

Microscopic Theory of Superconductivity

page 116

55

11.1

Introduction

59

11.2

6.3

and the range of coherence


surface energy and the intermediate state
Phase nucleation and propagation

61

.3

6.4

Supercooling in ideal specimens

65

1 1

.4

6.5

Superconductors of the second kind

67

11.5

6.6

The mixed

71

The
The
The
The
The

6.7

Surface Superconductivity

74

11.7

Experimental verification of predicted thermal

75

11.8

The

75

11.10 Electromagnetic properties

6.1

6.2

surface energy

state or

Shubnikov phase

1.6

116

electron-phonon interaction

117

Cooper

pairs

118

ground

state energy

energy gap at

120

0K

superconductor at

126
finite

temperatures

129

properties

vn Low Frequency Magnetic Behaviour of Small Specimens


7.1

Vffl

Increase in the critical field

7.2

High

threads and superconducting magnets

77

7.3

Variation of the order parameter and the energy


gap with magnetic field

78

The Isotope
8.1

8.2

field

Effect

XII

81

12.1

Introduction

12.2

Dilute solid solutions with non-magnetic im-

12.3

Compounds

83

Thermal Conductivity

87

9.2

Electronic conduction

9.3

Lattice conduction

9.4

Thermal conductivity in the intermediate

temperature thermal conductivity

139
141
141

141

Superconducting Devices

145

150
153

153

89

13.2

Superconducting magnets

154

Superconducting computer elements

157

3.3

94

95

10.2

The

96

10.3

Electromagnetic absorption in the far infrared

10.4

Microwave absorption

100

10.5

Nuclear spin relaxation

104

10.6

The tunnel

10.7

Far infrared transmission through thin

10.8

The Ferrell-Glover sum

specific heat

98

106
films

Superimposed metals

Research devices

95

effect

XIII

2.4

with magnetic impurities

13.1

Introduction

rule

136

87

93
state

The Energy Gap


10.1

134

81

Precise threshold field measurements

Low

Coherence properties and ultrasonic attenuation

purities

Discovery and theoretical considerations

9.1

specific heat

Superconducting Alloys and Compounds

IX

1.9

127

109
1

14

Bibliography

165

Index

183

Introduction
Although the fascinating phenomenon of superconductivity has been
known for fifty years, it is largely through the concentrated experimental and theoretical work of the past decade that a basic (though
as yet very incomplete) understanding of the effect has been reached.

Far from being an oddity of little physical

interest it has been shown


phenomenon of basic importance and with close
a number of fields. At the present time one important

to be a co-operative

analogies in

period in the development of the subject has been completed, and the
next is already well under way, with much effort in theory and experi-

ment to carry our understanding from the general to the particular,


from the idealized superconductor to the specific metal. Somewhat
coincidentally, there

now

also

is

great interest in possible practical

applications of superconductivity.

This

monograph

is

a largely descriptive introduction to super-

more than an undergraduate physics background, and written to serve two functions. It can be a first survey and
a stepping stone toward more intensive study for those who intend to

conductivity, requiring no

become actively engaged


tivity,

in the further development of superconducbe it in basic research or in technical applications. Such readers

from the extensive bibliography, listing more than 450


At the same time the book is sufficiently complete
description both of experimental details and of theoretical

will benefit

books and
in its

articles.

approaches to be a basic reference for those who wish to be acquainted


with the present state of superconductivity. It will enable them to
follow further developments as they appear in the scientific and
technical literature.

The contents of the book can be grouped into a number of sections


which treat the subject of superconductivity in successive layers with
increasing resolution of detail. The first three chapters introduce the
reader to the principal characteristics of bulk superconductors, and
treat these in

terms of the basic phenomenological models of London


and of Gorter-Casimir. With this section the reader thus acquires a
broad outline and a general understanding of the thermodynamic and
1

Superconductivity

the static electromagnetic behaviour of idealized, bulk superconductors. The treatment of the subject is then pursued in greater detail
along two essentially parallel directions. In the section comprising
Chapters IV-VII are discussed those aspects of the behaviour of

CHAPTER

Basic Characteristics

superconductors which lead to the non-local treatments of Pippard

and of Ginzburg and Landau. These more sophisticated phenomenological models account for an interphase surface energy, in terms of
which the later chapters of this section describe the intermediate state,
phase nucleation, propagation, and supercooling, superconductors
of the second kind, and the magnetic behaviour of specimens of
small dimensions. Chapters VIII-X can be read without a study of
the preceding section (IV-VII)

and

characteristics of a superconductor

describe in

much

detail those

which during the past decade

Perfect conductivity and critical magnetic field


behaviour
of electrical resistivity was among the first problems
The
investigated by Kamerlingh Onnes after he had achieved the liquefaction of helium. In 1911, measuring the resistance of a mercury
sample as a function of temperature, he found that at about 4K the
resistance falls abruptly to a value which Onnes' best efforts could not
1.1.

distinguish

zero. This extraordinary

from

phenomemon he

called

have indicated the microscopic nature of superconductivity, and


have led to the theory of Bardeen, Cooper, and Schrieffer. The
fundamental aspects of this theory are presented with a minimum of

superconductivity,

mathematics.

current will be induced which attempts to maintain the magnetic flux

The book closes with a chapter on the behaviour of alloys and compounds, and with one on superconducting devices.
In describing the principal empirical characteristics of supercon-

temperature,

When

Tc

and the temperature at which


(Kamerlingh Onnes, 1913).

a metallic ring

through the ring

at

have

tried to include only the

tinent experiments, or in the description of superconducting be-

to emphasize

has been slighted.

My

what aspects of the latter


whose work
The choice was not a judgement of the scientific
detail.

may appear

critical

field,

a constant value. For a body of resistance

R and

self-inductance L, this induced current will decay as

(LI)

7(0)exp(-i?//L).

key experiments through

which the phenomenon in question was established, as well as more


recent work which gives the most detailed or the most precise information. It is both unnecessary and impossible in a monograph of this
small size to be encyclopaedic either in the enumeration of all perhaviour in minute

appears the

exposed to a changing magnetic

is

/(/)

ductors

it

selection of

arbitrary, especially to those

value of such work, but rather of its didactic usefulness in illuminating


the elementary characteristics of superconductors.

/(/)

can be measured with great precision, for example, by observing

the torque exerted by the ring


carries

known

upon another, concentric one which

current. This allows the detection of

much

smaller

any potentiometric method. A long series of such


measurements on superconducting rings and coils by Kamerlingh
Onnes and Tuyn (1924), Grassman (1936), and others recently culminated in an experiment by Collins (1956), in which a superconducting ring carrying an induced current was kept below Tc for about two
and a half years. The absence of any detectable decay of the current
-21
during this period allowed Collins to place an upper limit of 10
ohm-cm on the resistivity of the superconductor.! This can be com-9
pared to the value of 10
ohm-cm for the low temperature resistivity
resistance than

of the purest copper.

There

is,

therefore,

little

doubt that a superconductor

is

indeed a

23
ohm-cm
t Quinn and Ittner (1962) have lowered this upper limit to 10"
by looking for the time decay of a current circulating in a thin film tube.

Basic characteristics

Superconductivity

perfect conductor, in the interior of which


field vanishes.

any slowly varying electric

A current induced in a superconducting ring will persist

indefinitely without dissipation.

Below Tc the superconducting behaviour can be quenched and


normal conductivity restored by the application of an external mag,

netic field. This field,


field,

and, as

shown

c , is

thick so that surface effects can be ignored, the critical current is that
which creates at the surface of the specimen a field equal to
c.
Smaller samples remain superconducting with much higher currents

than those calculated from this criterion, which

1 ,

it

1.2.

varies approximately as

Superconducting elements and compounds


I lists all

characteristic
(1.2)

investigators,

Alekseevskii

presently

and

Tc

known superconducting elements and their


In addition there have been found by many

in particular by Matthias and co-workers, by


and co-workers, and by Zhdanov and Zhuravlev (see

Table
Element

Normal

119

Cadmium

0-56

Gallium
Indium

3-407

Temperature T
Fig.

= Hc at T= 0K. It is convenient to introduce reduced co/

T/Tc and
,

h(t)

H (T)/H
C

2
ft l-/

The actual temperature variation of h


by a polynomial in which the

5-95

Lead
Mercury-a

718
4153

Mercury-/?

3-95

Molybdenum
Niobium

9-46

Osmium

0-7

Rhenium

1-70
0-49

in

terms of which
(I.2a)

is

more accurately represented

coefficient

of the

2
t

term

differs

from

unity by a few per cent.

The superconductivity of a wire or film carrying a current can be


quenched when this reaches a critical value. For specimens sufficiently

Technetium
Thallium

Thorium
Tin
Titanium
Tungsten

(gauss)

99
30
51
283

~20

~5

Lanthanum-/*

Ruthenium
Tantalum

109

014

Lanthanum-a

TC (K)

Aluminium

Iridium

ordinates

called Silsbee's

called the critical or threshold magnetic

in Figure

H *H [l-(lJ],

is

rule (Silsbee, 1916).

Table

where

1600
803
411
340

10

4-482

1944
65-82
201
66
830

11-2

2-39
1-37

171

162
306
100

3-722
0-40

~001

Uranium-a
Uranium-y

0-6

Vanadium

5-30
0-92

1310
53

0-75

47

Zinc
Zirconium
(cf.

~2000

1-80

Roberts (1963) for most references)

Superconductivity

Matthias, 1957; Roberts, 1961), a very large

Basic characteristics

number of

alloys

and

compounds which also become superconducting. Some of these


compounds consist of metals, only one of which by itself becomes
superconducting, some have constituents of which neither by itself is
superconducting, and some even are semiconductors. The possibility
of superconductivity in semiconductors and semimetals has been
discussed by M. L. Cohen (1964), and both GeTe (Hein et al., 1964)
and SrTi0 3 (Schooley et al., 1964) have been found to be superconducting at very low temperatures.

The Meissncr

effect, and the reversibility of the S.C. transition


conductor were placed in an external magnetic field, no
magnetic flux could penetrate the specimen. Induced surface currents
would maintain the internal flux, and would persist indefinitely. By

1.3.

If a perfect

same token, if a normal conductor were in an external field before


became perfectly conducting, the internal flux would be locked in

the
it

by induced persistent currents even if the external field were removed.

o o

A:He=0,

B%0.

C:0<H e<H o

B:H e=0

T>TC

T<TC

T<TC

T<TC

(b)

(a.)

(d)

(c)

Fig. 3

10

No. valence electrons/atom

Fio.2

The

C:0<Hp<H
e n c.

temperatures of superconductors range from very low


values up to 181K for Nb 3 Sn (Matthias et al., 1954). Matthias
(1957) has pointed out a number of regularities in the appearance of
critical

superconductivity and in the values of


the following

Tc

Superconductivity has been observed only for metallic substances for which the number of valence electrons
lies between

8.

(2) In all cases involving transition metals, the variation

H e =0,

T<TC

(d)
Fig. 4

Because of this, the transition of a merely perfectly conducting specimen from the normal to the superconducting state would not be
reversible,

and the

final state

of the specimen would depend on the

path of the transition.

of Tc with

number of valence electrons shows sharp maxima for Z = 3, 5,


and 7, as shown in Figure 2.
(3) For a given value of Z, certain crystal structures seem more
favourable than others, and in addition Tc increases with a high power
of the atomic volume and inversely as the atomic mass.
the

B:

the principal of which are

(1)

about 2 and

T<TC

As an example, Figures 3 and 4 show the flux configuration for a


perfectly conducting sphere taken from point A in Figure 1 to
point

by the different paths

ABC and ADC,

respectively.

distribution at C, as well as that at B, depends

The

final field

on whether one proceeded via Bov via D, and the irreversibility of the transition
is evident.
Careful measurements of the field distribution around
a spherical

Superconductivity

Basic characteristics

specimen by Meissner and Ochsenfeld (1933), however, indicated that

regardless of the path of transition the situation at point


is always
that shown in Figure 3c the magnetic flux is expelled from the interior
of the superconductor and the magnetic induction B vanishes. This is
:

called the Meissner effect,

tion

is

and shows that the superconducting transi-

connectivity of the body, so that

one has a superconductor with a


and only the hole may be
The magnetic properties of such a superif

hole, the Meissner effect occurs in the metal

threaded by magnetic

flux.

conducting ring are thus essentially determined by the relative size of


the diameter of the ring to the diameter of the hole.

reversible.

Figure 5 illustrates this by showing B vs. e curves both for a perfect


conductor and for a superconductor, taking the case of long cylindrical specimens with axes parallel to the applied field.
e is a uniform,

1.4.

The

The

specific heat

specific heat

of a superconductor consists,

metal, of the contribution of the electrons

(Cg). For a normal metal

>'
*^n

perfect conductor

/>

He

is

the

at

of a normal

like that

(Ce) and

that of the lattice

low temperatures

Wntt,gn = yT+A(T/0) 3
,

(1.3)

Sommerfeld constant, which is proportional to the density of


Fermi surface, is the Debye temperature, and

electronic states at the

cO

a numerical constant for

all metals. Experimentally the two contribe separated by plotting CJTvs. T 2 so that the
slope of the resulting curve is A/0 3 and the intercept is y.
In the superconducting phase

butions to

Cn can

superconductor

"a

C|

= Ces +Cgs

Cs (H=0) and C (H> c ) for tin as


measured by Corak and Satterthwaite (1954), displaying the characteristic features ofa sharp discontinuity in C,of the order
of 2y7c at Tc
and a rapid decrease of Cs to values below Cn varying about as T 3
It is customary to attribute the difference
between Cs and C entirely
to changes in C on the assumption that C is the same in
e
both phases.
g
This seems reasonable in view of the electronic nature of the superconducting phenomenon, and is supported by the absence of any
Figure 6 shows values of both

Applied Field He

Fro. 5

external

field.

In increasing

field

H =H

both specimens have

5=0

until

e
c , when they become normal and B =
e If the field is now
again decreased, the induction inside the perfect conductor is kept at
its threshold value B =
c by surface currents, and in zero field the
.

specimen is left with a net magnetic moment, as is illustrated in Figure


4d.

The superconductor, however, expels the flux at the transition and

returns reversibly to

its initial

state with

B=

for

< He < Hc

The vanishing of the magnetic induction, corresponding


pulsion of the magnetic flux,

is

to the ex-

the basic characteristic of all ideal

superconducting material with dimensions large compared to a basic


length which will be mentioned later. It is quite independent of the

observable change in the lattice parameters (Keesom and Kamerlingh


Onnes, 1924), and by the detection of only minimal changes in the
elastic properties (see, for instance, Alers

and Waldorf,

1961).

On this

assumption

^s

*-n

which allows one to determine


specific heat difference after

f-'es

Ces

^e

(1.4)

from measured values of the

Ce = y7"has been determined separately.

There has recently been some evidence that the lattice contributions
to the specific heat in the two
phases are not quite equal in the case of

Basic characteristics

Superconductivity

10

very low temperatures.


in

For

11

these elements possible small differences

Cg therefore do not much affect the validity of 1.4.


Ccs for tin calculated on the basis of 1.4 from the

Figure 7 displays

Figure

results in

6,

plotted logarithmically in units of l/yTc vs.

This shows that for l/t>2, one can represent

CJyTc =

A subsequent chapter
existence of a finite

in the

1//.

by the equation

aexp(-b/t).

will discuss that this is

gap

Ces

(1.5)

an indication of the

energy spectrum of the electrons

separating the ground state from the lowest excited state.

The number

of electrons thermally excited across this gap varies exponentially


it has become
apparent that such an energy gap determines the thermal properties
as well as the high frequency electromagnetic response of all super-

with the reciprocal of the temperature. In recent years

(deg.Kf)
Fig. 6

conductors, and that

it

must indeed be one of the principal features

of a microscopic explanation of superconductivity.


1.5.

Theoretical treatments

The macroscopic characteristics of a superconductor have been


subject of a number of phenomenological treatments of which
principal ones will be discussed in subsequent chapters. F.

the
the

and H.

London

(1935a, b) developed a model for the low frequency electromagnetic behaviour which is based on a point by point relation
between the current density and the vector potential associated with

a magnetic
electrons

field.

This implies wave functions of the superconducting


in the presence of such a field extend rigidly to

which even

the limits of the superconducting material

and then vanish abruptly.

A thermodynamic treatment and an associated two-fluid model based


1.0

1.5

2.0

2.5

3.0

3.5

on essentially equivalent simplifications were worked out by Gorter


and Casimir (1934a, b). These complementary theories provide highly

4.0

1/t

indium (Bryant and Keesom, 1960; O'Neal et al., 1964), so that 1.4
may not be exact for this element and possibly other superconductors
as well. Ferrell (1961) has suggested that this is due to a shift in the

phonon frequency spectrum. However, the superconducting elements for which reliable values of C exist are those with a relatively
high Debye temperature for which

Cg < Ce

and useful tools in the semi-quantitative analysis of many


problems involving superconductors. Their limitations become
apparent principally in situations in which size and surface effects are
successful

Fig. 7

in

both phases

down

to

important.

Pippard (1950, 1951) has shown that such effects become tractable
when one takes into account the finite coherence of the superconducting wave functions which is such as to allow them to vary only slowly
over a

finite distance.

This leads (Pippard, 1953) to a non-local

Superconductivity

12

and the vector


The equation has only

integral relation between the current density at a point

potential in a region surrounding the point.

been solved for a few special cases. In many instances, however, it


reduces to a modified version of the London equation, so that the
much simpler London formalism can then be used with the Pippard

CHAPTER

II

Phenomenological Thermodynamic Treatment

modifications (Tinkham, 1958).

Ginzburg and Landau (1950) have developed on a thermodynamic


basis

an

method of

alternate

treating the coherence of the super-

conducting wave functions. Their treatment

is

compatible with

Pippard's electromagnetic approach, and forms a highly useful com-

plement to

it.

A successful microscopic theory of superconductivity has recently


been developed by Bardeen, Cooper, and Schrieffer (1957). It is based
on the fact, established by Cooper (1956), that in the presence of an
attractive interaction the electrons in the neighbourhood of the

Fermi

surface condense into a state of lower energy in which each electron


is

paired with one of opposite momentum and spin. Bardeen, Cooper,

and

Schrieffer

(BCS) have been able to show that a finite energy gap

number of Cooper pairs


from the state with one pair less. This leads to the correct thermal
and electromagnetic properties to display superconductivity.
The attraction between electrons necessary to form Cooper pairs
can in principle be due to any suitable kind of interaction. The discovery (Maxwell, 1950; Reynolds et al., 1950) that for many superconducting elements the critical temperature depends on the isotopic
mass showed that for these substances the attractive interaction is one
between the electrons and the lattice. The BCS theory and its extensions have been worked out on this basis. However, the isotope effect

separates the state with the largest possible

is

The phase transition


Long before the determination of
2.1.

the reversibility of the superconducting transition by the discovery of the Meissner effect, attempts

had been made to apply thermodynamics to it by Keesom (1924), by


Rutgers (Ehrenfest, 1933), and in particular by Gorter (1933), who
virtually predicted the Meissner effect by pointing out that the success
of these early thermodynamic treatments strongly suggested the
reversibility

of the transition.

The discovery of the Meissner

effect finally enabled Gorter and


Casimir (1934a) to develop a full treatment of the superconducting
phase transition in a manner analogous to that of other phase transitions. They start with the fact that two phases are in equilibrium with

one another when their Gibbs free energies (G) are equal. The free
energy of a superconductor is most easily expressed by a diamagnetic
description developed in Chapter III, which attributes to the superconductor a magnetization
(H e) in the presence of an external

field

Then
V

G,(He ) = G/0)- j
For an

ellipsoid,

their

compounds

(see section 8.1).

Furthermore the

effect

pressure in transition metals does not correlate with the

temperatures as

it

) is

uniform, and

G s (H e ) = G s (0)- VJ M(He)dHe

of

(11.10

Debye

does in non-transition superconductors (Bucher

(H.l)

e.

He

apparently absent or considerably reduced in some transition metals

and

M(H

He

dvj M(He)dH

The last term

in this expression gives the

work done on

and Garland (1963a, b) have attributed these anomalies to the existence of overlapping bands in the
electronic energy spectrum at the Fermi surface. However, there is

by the magnetic

also a hypothesis that in transition metals the attractive interaction

parallel to the external field does the superconducting


into the normal one at a sharply defined value of
.

and Olsen,

1964).

Kondo

responsible for pairing

(1962)

may

be a magnetic one (Matthias, 1960).

field.

As

the magnetization

is

the specimen

diamagnetic, that

is,

negative, the field raises the energy of the superconducting specimen.


It will

be shown in Chapter III that only for a quasi-infinite cylinder

13

phase change

For

all

other

14

Phenomenological thermodynamic treatment

Superconductivity

is an intermediate state consisting of a mixture of normal


and superconducting regions. Even under these circumstances, however, any magnetic work is done solely on the superconducting portions, and for any shape of specimen this always equals, per unit

shapes, there

volume,
He

JM(H )dHe = -H 2 /Stt.

(II.2)

Thus one can

15

shows that the latent heat Q = T(Sn S s ) is zero


and is positive when c > 0. Thus there
isothermal superconducting-to-normal
absorption
of
heat
in
an
is an
transition, and a corresponding cooling of the specimen when this takes
place adiabatically. The resulting possibility of cooling by adiabatic
magnetization of a superconductor was suggested by Mendelssohn
(Mendelssohn and Moore, 1934) and has been used by Yaqub (1 960)
for low temperature specific heat measurements of tin.
Equation

II.5 also

at the transition in zero field,

write for any specimen

G,(/rc)

= (7,(0)+WJc2 /87r.

In the normal state the susceptibility

is

(H.3)

generally vanishingly small, so

that

G n (Hc) =

G(0).

Since the condition of equilibrium defining

H (T)
C

is

that

Gn (Hc) = G {H
S

C),

one has
01.4)

This is the basic equation of the thermodynamic treatment developed by Gorter and Casimir. As S = - (dG/dT) H differentiation
Pi
of 11.4 yields
,

Sn (0) - Sffl = - (VHJAtt) (dHJdT).


At T=

Tc

(II.5)

= 0, and Sn = Ss At any lower temperature, Hc > 0, and


shows that for < T< Tc dHJdT < 0. Hence

furthermore Figure

A further differentiation of II.4 yields,

the entropies of the two phases are equal at the critical temperature in
zero field; at any lower, finite temperature the entropy of the super-

conducting phase is lower than that of the normal one, indicating that
the former is the state of higher order. This ordering will later be
shown to follow from a condensation of electrons in momentum
space.

It

follows from Nernst's principle that

Sn = S s at T=

in this limit the slope of the threshold field curve

entropies of the two phases are also equal at

must pass through a maximum

at

0,

so that

must vanish. As the

T= Tc

their difference

some intermediate temperature.

Cs -Cn =

(VT/47r)[Hc (d

upon

H /dT

multiplication by T:

+ (dHJdT) 2).

(II.6)

Atr=ro ^ = 0,and
c

C,- Cn =

(*T/4tt)

{dHJdTfj^

Te

>

0,

(H.60

so that the thermodynamic treatment predicts the observed discontinuity in the specific heat. As the entropy difference between the
tvvo

Tc

phases passes through an extremum at

some temperature below

the specific heats of the two phases at that temperature

must be

16

Phenomenological thermodynamic treatment

Superconductivity

equal,

and

at even lower temperatures

course tend toward zero at

Cs is smaller than

C.

Both of

T= 0. The variation of Cs - C as a func-

tion of temperature, as well as that of Ss -S, are

shown

in Figure 8.

Thermodynamics of mechanical effects


The thermodynamic treatment developed thus far has ignored any
changes in the volume at the transition, as well as any dependence of
2.2.

17

There has been extensive experimental work on pressure effects on


by Swenson (1960) and sumrecently
Olsen
and
Rohrer
most
by
(1960). These latter
marized

the critical field. This has been reviewed

authors (1957) and, independently, also Cody (1958), have succeeded


in refining earlier work of Lazarev and Sudovstov (1949), and have

obtained for different superconducting elements empirical values of


the length

change of a long rod at the transition. (Andres et at., 1 962).

H on pressure as well as on temperature. In taking these into account

Differences in the behaviour of transition

one should begin by considering possible magnetostrictive field effects


on the volume in going from II. 1 to II. 1'. Ignoring this, however,
and noting (see Figure 1 1) that for the special case of a quasi-infinite

have been pointed out by Bucher and Olsen (1964).


The magnitudes of the several mechanical effects are exceedingly
8
9
small. Typical values for 8HJ8p are of the order of 10~ -10~

cylinder parallel to the external field the area under the magnetization

gauss/dyne-cm

curve up to any

field

value

H <H

c is

QABQ-Gjm =

equal to

one can write

j%tt,

(VJ*ir)Hl

01.7)

Differentiating this with respect top in order to obtain

V= (8GI8p) T H

and non-transition metals

-2

and the fractional length change of a long rod


-8
Using the above thermodynamic relations this
-7
yields a difference in the thermal expansion coefficient of about 10
per degree, and a fractional change in compressibility of one part
is

a few parts in 10

in 10

5
.

yields

V5 {He)- Vjm =

(H}l87r)(dVs !8p) T

(n.8)

Vn {Hc)-VM = *mv,EftT,p)l*n\
Vn {Hc)-Vs {0) = {H}l%ir)(dVJBp) T +{Vs HJAn)(dHcl*P)TComparing

II.9

with II.8 shows that the

side of the former

is

first

01.9)

superconductor

upon changing the field from zero to the critical value. It is the second
term which gives the actual volume change at the transition

V(HC )-VS (HC ) - (Vs Hc /47r)(dHc l8p) T

The

interrelation between magnetic

and thermal properties

magnetic and the thermal properties of a superconductor. Equation

term on the right-hand

just the magnetostriction of the

2.3.

One of the most remarkable features of the thermodynamic treatment


outlined in the preceding sections is the manner in which it links the

Similar differentiation of II.3 and II.4 leads to

(11.10)

This term exceeds the magnetostrictive one by more than an order of


magnitude. The derivatives of 11.10 with respect to T and to p yield

for example, indicates that quite independently of the detailed


shape of the magnetic threshold field curve, its negative slope indiII. 5,

cates that the superconducting

phase has a lower entropy than the


normal one. The quantitative verification of an equation such as II. 6',
called Rutgers' relation, provides the best available confirmation of
the basic reversibility of the superconducting transition. The following

taken from Mapother (1962), compares for a few particularly


favourable elements the specific heat discontinuity measured caloritable,

metrically, with

hold

its

field curves.

value calculated with II.6' from measured thres-

The agreement is seen

to be excellent:

expressions for the changes at the transition of the coefficient of

thermal expansion et=(l/V)(SV/dT) t and of the bulk modulus


K = - V{8pj8V). AtT= Tc,
= 0, this yields
c

-, = (U47T)(8Hc ldT)(dHc l8p\

(11.11)

Element
(millijoules/mole)

Indium
Tin

and

*- =

(K /47r)(dHc /8p)

2
.

(IU2)

Tantalum

9-75
10-6
41-5

9-62
10-56
41-6

18

Superconductivity

Phenomenological thermodynamic treatment

The relations between the thermal properties and the threshold field
curve of course also imply that

a specific temperature variation is


either assumed or empirically determined for one of the former, this
uniquely specifies the temperature variation of the latter. Kok (1 934),
for example, showed that if one substitutes into equation II.6 a parabolic variation of

c,

if

as given by equation

one obtains a cubic


It was mentioned in Chapter I that both
1.2,

temperature variation of Ccs


of these are only fair approximations to the actual temperature
dependence of these quantities, and that in fact the threshold field can
be represented more accurately by a polynomial which in reduced
.

co-ordinates has the form


h(t)

The

= 1-

a n t\

T=

(see

directly to

polynomial

to neglect

is

must vanish, as otherwise Ss S would not


equation II.5), and 2> = 1 to make //(l) = 0. If

in the lattice specific heat,

~^f=(MWHHT^-a

and one continues


follows that

it

...)x

argument: As shown by equation II.5, Sn > Ss and since S varies


linearly with T, Ss must approach zero with some power of T greater
than unity. Hence one can write
,

0,

Coc T l+X ,

so-called phenomenological two-fluid

tivity

have in

two-fluid

common two

The system

dependence of

Ces

is,

no matter what

CJT^Q

exhibiting superconductivity possesses an ordered

or condensed state, the total energy of which


order parameter. This parameter

is

T= Tc to unity at T= 0K, and can thus


which

is

characterized by an

generally taken to vary from zero

be taken to indicate that

finds itself in the superconducting

(2) The entire entropy of the system is due to the disorder of noncondensed individual excited particles, the behaviour of which is
taken to be similar to that of the equivalent particles in the normal
state.

In particular, two-fluid models

make

the conceptually useful

assumption that in the superconducting phase a fraction #" of the


conduction electrons are 'superconducting' electrons condensed into
an ordered state, while the remaining fraction 1 #" remain 'normal'.
artificiality

of this division cannot be overemphasized;

its

use-

fulness will presently appear.

the precise temperature

as T-+0.

Applying

to

The free energy per unit volume of the normal electrons continues
be the same as that of electrons in a normal metal, that is

where y
(l/27T)a 2 (Hl/Tl).

(11.15)

'

2
gn (T) = -\yT

this limit to

equation 11.14 thus yields

models of superconduc-

general assumptions:

'

It follows, therefore, that

(11.16)

model

The Gorter-Casimir

The

The

CJTozT*.

and

T-+0, dHc /dT-+0. One

state.

(11.14)

the two terms on the left-hand side, the second just equals the
Sommerfeld constant y. The first is subject to the following general

so that

recalling that as

2.4.

fraction of the total system

Of

T l+X ,x >

and

Both of these last equations are exact expressions which permit the
evaluation of the Sommerfeld constant from a detailed knowledge of
the threshold field curve. Mapother (1959, 1962) has carried out a
searching analysis of the extent to which magnetic and thermal data
can actually be correlated in practice without introducing excessive
errors due to extrapolation; Serin (1955) and Swenson (1962) have
also discussed the relation between the two types of data.

at
2
2t

x(2a2) + (-2<7 2 /-...)}.

o:

II.6,

2
y = -(l^Tr^Hl/T^ihd^rldr )^.

(1)

substituted into equation II.6,

any changes

equation

then obtains

this

equivalent expression results from applying the above argument

An

(11.13)

first coefficient a,

vanish at

19

is

(11.17)

the Sommerfeld constant. For the 'superconducting' elec-

trons g (T)
s

is

taken to be a condensation energy relative to the normal

CHRIST'S

COLLEGE

LIBRARY

20

Phenometwlogical thermodynamic treatment

Superconductivity

phase, and the considerations of the


this to

g s (T)

The

first

section of this chapter

show

be

#" of

-HllSrr.

The value of a must be chosen so as to give a reasonable fit to


mental data. With a = one obtains

experi-

QOT.n =

(11.270

(H.1J

volume of the superconducting phase


electrons and 1 of n electrons is

total free energy per unit

containing a fraction

iT(J)
s

a(\

- fT)gn {T) + b(iT)8s (T).

(11.19)

The simplest choice of a(l-iT) = 1-iT; b(iT)=ir, makes


G&if, T) a linear function of 1P, so that the equilibrium condition
(3C/air)r = can be satisfied for only one value of Tat which iT can
assume any value between and l This would mean that for any value
of
the normal and superconducting phases can be in equilibrium
at only that one temperature, which is not the case. Thus it is necessary
to choose a(l - iT) and b(iT) with more care. Gorter and Casimir
.

(1934b) chose

3yTc t\

SS(T) = yTc t\

'

therefore

G (ir, T) =

21

= l-/4

(11.260
(11.250

and

y = (WirKHlfT?).
Here again
which

is

is

(11.230

the cubic temperature variation of the specific heat

only an approximation. Clearly no value of a will change

equation 11.27 into an exponential expression.

comparison of

=
representation of h(t) to a

equation 11.23' with equation 11.15 also shows that the choice a

makes a 2 =\, and reduces the polynomial

parabolic form. This not only indicates once again the interrelation

between the magnetic and thermal properties, but also points up that
a(l

- #") =

- ir)\

(l

b{iT)

Hr,

(11.20)

the Gorter-Casimir
tively.

so that

G3 (ir,T) = -\{\-Hr)*yT 1 -inill%iT.

(11.21)

Within

model can be used

this limitation,

at best only semi-quantita-

however, the concept of the two inter-

penetrating 'fluids' of condensed

and uncondensed electrons

is

very

useful in obtaining a semi-quantitative understanding of many super-

Applying the equilibrium condition yields


a(l

- *0"~ m
'

conducting phenomena, and will be used repeatedly in subsequent

2
HllA-n yT

(IL22)

chapters.

There have been a number of attempts

which at

Tc

with

1T - 0,

p. 280) to improve
model so as to yield
more nearly the correct exponential variation of Ces and the corresponding non-parabolic dependence of h(t). These modifications have
either tried different functional forms for a(l - 1T) and b(iT) in

reduces to

2
y = (WttxHHIIT .).

Substituting 11.23 back into 11.22:

(\-1T)-

(TJT)

(11.24)

Oi.:

Putting this back into 11.21 and differentiating to obtain other


thermal quantities yields

s s (ir,T) = yT(\-iry = y rc r (1+0[)/(1 a)

Some of these

variations do yield considerably better equations for


and magnetic superconducting properties. However, the
principal virtue of a two-fluid model is to provide a conceptual tool
of primarily qualitative nature, and the various suggested improvements rarely add much to the basic physical picture of the two groups
the thermal

Hr = |_^/(i~0

f electrons.
,

and
[(l

[5],

equation TJ.19, or have introduced additional adjustable parameters.

r2

so that

Cs (iT,T) =

(see

the quantitative aspects of the Gorter-Casimir

+ a)/(l-a)]yrc / (l+a)/(, - a \

Static field description

CHAPTER

III

this

23

B = H+4ttM,

As
description

is

equivalent to attributing to the superconductor a

magnetization per unit volume

Static Field Description


3.1. Perfect

M/=

diamagnetism

susceptibility of

the magnetic induction at the interior of a superconductor is due to


induced surface currents.! In the presence of an external magnetic

3.2.

of an external

field

H e is, therefore, the following:

in the interior: B,

H,-

M,-

0,

where M,-

is

at the surface:

3S ^

0,

outside:

Be =

U +H

1/47T.

The influence of geometry and the intermediate state


The great convenience of the diamagnetic mode of description is
illustrated by considering an ellipsoidal superconducting specimen in
an external field e which is parallel to the major axis. The conventional proof, that inside a uniform ellipsoid B, H, and
are all constant and parallel to
e is independent of susceptibility and therefore

the magnetization

s,

where

s is

where

is

B,

=
=

Be =

0,

H,

0;

and

H +H
e

0,

M, ^

0;

s,

where now H,

is

which

is

cf. [1], p.

are, respectively, the

found to have the value of -e\2mc (Kikoin and Goobar,


50 and p. 193 [2], p. 83).

III.l

and

III.2 yields

M,= -H /47r(l-D)
e

(IU.3)

and

H,
In the
distorted

22

semi-major and semi-minor axes, and

- (1 -b 2 /a 2 ) 112 For an infinite cylinder with its axis parallel to He


D = 0; for an infinite cylinder transverse to the field, D = \, and for
a sphere, D-l/3.
e

the field due to the

t That electron currents and not, for example, spins are responsible for
the diamagnetism of a superconductor is demonstrated by its gyromagnetic

1940;

a and b

Combining

magnetization of the sample.

ratio

For an

given by

is

it

at the surface: 3 S
outside:

(IH.2)

-(HMS-'>

this description is

in the interior:

= B.-AirDMu

the demagnetization coefficient of the specimen.

of revolution this

It is this field which causes the distorted field distribution near a superconductor as shown in Figure 3c.

formally correct, it is much more conby an equivalent one which treats the superconductor in the presence of an external field as a magnetic body with an
interior field and magnetization such that

show that

unnecessary)

H,

the field due to the sur-

ellipsoid

Although

now

where 3 S is the surface current density; and


e

applies to the superconductor. Further standard treatments

(with vector notation

face currents.

venient to replace

per unit volume;

(HI.1)

which means that the superconductor has the ideal diamagnetic

Even in the absence of a microscopic explanation of the phenomenon


of superconductivity, it is reasonable to assume that the vanishing of

field, the magnitude and distribution of this current is just such as to


create an opposing interior field cancelling out the applied one.
formal description of a macroscopic superconductor in the presence

-(1/4tt)H,

= HJ{l-D).

(IH.4)

neighbourhood of the superconductor, the external field is


by the magnetization of the specimen. It follows from the

Static field description

24

Superconductivity

continuity of the normal

component of B and of the

tangential

com-

He

ponent of //that for an ellipsoidal specimen the exterior field distribution is as shown in Figure 9. At the equator of the specimen

(a) transverse

cu nder
,

and

H sphere

\p)

H~ = H,= He l(l-D),

25

(III.5)

(aX

!i

(c) longitudinal

cylinder

at the pole

Hp =

Bi

(IH.6)

0.

For the longitudinal infinite cylinder with axis parallel to e ,D =


and eq = e The exterior field at the surface of the specimen is,
therefore, everywhere the same, and the cylinder remains entirely
superconducting until the applied field becomes equal to the critical

\\

Superconducting state

\\

Intermediate state

2H c /3

H c /2

Applied

He

Field

Fio. 10

Fig. 9

value

c.

The entire body then becomes normal. The magnetization


is shown in Figure 10, in which for con-

curve for such a specimen


venience

For

all

-A-nM is

plotted against

other ellipsoidal shapes,

e.

D ^ 0, and the non-uniformity of

the field distribution around the superconductor raises the question

of what happens when

Hcq = H > H
c

e.

To assume

that a portion of

shown in
would lead to a contradiction: the boundary between the
= C) but in the
superconducting and normal regions occurs where
now normal region the field would equal e
c There is, in fact,
no simple, large-scale division of such a specimen into normal and
superconducting regions, which allows a field distribution such that
< c in the latter, and
C in the former,
C at the
the specimen near the equator then becomes normal, as

Figure

Fig. 11

11,

H <H

H>H

boundaries.

H H

H H
\

H=H

Instead one must postulate, as was


b y F. London (1936), that once
>

first

(1

done by

- D)HC

Peierls (1936)

and

the entire specimen

is

subdivided into a small-scale arrangement of alternating normal


and superconducting regions, with B =
c in the normal regions, and

**

in the others.

The

distribution of these regions varies in such a

Superconductivity

26

way

that the total magnetization per unit

linearly

from

to

Mi = -HJ47r(l - D) = - BJ4*r

at

at

Hence, for

(1

H
H

c {\

- D),

c.

Bi

is

as the

when

the external field

is

reduced to zero, for as long

specimen remains superconducting.

(UI.8)

c,

trapped even

(III.7)

H -A-nDMi = H
H -(MD)(H -H
e

Thus if such a non-ideal specimen is placed in a


magnetic field sufficiently high to make it entirely normal, and the
field is then reduced, the anomalous regions will become superconducting before the bulk of the specimen. Should some of these regions
high critical fields.

be multiply-connected, then the flux threading them at the moment


of their transition into superconductivity can no longer escape, and

-D)Hc <He < Hc


M,= -(l/4irZ(/fc -fQ,
Hi

27

Static field description

volume changes

(III.9)

e).

Magnetization curves for a transverse cylinder (D = I) and for a


sphere (D = 1/3) are also shown in Figure 10. Note that the area under
each of the curves

is

given by

j MidHe = -H?I8tt.

(111.10)

just the magnetic work done on the specimen in raising the


from zero to
as cited in equation II.2. In the region
c
in which the specimen is neither entirely
(1 D)
c =5
e ^
c
normal nor entirely superconducting, it is said to be in the intermediate

This

is

field

state.

VI

The

H H

detailed structure of this state will be discussed in

Chapter

Applied Field H e

at this time it is only necessary to emphasize that this intermediate

state exists, in

some

field interval, for

Fig. 12

any geometry other than that

of a quasi-infinite cylindrical sample parallel to the external

field.

Trapped flux
important to distinguish the reasons and conditions for the intermediate state from those giving rise to the phenomenon of trapped
flux or the incomplete Meissner effect. As mentioned earlier, the
3.3.

It is

magnetic flux threading a multiply connected superconductor is


trapped by an indefinitely persisting current, and cannot change
unless the superconductivity of the specimen

is

quenched.

similar

As a result, after it has once been normal in an external field, such


an imperfect specimen is less than perfectly diamagnetic in an external
field

H< H

shown
the same

c,

and

retains a paramagnetic

moment in

units as in Figure 10.

The

ratio of

This

to

-H

c is

called the

The perfect conductor


To emphasize once again the
3.4.

difference between a perfect conductor

a nd a superconductor,

superconductor. Strains,

ment of the former, as developed by Becker

fections can create inside a superconductor regions with anomalously

field.

fraction of trapped flux.

situation can arise in a simply-connected but

non-homogeneous
concentration gradients, and other imper-

zero

in Figure 12 for a long cylinder parallel to the field, using

it is

useful to outline an electromagnetic treat-

the discovery of the Meissner effect.

et al. (1933) just before

Superconductivity

Static field description

In a perfect conductor, the equation of motion for an electron of


m and charge e in the presence of an electric field E does not
contain a retarding term and would simply be

add to Maxwell's equations the following two relations in order to


treat the electromagnetic properties of a superconductor:

28

mass

mv =
In terms of the current density J

of the electrons, one can write

where

eE.

form

in the

E=

(47rA /c ) J,

011.12)

A2

mc2/4wne2

(IE. 13)

length,

-6

(A)

(47rA

field

/c)curlJ+H

by a vector potential curl A

gauge such that div A

= 0,

(B)

0.

= H and

choosing a

(B) reduces to
4ttA

J+A

(BO

0.

for a density of

atom it has a value of

electrons corresponding to one electron per

order of 10

and

and
Replacing the

The parameter A has the dimensions of

E = (^AVJ),

(IU.11)

= nes, where n is the number density

III.l

29

the

cm.

Note that (A)

is

and thus describes the property of


and III.4 is

identical to UI.2,

perfect conductivity, but that the difference between (B)

Using Maxwell's equation curlE


2

(4ttA /c) curl

= -H/c,

J+H

one finds that

the important

(UI. 14)

0,

and applying another Maxwell equation curlH

Solution of this for any geometry

V 2 H = H/A 2

decays exponentially
(111.15)

any specimen
which decreases exponentially as one

that the solution of III. 15 for

geometry yields a value of

For a
direction from the plane x =

enters the specimen.

V 2 H = H/A 2

OH. 17)

= 4tt3jc yields for the

perfect conductor the equation

Von Laue (1949) showed

one that application of Maxwell's equations now leads

to

now shows that H, and not only H,

upon penetrating

into a superconducting specimen. For the semi-infinite slab described above, the solution of III. 17

is

H(x)

H(0)exp(-x/A),

(IU. 18)

semi-infinite slab extending in the x0, the

appropriate solution

which shows that for x >

is

A,

H(x)

0, in accordance with the Meissner

effect.

H(x)
Clearly, for x

H(0) exp ( - x/X).

P A, H(x) 0. Thus equation III.

interior of a perfect

6 confirms that in the


cannot change in
had when the specimen became perfectly

conductor the magnetic

time from the value


conducting.

it

(UI. 1 6)
1

field

Clearly the London equations (A) and (B) do not, in fact, yield the
complete exclusion of a magnetic field from the interior of a superconductor. Instead, the*y predict the penetration of a field such that
it

decays to

the

London

/e

of its value at the surface in a distance

penetration length.

Its

A.

This

is

called

existence has been fully confirmed

experimentally, although empirical values are consistently higher


3.5.

The London equations

for a superconductor

The

incorrectness of IU.16 was demonstrated by the discovery of


Meissner and Ochsenfeld (1933) that regardless of the magnetic history of the specimen, the field inside a superconductor always

vanishes. F.

and H. London (1935a, b; see

[2])

therefore proposed to

than those predicted by the defining equation ni.13, as will be discussed in a later chapter. The existence of this slight penetration of an
exterior field

must be taken into consideration in the discussion of


superconducting thin films, wires, or colloidal particles, and in a
detailed treatment of the intermediate state.

Static field description

Superconductivity

30

Applying curlE

= H/c to equation (B), one obtains


curl[E-(47rA

/c

J]

By defining a

particle density

0,

/i(R)

E (47rA

showing that

/c ) J

postulates.

particles of charge q described

by the wave

function

^/(r
the

mean

,r2 ,...,r N ),

current density at a point

H(r a )

magnetic

curlA(r a )

(III. 19)

given by

tnc
field,

.drN

A(ra)

The mean
is

0,

(111.24)

field

(H< i/c ).'

local value of the carriers

...dr N

Q,

and the current density

(^)[(4ttA /c)J + A].

same gauge as that leading to (BO, 111.25 for a simply connected


superconductor reduces to

(IH.21)

0.

W=

follows that
2

...J

^-c A(r )W*V8Ql.-r x )dr

(111.25)

In the

in the

= W v + (9/c)A,

p,
t

momentum

which can be rewritten as

W=W

therefore, one assumes that the wave function


is perfectly rigid
X
under the application of a magnetic field, that is, that
F() always,

-2J

A(R).

given by

If,

J(R)=

mc

brought into a magnetic

(111.20)

[ fl^tnv^o-^VamscR-rjx
xdr

it

is

p
.

vanishes, so that

then

QSU8)

wave functions of the superconducting carriers. In F. London's own


words ([2], p. 150): '...superconductivity would result if the eigenfunctions of a fraction of the electrons were not disturbed at all when

- ^-AOrJ V * f U(R-r.
8(R- ra)</r,

But if the particle density n(R) is a sufficiently smooth function so that


one can replace it by a constant n, then in view of the defining equation
III. 13, 111.24 is seen to be identical to (BOThus the London equation (B) or (B') implies that the magnetic

presence of a field

N
s

...dr N

A possible explanation of this is contained in the London equations


1

>=

= -n(R)

J(R)

themselves.

equation 111.22 can be written as

the system

In the absence of a

properties of a superconductor are due to a complete rigidity of the

R in the presence of a

field

is

j ...j P*Y8(R.-ra)dr

<f>

In a system of

grad<,

where is a scalar. In the most general case of a multiply connected


superconductor or a superconducting portion of a current-carrying
circuit, one cannot prove that < vanishes. Hence (A) does not always
follow from (B) and the perfect conductivity implicit in (A) and the
perfect diamagnetism in (B) must be considered as independent

...drN . (111.22)

0.

(IIL26)

The London equation thus implies that superconductivity is due to a


condensation of a number of carriers into a lowest momentum state
P s = 0. By the uncertainty principle this requires the essentially unlimited spatial extension of the appropriate wave functions, and
makes it impossible for them to be affected by local field variations.
It also follows from 111.26 that
v*

= -(qlmc)A,

(IH.27)

'

32

Superconductivity

33

Static field description

showing that in a simply connected superconductor the charge flow


is entirely determined by the externally applied field, and exists only
in

its

presence.

3.6. Quantized flux

London already observed ([2], p. 151] that the unlimited extension


of the wave function of the superconducting charge carriers has a very
fundamental consequence in a multiply connected superconductor.
F.

Consider, for example, a superconductor containing a hole. The wave


functions must then be single valued along any closed path enclosing
orbit

By analogy

wave functions in an atomic


one can then apply the Bohr-Sommerfeld quantization rules

the hole.

and require

to the electronic

that for the superconducting charge carriers

p-dl

nh,

(IU.28)

(J>

and the fluxoid

is

just equal to the total flux associated with the hole.

This flux is thus seen to be quantized.


The quantization of flux was verified experimentally by Doll and

Nabauer (1961) and by Deaver and Fairbank (1961). These


ments have shown that the quantum of flux is given by

A This shows that q

kC

= 2e,

2x10

that

is,

gauss-cm2

experi-

that the superconducting charge

of electrons. It has already been mentioned that


indeed this is the fundamental premise of the microscopic theory.
A number of authors (Byers and Yang, 1961; Onsager, 1961;
Bardeen, 1961b; Keller and Zumino, 1961; Brenig, 1961) have

carriers are pairs

extended the

London argument

for flux quantization in a rigorous

and Yang as well as Brenig have shown


explicitly that the quantization is due to a periodicity of the free
energy of the superconductor as a function of flux. The free energy of
the normal phase is essentially independent of flux, and there must
fashion. In particular Byers

along any path enclosing the hole. According to 111.25

means

jMttA 2
(b

Since

this

then

that

m
J-dl+&>A-dl =

he

n-

(m.29)

H = curlA, the contour integral of A equal to the surface


H over the area enclosed by the contour, and this in turn
is

integral of

Thus

H-c/S

he

J<fl+0 =

0.

(111.30)

(111.31)

London called the left-hand side of this equation


that according to 111.31 such a fluxoid

is

a.

fluxoid,

and we see

quantized in integral

multiples of

he

&Note that
compared

if

the contour

is

This flux periodicity of


(1961).

equals the magnetic flux <P threading the contour:

<J)A-rfl=

therefore occur a corresponding periodic variation of the critical


temperature at which the free energies of the two phases are equal.

(H1.32)

taken at a distance from the hole large

to the penetration depth A, the current density vanishes,

Tc

has been observed by Little and Parks

The Pippard non-local theory

35

As a consequence it is impossible to test the validity


penetration law, such as, for example, the London
particular
of any
small specimens.

CHAPTER

IV

relation III. 17,

The Pippard Non-local Theory


The penetration depth A
The London equations lead
4.1.

to an exponential penetration of an

externally applied magnetic field into a superconductor, so that the

penetration can be characterized by the depth A at which the


fallen to 1/e

of

its

field

has

value at the surface. Quite in general, and inde-

pendently of any particular set of electromagnetic equations for the


superconductor, one can define the penetration depth for an infinitely
thick specimen

by

H{x)dx.

(1V.1)

This would apply equally well to an exponentially decaying field as


to one, improbable though it may be, which remains constant to a

and then vanishes suddenly.


Shoenberg ([1], p. 140) has pointed out that in this way one can
treat problems involving either very thick specimens (thickness a > A)
certain depth

or very thin ones {a

<^ A)

independently of a detailed knowledge of

the appropriate electromagnetic equations. Using IV. 1 to calculate


the ratio of the magnetic susceptibility \ of a sample into which the
field has penetrated, to the susceptibility xo of an identical

applied

sample from which the field is entirely excluded, he finds equations


of which the following are applicable to a plate of thickness la in a
uniform field parallel to its surface

One can then deduce directly the variation


question, without having to make any
parameter
in
of A with the
assumptions about the true penetration law. The results of such
measurements can therefore help to choose between different electromagnetic theories if these predict different parametric dependences of

field,

x/xo

The

detailed

IV.2 and

all

form of the

=
=

- A/a
2

aa /A

field

for

>

(IV.2)

A,

The oldest method of measuring the penetration depth consists of


determining the magnetic susceptibility of samples with large surface
to volume ratios to make the penetration effects appreciable.
Shoenberg (1940) measured the temperature dependence of the susceptibility of a mercury colloid containing particles of diameter
between 100 and 1000 A. Desirant and Shoenberg (1948) used composite specimens consisting of about 100 thin mercury wires of
3
diameter about 10" cm, and Lock (1951) carried out extensive

measurements of the magnetic behaviour of thin films of tin, indium,


and lead. Casimir (1940) suggested a method using macroscopic
specimens in which the mutual inductance was measured between
two coils closely wound around a cylinder of superconducting
material. It was applied successfully by Laurmann and Shoenberg
(1947, 1949), by Shalnikov and Sharvin (1948), and most recently
with certain refinements by Schawlow and Devlin (1959), and by

McLean

for a

<

(IV.3)

A.

penetration does not enter at

all

into

does so only through the numerical parameter a

in the

equations for

(1960).

A superconductor has finite surface impedance at high frequency,


this

impedance

is

limited in the superconducting phase by the

penetration depth A, as in the normal phase

it is

limited by the skin

Pippard (1947a) was the first to use this as a means of


measuring A, and he and his collaborators have carried out a large
number of experiments at different frequencies and varying experimental conditions (see Pippard, 1960). Basically all these measurements involve observing the change in the resonant frequency of a
depth

other equations for large specimens of other shapes, and


34

impurity content, etc.

the penetration depth.

and
x/Xo

by measurements on large specimens; with very small

specimens this can only be done if one can determine absolute values
of a and of A, which is very difficult. On the other hand, one can
measure the variation of the susceptibility of large or of small specimens with any parameter affecting only A: temperature, external

S.

36

Superconductivity

cavity containing the specimen

when

The Pippard non-local theory

the specimen passes

normal to the superconducting phase. At

T< Tc

from the

where \<8, these

changes are proportional to 8 - A. If 8 is independent of temperature,


as is the case for a metal in the residual resistivity range, then any
temperature variation of the observed changes must be due to the
temperature variation of

Dresselhaus et

A.

al.

(1964) use instead of a

cavity a rutile resonator to which the specimen

is

coupled, and also

observe changes in the resonant frequency.

measurements, however, shows a small deviation from IV.5 at


/

< 0-8, which becomes particularly pronounced at low temperatures.

This deviation

Schawlow (1958), Jaggi and Sommerhalder (1959, 1960), and most


al. (1960) have measured the penetration of a
field

through a thin cylindrical film of thickness

less

y(.t)

barely discernible in the normal plot of A(r) vs. y(t),


'2

1400p

than

1200 >

the penetration depth.

4.2.

is

= (1-f4) -1

but is displayed strikingly in Figure 13,


,
which shows for Schawlow's results (1958) the variation of the slope,

where

recently Erlbach et

magnetic

1000

The dependence of A on temperature and

field

According to the London theory, an external magnetic field penetrates into a superconductor to a depth characterized by (see equation HI. 13)

where n s

is

the

(mc

It is

400

be the only temperature-dependent factor


in fact the Gorter-Casimir two-fluid

200

number density of the superconducting electrons.

reasonable to expect

this to

and

in this defining equation,

BCS
THEORY

d y 600

2 112

l4irn se

model assumes that


n s (t)

where i^it)

is

#xq*m

a1.0

av.4)

1
2.0

1.5

25

depth one obtains

3.5

4.0

4.5

5.0

A(0)

the penetration depth at

(IV.5)

(IV.6)

T= 0K. Very near T

IV.5 can be written

The solid line indicates the values of dX/dy calculated


by Miller (1960) on the basis of the BCS theory; the experimental
results appear to deviate less from IV.5 than is predicted by theory.
Furthermore it appears that in impure specimens no deviation from
IV.5 can be found at

as

*-Ww.

Fig. 13

dXjdy, with y(t).

= A(0)/(l-/4) ,/2
= (mc 2/47r/;/0)e 2) ,/2

A(/)

is

3.0

the order parameter.

The temperature dependence of


is given by 11.25', so that substituting this and IV.4 into the defining equation for the penetration

where

37

sented to a very high degree of approximation by IV.5. This is a


striking success of the phenomenological theories discussed in the
preceding chapters. Close inspection of the recent very precise

The slope of A(r)


(IV.7)

Daunt et al. (1948) were the first to point out that the empirical
temperature variation of the penetration depth can indeed be repre-

all (Waldram, 1961).


plotted as a function of y(t), as well as the inter-

cept, yield values for A(0) if one ignores the small deviations from
IV.5. Appropriate empirical values for pure bulk samples are shown
in

Table

II.

They exceed by a

expect from the

London

what one would


one makes rather

factor of about five

definition IV.6, unless

38

The Pippard non-local theory

Superconductivity

unlikely assumptions of low densities of superconducting electrons


or of a large effective mass. Experiments on very small samples, and

measurements on impure metals, yield even higher values of A(0),


although none of the factors in IV.6 appear to depend on size or
purity. This failure of the

London theory

will

be discussed

in the

Table

II

Al

A(0)(A)

Sn

500
1300
380-450*
640
440
390
510

Tl

470-600*
920

Cd

Hg
In

Nb
Pb

is

quite considerable.

Pippard pointed out that to assume that the entire entropy change
takes place in the thin layer into which the field penetrates would,
therefore, result in an unreasonably high entropy density in this layer.

Yet this is just what one is led to believe by the London model, according to which the superconducting wave functions or, in two-fluid
language, the corresponding order parameter #", remains rigidly

following sections of this chapter.

lement

39

difference between the normal and superconducting phases, which

unchanged by the application of an external

Reference

field.

Any change in the

Faber and Pippard, 1955a


Khaikin, 1958

Laurmann and Shoenberg, 1949


Lock, 1951

McLean and

Maxfield (1964)

Lock, 1951
Pippard, 1947; Laurmann an
Shoenberg, 1949; Lock, 1951
Schawlow and Devlin, 1959
Zavaritskii, 1952
* Anisotropy.

Z0
Pippard ( 1 950) investigated the change of the penetration of a small,
r.f. field

e is

kMc/s) at a given temperature as an external d.c. field


from zero to the critical value. This change, divided by

the penetration depth in zero

field, is

plotted against temperature in

There are clearly two effects: one at low temperatures


(which Bardeen (1952, 1954) has shown to follow from an extension
14.

London equations to include non-linear terms), and one near


This latter involves a change in A with
in just that region in
which A varies appreciably with temperature. By a thermodynamic
derivation Pippard has shown that this temperature variation of A
leads to a dependence of the superconducting entropy on field. He
of the

Tc

3.0

35

TC

(9-4

raised

Figure

25
T(K)

finds that

S(He )-S(0) =

^^V/U

Fig. 14

thermodynamic functions with field must therefore be confined to the


thin layer into which the field penetrates.
Recent measurements of the field dependence of the penetration
depth at
and 3 kMc/s (Spicwak, 1959; Richards, 1960, 1962; Pip1

pard, 1960; Dresselhaus et

al.,

1964) have

certain unexpectedly complicated features.

of the change frequency dependent, but even

av>8)

for a superconductor of total surface A, assuming IV.5 to hold. Near


Tc this change contributes as much as one-fourth of the total entropy

is

sign can change under


an increase in the pene-

its

certain conditions. In particular there can be


tration depth when the applied field

~/4)]

shown that this effect has


Not only is the magnitude

parallel to the specimen surface,

and at the same frequency a decrease when the field is perpendicular.


Bardeen (1958) and Pippard (1960) have suggested that these complexities may be due to field induced deviations of the superconducting
and normal electron densities from their equilibrium values.
4

40
4.3.

The Pippard non-local theory

Superconductivity

41

a broadened transition even for ideally homogeneous samples.


Goodman (1962c) has discussed this in some detail.

The range of coherence

The unreasonably high entropy

density in the surface layer led

Pippard (1950) to propose a basic modification of the London model,


according to which the order parameter changes gradually over a

which he calls the range of coherence of the superconducting wave functions. In terms of the microscopic theory this
certain length f ,

distance can be considered as the typical size of the

4.4.

The Pippard

non-local relations

In 1953 Pippard measured the penetration depth in a series of dilute


alloys of indium in tin, and found that the decrease in the normal
electronic

mean

free

path of the metal was accompanied by an

Cooper pairs. Any

change in the thermodynamic functions of course extends over as


wide a region as the change in the order parameter, and thus a value
> A would correspond to a more reasonably small value of the fieldinduced entropy density.
Pippard (1950) obtained an estimate of the range of coherence of
the order parameter by minimizing the Gibbs free energy of the superconductor in the presence of an external field. The resulting relation
between the fractional change of the penetration depth and the ratio
A(0)/ allows him to estimate from his experimental data on the field
-4
effect on A that M 20A(0) 10
cm. Such a distance is much larger
than the smallest colloidal specimen or the thinnest films in which

10-

is still known to exist, and it is therefore of great


importance to realize that, to quote Pippard (1950, p. 220) the range
of order must therefore not be regarded as a minimum range necessary

superconductivity

'

for the setting

up of an ordered

state,

10

20
)0

Strong support is given to the existence of this range of coherence


by the extreme sharpness of the superconducting transition under
suitable conditions.

De Haas and Voogd

of tin taking place within a range


of one millidegree, and a sharpness approaching this value has come
to be the criterion for the quality of a specimen of suitable shape and
orientation. Applying a simple statistical argument, Pippard shows

would create a broader

40

50

(cm)
Fig. 15

(1931) have observed

resistive transitions in single crystals

that fluctuations

.L
30.

but rather as the range to which

order will extend in the bulk material'.

transition unless the super-

appreciable rise in the value of A(0). This has been confirmed by

Chambers

(1956),

and by Waldram

(1961),

whose

results are

shown

Figure 15. Such a dependence of A(0) on the mean free path is quite
incompatible with the London model, for clearly none of the para-

in

meters in the defining equation IV.6 varies appreciably with the

mean free

conductivity of a bulk sample can be created or destroyed only over

electronic

an entire domain of diameter M 10A(0).


In the next section as well as in 6.5 it will be shown that the range of
~
coherence is much smaller than 10 4 cm in low mean free path alloys
as well as in certain pure metals. For such materials one would expect

This experimental result, added to the previous questions which had


been raised about the correctness of the London phenomenological

path.

treatment, led Pippard (1953) to develop a fundamental modification


of this model, based on the concept of the range of coherence of the

42

Superconductivity
will

it

be

remembered, lead to the relation


A(R),

(IV. 10)

4-nXl
A^,

= mc 2 /4ime 2

One way

They do not, however,

so that one can also write this as

J(R)

mean

free path

is

London value IV.6 by

According to Pippard this is because IV.P1 does


not correctly describe the relation between current density and the
vector potential in such a case. PI still implies, as does equation
III.B', the basic London idea of a wave function which is completely
factor of four to five.

A(R).

(IV.ll)

to introduce a dependence of the penetration depth

electronic

satisfactorily explain the finding that A(0) in

pure, bulk superconductors exceeds the

2
IW~

These equations satisfactorily explain the onset of a mean free path


on the penetration depth at a critical value of /, as found by
Pippard and others, as well as the very large penetration depth values
obtained from experiments where / is limited by boundary scattering.
effect

J(R)= where

43

The Pippard non-local theory

superconducting phase. The basic London equations,

on the

to write
rigid

under the application of an external

field

momenta are ordered or correlated over an


(IV.P1)

distance over which

H and A vary

is

because the electronic

infinite distance.

Thus the

quite immaterial the same kind of


;

would hold if the field varied very slowly as if it varied very


But according to Pippard the range of momentum coherence
4
not infinite but only about 10~ cm, so that the electromagnetic

relation

where is a constant of the superconductor in question, and (/) a


parameter depending on the mean free path /. It is evident from the
analysis in Chapter

HI

that IV.P1 leads to an expression for the field

penetration into a semi-infinite slab which has the

H{x)

London form:

H exp(-x/X),

rapidly.
is

response of the superconductor should be affected profoundly


field varies

only

if

The
but where

now

As experimentally
that (/)

is

XL

(TV. 12)

found to increase with decreasing

must decrease as

is

(/)

(/)

->/as /->0. This

is

if

is

somewhat analogous

to the

problem of electrical

conductivity in a normal conductor, for which the relation

a(/)E(R)

(IV. 14)

decreases.

tends toward this value as


the case

if,

/-> a>,

the

could apply

clear

with the range of coherence of the pure superconductor,

and assumes that

A relation like PI

the field varied slowly over a distance of the order off.


situation

J(R)
/, it

As a first step toward a modification of the London model, Pippard


identifies

rapidly over this distance.

but that

for example,

is

valid only

if

E(R) varies slowly over a distance of the order

of/.

An

which
varies inversely as the square root of the frequency. At sufficiently low

applied alternating field penetrates only a finite distance,

temperatures and high frequencies, the electronic

mean

S,

free

path in

normal metal may be longer than this skin depth, so that electrons
may spend only part of the time between collisions in the field penetrated region. Pippard (1947a) showed how this makes the electrons
less effective as carriers of current and leads to a higher surface
resistance, as observed by H. London (1940) and Chambers (1952).
Under these conditions Ohm's law (IV. 14) can no longer be a valid
approximation; the current at a point must be determined by the
the

(IV. 13)

where a is a constant of order unity. (/) is thus an effective range of


coherence which has a size (w 10 " 4 cm) characteristic of the metal in
a pure superconductor, but which becomes limited by the normal
electronic

10- 4 cm.

mean

free path as the latter

becomes much smaller than

integrated effect of the field over distances of the order of the


free

mean

path (see Pippard, 1954). The details of this so-called anomalous

44
skin effect were

worked out by Reuter and Sondheimer

(1948),

who

derived that

JCR)_

3a CR(R-E)e- R "dT

4ir/J

(IV. 15)

is just the expression IV.9 for the range of coherence derived


from an uncertainty principle argument, with a = 018.
From P2, the penetration depth A as defined by IV. 1 can be

This

evaluated explicity in two limiting cases:

where a is the d.c. conductivity and / the mean free path. The form
of this equation ensures that in the case of a rapidly varying field the
current density at a point R is determined by the integral of the field
over a distance comparable to the mean free path /.
In a superconductor of range of coherence , the current density
at a point in the case of a rapidly varying field should also be deter-

mined by an integral of the field over a distance of the order of g, and


not, as is implicit in the London equation as well as in PI, by the field
variation over a quasi-infinite distance. Because of this analogy to the

anomalous conduction

in a

normal metal, and because some special


known, Pippard (1953)

solutions of equation IV. 15 were already

proposed as the basic relation for the electromagnetic response of a


pure superconductor the equation

3ne 2
J(R)= -

4n$

Somewhat misleadingly, as
London equation is a truly

R(R-A)e~ R ltdr

mS

this

(IV.P2)

R*

erroneously implies that the basic

local relation,

P2

is

called the Pippard

non-local relation.

Sommerhalder (1962) have observed this effect.


The validity of P2 is strongly supported by Bardeen's proof ([5],
pp. 303 ff.) that an energy gap in the single electron spectrum requires
a non-local relation between current density

BCS

and vector potential. In


theory leads to a relation entirely equivalent to P2 if one

for

5<

for

g> A, (Pippard limit).

A,

(London

feM

nv

/7re(0),

striking

BCS

value

g,

= 0l8hv

/k B Tc

(IV.17)

(IV. 1 8b)

agreement with the

chapter will mention

BCS value

of 0-18, as cited in IV.17.

how measurements

of the transmission of

strong support to this value.


Peter (1959) has solved the Pippard non-local relation
case of cylindrical superconducting films of thickness
radius

r.

He

finds that

film to a value

the energy gap at 0K. Substituting the


3-52k B Tc IV.16 becomes

(IV. 18a)

infrared radiation through thin superconducting films lends further

(IV. 16)

is

limit),

The second of these is the one applicable to the case of an infinite


mean free path, and correctly predicts a penetration depth into very
pure superconductors which is much larger than the London value.
IV. 18a is identical to IV. 12 obtained directly from PI. This is of
course a reflection of the fact that PI is the limiting form for <^ A of
the more general equation P2.
Equations IV. 18 show that the range of coherence of a superconductor can be calculated from absolute values of the penetration
depth. Faber and Pippard (1955a) have in this way obtained values
-5
of 2-1 x 10
cm for tin, and 12-3 x 10 -5 cm for aluminium. These
values differ very much, but when substituted into equation IV.9
together with known values of Tc and v (from anomalous skin effect
data [Chambers, 1952]), both correspond to a = 015. This is in

Hi such

an external

field

's

P2 for the
d < A and

penetrates through the

that

HJH^Vrd^faFitld).

&=
2e(0)

V(Zolt)*L
1/3

assumes

where 2e(0)

A.-

later

This relation leads to a reversal of the phase of the magnetic field


penetrating into a superconductor (Pippard, 1953). Drangeid and

fact the

45

The Pippard non-local theory

Superconductivity

(IV. 19)

the range of coherence in a specimen of unlimited mean free path,

and can be calculated from IV.17; is the actual range of coherence


in the film, and A should be the London penetration depth as calculated from IV.5 and IV.6. Schawlow (1958), however, has shown that

46

good agreement with

Superconductivity

The Pippard non-local theory

measurement on

(Gor'kov 1959b) that the mean free path dependence of this aspect of
the range of coherence is given by

can be found by
substituting for A the empirical value for bulk samples (510 A) and
considering as being determined by the size-limited mean free path
of the electrons in the films. A similar analysis has been used by

Sommerhalder
It is

his

tin films

(1960).

now generally accepted that whenever one applies the equation

of the Pippard theory (or those of the Ginzburg-Landau treatment


to be discussed presently) to the case of small or impure specimens,

oX-

1/2

47

(IV.21)

(0

a function of the mean free path shown graphically by Gor'kov


and approximated to within about 20 per cent by the simple expression
(Douglass and Falicov, 1964)
x(/) is

one obtains good agreement by using for the


in

a bulk sample, not the London value

A^.

ideal penetration depth


but rather the depth deter-

mined experimentally. For example, the results of Whitehead (1956)


on the magnetic properties of mercury colloids were shown by
Tinkham (1958) to be in excellent agreement with the prediction of
the London limit of the Pippard theory if one modifies equation
IV.18b and writes

where X b

is

now

\W

In the limit

<^

IV.21 thus reduces to

The
with

(IV.20)

relatively

mean

= V(to 0-

the empirical penetration depth for a bulk sample

(IV.23)

slow variation of this aspect of the coherence length


path is essentially due to the fact that not every

free

electronic collision destroys the superconducting coherence.

The other aspect of the range of coherence

and takes the place of the London value A L The mean free path / is
limited by boundary as well as by impurity scattering. Ittner (1960a)

(IV.22)

(s4

x(/)

is

that

it

determines the

distance over which the magnetic field or the vector potential at a

given point influences the current density. This

is

expressed by the

has similarly found that such a modification of the Pippard equations


adequately predicts the results of the observations by Blumberg (1 962)

Pippard equation IV.P2.


actual

mean

critical field of moderately thin films. In analysing the magnetic


behaviour of small (or very impure) specimens, for which * / <^ A, it
is

thus in general possible to obtain adequate precision without

attempting to solve the

difficult relation IV. P2. Instead one can use


IV.20 to calculate the penetration depth, and then substitute this value
of A into the London equation IV.10.
In discussing the mean free path dependence of the coherence

length one must

superconductor

remember

in

that

two subtly

tioned in Section 4.3,

is

it

is

related to the behaviour of a

different ways.

One of

these, as

men-

the distance over which the order parameter

of the superconducting phase varies. It is this aspect which, for


example, in Chapter VI will enter into the discussions of the width
of a boundary between the normal and superconducting phases.
It

follows from Gor'kov's analysis of the influence of impurities

is

important

in this application is the

distance between electron collision, so that

tion IV. 13 applies. This

of the

What
means

now

equa-

that

(IV.24)

/ <g
It is this mean free path dependence which enters, for
example, into equation IV.20.

for

The Ginzburg-Landau phenomenological theory

49

With these one then finds from V.3, remembering that the free energy
difference between the phases equals the magnetic energy, that

CHAPTER V
=
Ht2 _

The Ginzburg-Landau

47ra

47r(rc -r) 2/fl - x2


(V.5)

P
Near

Phenomenological Theory

Tc

indeed

is

\8T/ T=Te

Pc

known

to vary linearly with (Tc -T), so that

the correctness of equation V.5 justifies the assumptions V.4. All

In 1950 Ginzburg and Landau (G-L) introduced a phenomenological

further

thermodynamic manipulations are now

approach to superconductivity which, like that of Pippard, modifies


the absolute rigidity of the superconducting order parameter or wave
function which is implicit in the London model. Although the theory

and

other conclusions drawn using V.4 are restricted to tempera-

was

originally formulated so as to reduce always to the 'local'

London equations in zero field, Bardeen

(1954) has shown that it can


be modified so as to be compatible with a non-local equation of the
Pippard type. Furthermore, Gorkov (1959, 1960) has derived the G-L

equations, under certain conditions, from his formulation of the

BCS

G-L introduce an order parameter >p which they normalize so as to


2
= n s where n s is the density of the superconducting elec\*fi\

ip is thus a kind of 'effective' wave function of the superconducting electrons. According to the general Landau-Lifshitz theory
of phase transitions (1958), the free energy of the superconductor

trons,

depends only on
peratures near

tures very near

and can be expanded in series form for temIn the absence of an external field, the supercon-

\ifj\

Tc

ducting free energy (per unit volume, as are all equations


(7,(0)

= GB (O) + o#| 2 +(|8/2)|0| 4

Minimizing the free energy with respect to

2
\<p\

listed) is

then

(V.l)

Tc

Both Bardeen (1954) and Ginzburg (1956a) have

full superconducting range


by introducing different forms for a(T) and @{T), the former using
expressions based on the Gorter-Casimir two-fluid model.
The outstanding contribution of the G-L model in any temperature
range arises from its ability to treat the superconductor in an external
field

H H

now increased not only by the


e
c The free energy G s (He) is
volume term H*/9v, but also by an extra term connected with
.

the appearance of a gradient of

M>l

=-/A

i/j,

as

ifj

is

not completely rigid

in the

Such a gradient would contribute to the energy in


analogy to the kinetic energy density in quantum mechanics which
depends on the square of the gradient of the wave function. Introducing this extra energy is equivalent to requiring that not change
presence of

e.

*/>

too abruptly.

One is thus

led to a concept of gradual, extended varia-

tions of the superconducting order

parameter quite analogous to

Pippard's model of the range of coherence.


In order to preserve gauge-invariance,
term to be

yields the zero field

G-L assume the extra energy

J-L/^-^aJ

equilibrium value
2

possible, but they

considered extensions of the model to the

usual

theory.

make

all

(V.6)

(V.2)

In the immediate vicinity of Tc one can assume that the coefficients a

where A is the vector potential of the applied field, and e* a charge


which, as stated in the original version of the theory, 'there is no
reason to consider as different from the electronic charge'. Modifica-

and

tions of this view will

from which

/?

(7,(0) - (7(0) =

-<x /20.

(V.3)

have the simple form


*(j)

jB(D

= p(Tc ) -

(rc -r)(sa/ar) r=re

t2

(V.4)

and

48

fie

be discussed presently.

G-L thus write

G s(He ) =

m~M
*

GM + fr+^A
Uc-W*~A

12
-

(V.7)

50

The Ginzburg-Landau phenomenological theory

Superconductivity

One must now minimize


leads to the

with respect to both

this

tp

and to A, which

Note that with

geometry

this

two equilibrium equations:

H=

curlA

~
dz

2m\

(V.G-L1)

The meaning of

dtfj*

these equations

becomes

clearer

by introducing a

dimensionless parameter k defined by

V2 A =

-^J, =
c

t^'w^-W^
mc

2e*
KT

:=

h
Aire*"'

+
In a very

weak

stant (that

is,

A.

4>

A-ne*

0o>

l^ol

iff

remains practically con-

and G-L2 reduces

mc'

Ag-

where

A=

Aire*

to

The

mc2

mentally,

C A.

(V.8)

and

be expressed.

G-L

c is

field, k,

and

are the three

theory which are to be determined experi-

in terms of

(V.12)

4ne* 2 ifa

again denotes zero

subscript

parameters of the

(V.ll)

<

(V.G-L2)

the function

V0 = 0,

V2 A

I0I

mc

HxO,

field,

rigid),

.,/ZcAq,

.-,

which various

the bulk critical

field.

field

is

weak

tion depth of a superconductor in the

and

size effects

can

the empirical penetrafield limit,

and

is

the

quantity which through equation V.12 determines the zero field

Here, as in V.2, the subscript

denotes the zero field value. This of


London's expression (B). Non-local versions of the G-L
treatment are obtained by substituting an integral expression for the
second term in V.6. In its present local form the G-L treatment is
course

is

just

restricted to temperatures near

Tc

for

two reasons:

in the first place

equilibrium value of the order parameter

i/jq. For a bulk sample conwas discussed in the previous


chapter, and this in turn affects both O and k.
k can be determined in a number of ways, two of which follow
directly from the defining equation V.
In the immediate vicinity of
Tct the experimental variation of A (/) can be expressed by IV.7:

taining impurities A

increases, as

1 1

because of the simple forms assumed for the functions a(T) and fi(T),
and secondly because only near Tc is A > and can the non-local

electromagnetic character of superconductivity be ignored.

The

set

G-Ll and G-L2 of coupled non-linear equations

in

ifj

and

A have been solved for essentially one-dimensional problems. Taking


the z-axis to be normal to the infinite superconducting boundary, the
field

H along the>--axis, and the current J

x-axis,

and potential

Also one can write

A along the

one obtains (using V.l)


2
</

dz

-tN**w^*-t^-*
^4
A
2

dHr
dT T=Te

so that

dH

8/jV dT T=T

mc 2

(V.10)

xAT,

Thus k

is

xT 2 xX 4 (0).

seen to be temperature independent, at least for

(V.l 3)

T& Tc

52

The Ginzburg-Landau phenomenological theory

Superconductivity

One can
from the

also use the expression for the penetration depth derived

BCS theory to be,

very near

Tc

larly in

53

measurements used to evaluate A L (0), and in part, particuthe case of aluminium, due to the large value of because of

skin-effect

>

which non-local conditions set in very close to Tc The values of k


calculated from supercooling are probably the most reliable.
In terms of the parameters k, Aq, and
c , equation (V.9) reduces
.

-^
V2

(l

_-./!

_y?/ik-

J/2

V2 \ATj

so that
,*2

,2

4/i

2
,

dH

dT t=t

x7?xAf(0).

(V.14)

where now A L (0)

is

the

London penetration depth

111.13 using the actual free electron density n. This

from the value of the normal

state

calculated

from

resistance

to

Ginzburg (1955) pointed out already before the formulation of


theory that values as calculated from V.13 and V.14 could
be made to agree very well with those obtained from supercooling
data by taking e* = 2 or 3e. More recently Gor'kov (1958) has formution.

BCS

BCS microscopic theory in


terms of Green's functions, and was able to show (1959, 1960) that
the G-L equations G-Ll and G-L2 are identical to his expressions
lated the electromagnetic equation of the

Tc when ip is taken to be proportional to the energy gap, and


when one takes e* = 2e. This again is an indication that the current
carriers in superconductivity are the doubly charged Cooper pairs.
near

this value

of e*, V.13 and V.14 yield

*c= 108 xlO 7

and

= 216xl0

= 0, V.9' is satisfied in the absence of an external


{A = 0) by ip2 = ipl; difi/dz = 0. In other words, the presence of
the phase boundary as such has no influence on the function tp, which
the boundary, z

field

has the same value

tp

everywhere. In the presence of an external

field

however, this solution no longer applies, and one must integrate

V.9' and V.10 with the boundary condition


the condition

H = dA/dz = H

and

e,

difi/dz

tp

= ipl for z--co, and


= 0. This integra-

for z

cannot be carried out exactly. Neglecting higher order terms,


however, one finds equations for
and for A as functions of z. At
tion

i/j

= 0,

the value of

tp is

CV 13)
'

4(+V2)

^o

m 0- 1 , this equation predicts a decrease of ip by only


about 2-3 per cent when e = c It is not surprising, therefore, that
the change in penetration depth with field is also very small. This can
With values of k

dH

dT T=T

tMo),

(V.130

= ipl, and

dz

At
is

use results on supercooling, as will be discussed in a subsequent sec-

With

tp

can be calculated

anomalous skin

(Chambers, 1952).
Another method of calculating k for a given superconductor

the

Far from the phase boundary, for z-> o,

dHc

dT T=T

Tc X L (0).

(V.140

be calculated formally by using the defining equation IV. 1 from which


one finds that, with a weak measuring field normal to

For

tin,

the

first

of these yields k

= 0-158,

values which are in excellent agreement.

the second 0-149,

two

8(k+V2) 2 # 2
(

For indium, however, the

respective values are 01 12 and 0051 (Davies, 1960; Faber, 1961).


For aluminium, the equations yield 005 and 001 (Davies, 1960).
This lack of agreement may be in part due to errors in anomalous

An

For a measuring

field parallel to

(V.16)

e,

the effect

is

tripled.

54

Superconductivity
It is

ifj

evident that in the limit k-*0, the effect of the external field on
vanishes, so that one returns to a situation formally

and on A

CHAPTER

London picture. It must however be noted that even


deduced from the empirical value of A As a result

equivalent to the
for k

= 0,

one can

j/tq

is

in certain cases,

VI

such

as, for

example, the treatment of very

thin films, allow k to vanish without necessarily reducing the


treatment to the London one.

The Surface Energy

G-L
6.1.

The surface energy and

the range of coherence

wave
on a
boundary between the superconducting and normal phases. H.
London (1935) already pointed out that the total exclusion of an

Closely tied to the range of coherence of the superconducting


functions

is

the existence of an appreciable surface energy

external field does not lead to a state of lowest energy for a super-

conductor unless such a boundary energy exists. In the presence of


an excluded external field,
e the energy of a superconductor in-

creases by Hg/Sir per unit volume. It would, therefore, be energetically

more favourable

for a suitably shaped superconductor to divide up


number of alternately normal and superconducting
layers such that the width of the latter is less than A, and that of the
former very much smaller than that. The resulting penetration of the
into a very large

external magnetic field into the superconducting layers

much reduces

the magnetic energy of the sample, while the extreme narrowness of


the normal layers keeps negligible their contribution to the total free

energy. This situation

is

made energetically unfavourable by the


To make each superconducting layer

existence of a surface energy.

narrower than A, a slab of thickness c/must have d/X such layers. This
is avoided by an interphase surface energy cc, per unit surface whose
contribution exceeds the gain in magnetic energy, that

is:

Hid

2d

'

(VI.l)

8tt

where the energies have been calculated for a volume of slab of unit
surface area.

Hence

>

XH[
28tt

55

(WW)

Superconductivity

56

The surface energy

It is convenient and customary to express the surface energy in terms


of a parameter A ' of dimensions of length, such that

C is the configurational boundary such that if #" dropped


C after being constant up to that point, one would

sharply to zero at

(VI.2)

have the same superconducting free energy as the actual amount. The
free energy per unit volume of the superconductor is lower than that

(VI.3)

normal state by an amount Hc/Stt. A configuration boundary


shown on the inside of the magnetic boundary is essentially equivalent to a reduction of the superconducting volume and hence an
increase in the total free energy by an amount equal to HcI%t times
the distance C-M per unit area of interphase boundary. The intro-

8tt

Thus one

Similarly

57

in the

sees that

as

A'>

duction of the Pippard range of coherence thus leads to a configurational

boundary surface energy A

'

w . From this one must subtract

the decrease in energy due to the penetration of the


indicates that the distance

C-M

field.

Figure 16

corresponds to the resulting net

surface energy parameter

A m -A.

Position
Fio. 16

is

H H

the condition for the diamagnetic behaviour of superconductors.!

Empirical values of

'

for pure superconductors turn out to be an

is

of the externally applied

field

iV and

along a direction perpendicular to

One can

two effective boundaries, indicated by


and C.
is the magnetic boundary defined so
that if inside the superconductor B =
and then dropped
c up to
off sharply to zero, the total magnetic energy would equal the actual
value, given by the integral of BH/Stt over the entire superconductor.
the s-n interphase boundary.

define

t F.

London

detailed field

pp. 125-130) has shown that taking into account the


penetration leads to the condition A' > A.
([2],

field

equal to

H(z)-Hc
-MH = -

intimately related to the Pippard range of co-

herence. Figure 16 shows the variation of the order parameter

ducting phase the increase in the free energy in the region where
is changing is given by V.6; in addition there is a reduction of the

energy due to the penetration of the

order of magnitude larger than the penetration depth.

The surface energy

(VIA)

The condition for the Meissner effect is that f > A, i.e. that A > 0.
The Ginzburg-Landau theory was formulated so as to lead explicitly to the existence of a surface energy, which arises as in the
Pippard approach from the gradual variation of the order parameter
*p over a finite distance, from the zero value in the normal region to
its full equilibrium value in the superconducting domain. Again the
surface energy is that amount which is needed to equate the energies
= c In the superconof the two phases in equilibrium, with
e

(VI.5)

If,-

4tt
is the value of the penetrated field at any point inside the
superconducting region. Thus the surface energy is given by the
integral of the difference between the superconducting and normal

where H(z)

free energies

over the entire superconducting half-space


00

G s {H,z)-

H:

H(z)Hc H\
.

4tt

+-P-G

/I

(0)--^

(VI.6)

The surface energy

Superconductivity

58

tent,

which gives for A

59

and ultimately to become negative. This has indeed been inferred

by Pippard (1955) and by Doidge (1956) from their studies of flux


trapping and the superconducting transition in dilute solid solutions
of indium in

Direct measurements of J in such alloys by Davies

tin.

(1960) has demonstrated

CO

its

decrease with shortening

/,

and Wipf
work

(1961) has traced this decrease to actual negative values. All the

where A
ductor,

is

the empirical penetration depth into a bulk supercon-

and k the dimensionless parameter defined

in the previous

chapter. This equation requires numerical integration.

For k <^

A becomes negative at a critical concentration of


approximately 2-5 atomic per cent of indium in tin.

cited indicates that

Changes of A with decreasing mean

it

reduces to

A =

1-89-

(VI.8)

The

free path also follow

numerical integration of VI.7, which yields that

is thus, according to the G-L


10A for most pure elements. The
the G-L model and Pippard's range of

thickness of the transition layer

This prediction

is

>

A<

from the

for
(VI. 10)

1/V2.

in good agreement with the work on tin-indium


Chambers (1956) found that the addition of 2-5

theory, of the order of A /k

alloys just cited.

intimate relation between

atomic per cent of indium to tin about doubles the penetration depth

coherence
first

is

shown by Gorkov's

principles.

terms of the

He

finds

critical

velocity of the metal.

derivation of

G-Ll and G-L2 from

an expression for the

G-L parameter k in
momentum and

temperature and the Fermi

Using equations IV. 17 and 111.13,

compared

to pure tin, so that according to the defining equation k

should be increased by a factor of approximately four. This would

make k

0-6,

which

close to the theoretical value of 0-707.

The surface energy and the intermediate state


Chapter II mentioned that a superconducting specimen with a demag-

6.2.

A
0-96

?-

(VI.9)

50

G-L

theory are of comparable

size.

netization coefficient

magnetic

Comparing VI.8 and VI.9 shows that, as expected, the Pippard range
of coherence g and the surface energy parameter A as derived from
In short, both approaches

necessarily lead to a positive surface energy because both require that

field

e satisfies

stantiated the suggestion of

Landau

It therefore also follows from both theories that the surface energy
must decrease and may even become negative when the range of coherence decreases and the penetration depth increases. Equations
IV.12 and IV.21 show that this is just what happens to A and to when

one

would, therefore, expect A to decrease with increasing impurity con-

c.

All

(1937, 1943) of a laminar struc-

normal layers

approaches

B = Hc

c.

while

Landau

B=

Clearly the width of the laminae

is

further suggested that in

in the superconducting ones.

strongly influenced by the

nitude of the interphase surface energy A. Indeed

field.

free path of the superconductor decreases. In alloys

-D)HC <HC < H

normal and superconducting layers. The thickness


of the normal layer grows at the expense of the superconducting one
as the external field

mean

(1

ture of alternating

the

length comparable to the difference between this distance and the

the inequality

experiments on the detailed structure of this state have generally sub-

the characteristic superconducting order parameter vary over a finite

penetration depth of an external magnetic

D is in the intermediate state when the external

distance. Both, therefore, obtain a net surface energy parameter of

the

is

this simplifies

to

the

as

for an infinite plate of thickness


(>

1),

the

sum a

Landau

mag-

finds that

L oriented perpendicularly to the field

of the thickness of the superconducting layer, a s ,

and that of the normal one,

a, is given

by

LA
(VI.11)

60

The surface energy

Superconductivity

where

is

a complicated function of the ratio of the external to the

e jHc Numerical values for Y(HJH have been calcuby Lifshitz and Sharvin (1951). A typical result is a value a 1 -4
= 0-8. A similar equation has also been
for L = 1 cm and
C
derived by Kuper (1951), who predicts numerical values which are
smaller by a factor of two or three. Typical experimental results fall
in between these predictions.
These results have been obtained by a variety of methods, all
making use of the fact that in the intermediate state lines of flux pass
only through the normal laminae, and emerge from the specimen
wherever these laminae end on the surface. A number of authors
(Meshkovskii and Shalnikov, 1947; Shiffman, 1960, 1961) have

critical field

lated

mm

HJH

i\,\t\

e2&

&&

The

61

series

superconducting laminae. The gradual shrinking of these areas with

and the corresponding growth of the light, normal


The domains show a peculiar type of
corrugation, not predicted by the Landau model, and adding to the
surface to volume ratio of the laminae.
increasing field

regions

is

clearly visible.

Phase nucleation and propagation


H. London (1935) pointed out that the existence of a positive surface
energy at the interphase boundary must under suitable conditions
give rise to phenomena analogous to superheating and supercooling
6.3.

in the

i\i\i\i\

shows a

of photographs obtained by Faber


(1958) with superconducting tin powder on an aluminium plate,
taken with increasing external field oriented perpendicularly. The
dark areas are covered with powder and are therefore the ends of the
frontispiece

more familiar phase transitions. In

phase transition cannot exist at

all if

fact a stable nucleus for the

the surface energy

is

everywhere

Indeed there are many experimental observations that when


a specimen is placed in a greater than critical magnetic field which is
positive.

then reduced, the normal phase persists in

fields less

the superconducting equivalent of supercooling.


Fig. 17

tion curve illustrating this

'supercooling'

passed very fine bismuth wire probes across the surface of a specimen,
and observed the magnetoresistive fluctuations in the probe resistance

when passing from

and cluster on the ends of the superconducting laminae, as shown

schematically in Figure 17; the latter will be attracted by flux and

move onto

the ends of the normal laminae.

The

resulting

powder

patterns can be easily seen and photographed. Another optical

method

is

shown

c.

This

is

in Figure 18.

The degree of

characterized by the parameter Si

= H IHc
t

this

or by

the parameter
<j>,

m 1-af - {Hl-Hf)IHl

(VI.12)

the end of a normal lamina to that of a super-

conducting one. Others have spread on the surface of a flat specimen


fine powder, superconducting (Schawlow et al, 1954; Schawlow,
1956; Faber, 1958; Haenssler and Rinderer, 1960) or ferromagnetic
(Balashova and Sharvin, 1956; Sharvin, 1960). The former will shun
flux

is

than

A typical magnetiza-

consists of placing a thin sheet of magneto-optic glass (for

tin, 5/ is commonly of the order of 0-9; in aluminium the degree


of supercooling is usually much larger, and values of St as low as 002

For

have been observed.


Superheating is the name given to the persistence of the supercon.

normal phase, probably because of large local field values resulting


from demagnetization effects. Centre portions of long tin rods could

made to superheat to S = 1-17.


Much information on the nucleation

example, cerium phosphate glass) on the specimen surface, and


observing the reflection of polarized light (P. B. Alers, 1957, 1959;

be

De

and on

Sorbo, 1960, 1961).

fields above
This is very rarely observed.
c
Garfunkel and Serin (1952) have shown that this is so because the
ends of any conventional specimen cannot resist the initiation of the

ducting phase at

its

of the superconducting phase

relation to the surface energy has been obtained by

Faber

62

Superconductivity

The surface energy

on supercooling in tin
and aluminium. His technique consisted of winding on a long cylindrical specimen several small, spaced coils the field of which could be
made to add or to subtract from a field produced by a large solenoid
surrounding the entire sample. With the sample normal, the field of
the solenoid could be lowered to some value between H, and
c and
the field could then be lowered locally by a suitably directed current
through one of the smaller coils. The superconducting phase then
(1952, 1955, 1957) in a series of measurements

nucleated in the portion of the sample under the

coil,

and spread

63

some of which exist in even the purest specimens. This


supported by Faber's finding that any handling of the specimens
between measurements could change the location and effectiveness of

of local strain,
is

the nucleation centres. Strained regions probably contain a high

density of dislocations.

By correlating the size of the nucleating field t with the time it


took to be effective, Faber could deduce the depth of the nucleating
flaw below the sample surface, and found this always to be between
-4
-3
10
and 10
cm. Etching down the surface to this depth would
uncover further flaws extending to a similar depth. It is thus reason-4-10~ 3
cm as being the approximate size of the
nucleating flaws. At temperatures well below Tc this length is considerably bigger than the width of the interphase boundary, and one
able to take 10

can therefore imagine such a flaw to consist of a region of negative


surface energy surrounded by a shell across which the surface energy
increases to the

(1952) has

normal

shown

positive value of the bulk material.

that there

is

Faber

a potential barrier against the further

growth of this nucleus until one has reached a degree of supercooling


such that
,
<f>,

Hc

H,

Applied Field

He

Fig. 18

where A

is

flaw size,

tected by pick-up coils distributed along the specimen.

At a

given temperature the degree of supercooling varied confrom point to point in a given specimen but at a given point

siderably

frequently remained reproducible even when in between measurements the specimen was warmed to room temperature. This indicated
that nucleation must occur at particular spots, some of which promote nucleation more effectively than others. As the surface energy
can be lowered and may even become negative due to strain, it is
reasonable to assume that the spots favouring nucleation are regions

(VI. 13)

the surface energy parameter, r a length of the order of the

and

//

a small constant determined by the flaw's shape and

demagnetization factor.
perature variation of

rapidly throughout the sample. In this fashion supercooling could be


studied at different portions of the sample. The transition was de-

-+n,

</>,

The measurements in fact show that the temis very much like that of A, as determined

from other experiments.


Both Faber (see Faber and Pippard, 1955b) and Cochran et al.
(1958) found that supercooling was much enhanced after a specimen
had been placed temporarily in a field much higher than the bulk
critical value. This shows that certain superconducting nuclei can be
quenched only by such a high field and supports much other evidence
that in a non-ideal specimen there can exist small regions of high
strain which remain superconducting in very high fields.
By means of a series of pick-up coils along his specimens, Faber
(1954) was able to observe the propagation of the superconducting
phase once the transition had been initiated at some nucleus. From

CHRIST'S COLLEGE
IDDADV
I

Superconductivity

The surface energy

growth of the superconducting phase


The nucleus, which is always near
the surface, first expands to form an annular sheath around the specimen. This sheath then spreads along the entire length of the specimen
with a velocity of the order of 10 cm/sec, and finally the superconducting phase spreads inwards to fill the entire sample.
The growth of a superconducting region is limited principally by
the interphase surface energy on the one hand, and by eddy current
damping on the other. If there were no surface energy, the superconducting phase could propagate by means of very thin filaments
which displace no magnetic flux and therefore create no retarding

The second of these functions appears to give a good fit to various


results for tin over a rather wide range of temperature, but Faber's

induced currents. For a sheath of

equation IV. 17, as well as empirical values for A

64

his results he infers that the

occurs in a series of distinct stages.

finite thickness, on the other hand,


which propagates in the presence of an external field c eddy currents
are generated, and the magnetic energy gained in the phase transition
is balanced by the unfavourable surface energy as well as the eddy
current joule heating. Faber (1954, 1955) has shown that the resulting
velocity of propagation for very pure specimens under optimum conditions is given by

A(I/o)A- (lHc

aluminium data can be represented only by the first of these. There


seems to be a definite difference in the temperature dependence of the
surface energy for these two metals which is at present not understood.

The uncertainty
The

table

below

lists

for tin

and for

Element

way by Faber, as well as those


ways by Davies (1960), Sharvin (1960), and
Shiffman (1 960), can be fitted by a number of empirical functions of
temperature. According to the G-L theory, the surface energy should
have the same temperature variation as A at least very near Tc where
k is independent of temperature. Hence one would expect
different

= J(0)(l-/4 )- ,/2

(VI.15)

Aluminium
Indium

18

Tin

0-50

16
4-4

0-64

2-3

2-3

0-51

Supercooling in ideal specimens

Near

Tc A becomes
,

large,

and the flaws

nucleation centres. Measuring

some information on supercooling in


(1957) has found for aluminium, S,
al.

lose their effectiveness as

in this region can, therefore, give

unflawed material. Faber

ideal,

= 0036,

1.

(VI.150

for In 016,

and

for

Sn

(1958) for aluminium are in reasonable

agreement. These results can be compared with theoretical predictions arising from the G-L model. Equation V.9' has an interesting

consequence with regard to the normal phase. One would expect that
with
described by the equation would be
e S*
c , the half-space
= 0. This is indeed a solution, but the equation
entirely normal, with

H H

</<

is

also satisfied by a second solution with

this solution

#/>

<^ 1,

so that H(z)

ifj

# 0.

Assuming that

for

everywhere, and remem-

becomes
for/

(0).

3-4

bering that in the geometry chosen A(z)

which can also be written

j-o--w

number

10 5 Ao(0)
(cm)

10 5
(cm)

10 <d(0)
(cm)

0-23 values of Cochran et

values of A(T) obtained in this

A(t)

of course

the best available values of ^(0) for a

The

of metals, from a comparison of all available experimental data. Also


listed are values of > the range of coherence, as calculated from

6.4.

where / is the electronic mean free path in the normal phase, a the
normal electrical conductivity, and A is a constant of the specimen.
By measuring the temperature variation of v, Faber has used this

measured in

dependence of

(VI. 14)

equation to obtain the temperature variation of


aluminium.

in the temperature

introduces a degree of doubt about the extrapolated value at 0K.

-H ]/H y
e

65

7ES

= H(z)z,

the equation

Superconductivity

66

The surface energy

This has the form of a wave equation for a harmonic oscillator, which

known

is

(which

is

which vanish for z= co


the required boundary condition for the normal phase) if
to have periodic solutions

tp

V2^(/+i), =

67

For superconductors with a dimension small compared to X /k


the order parameter is essentially constant throughout and one can
solve the G-L equation with the simplifying assumption k a 0. The
,

critical fields

of supercooling are then given by

0,1,2,...

tic

In other words, for any value of

k,

conductor becomes unstable with regard to the formation of laminae


of superconducting material when

HJH =
C

K/(n +

of which the highest value, with n

J)

< B and

the field

IHc = V(2)k.

c2

is

c2 is

2V5^H
a

V8jHc

that of

and

(VI. 18)
1

/V2.

most pure superconductors,


field to which the

then the lowest

normal phase can persist in a metastable fashion.


c2 is thus the
lower limit to which an ideal superconductor can be supercooled,
and therefore in the region very near Tc one would expect the experimental value of S t to equal \/(2) k (Ginzburg, 1956, 1958a; Gor'kov,
1959b,

= 0, is

c2

= V6-/Z,

for a slab of thickness la,

(VI. 17)

V2,

A distinction must now be made as to whether k < / V2 or k>

C2

for a sphere of radius a,

c2

In the former case, which

the normal phase of the super-

c2

for a cylinder of radius r (Ginzburg, 1958a).

Hc2

For all

these geometries

decreases with increasing specimen size, approaching

tonically the value given

mono-

by VI. 18, which depends only on the value

of k characteristic of the material.

The compatibility of the G-L theory with the Pippard range of cosize, or mean free

herence under those circumstances of temperature,

c).

The values of k calculated in this fashion from Faber's measurements of St are: 01 64 for tin, 0112 for indium, and 0026 for aluminium. The first two of these agree very well with k values deduced
from experimental penetration depths. In aluminium the lack of
agreement is probably due to the appearance of non-local effects very
close to Tc Ginzburg (1958b) has noted that this is more likely to

path which eliminate the need for a non-local electromagnetic formulation is brought out once again by the similarity of VI. 18 with the
corresponding expression derived by Pippard (1955). He finds, also

by minimizing the free energy, that


2 3A o
Hc2 - V jn
n

c-

(VI. 19)

invalidate calculations involving the penetration depth than those

regarding the surface energy and supercooling. Non-local effects

become important
when

for the former

when >

for the latter only

This differs from VI. 18 only by a numerical factor of order unity


since

6.5.

Thus K-values calculated from supercooling data are probably


most reliable, except for the effect discussed in section 6.7.

the

/.

Superconductors of the second kind

According to equation VI. 10, the surface energy becomes negative


when k > l/\/2. A similar conclusion follows from the Pippard nonlocal model when A > $ (equation VI. 4: seeDoidge, 1956). The existence of a positive surface energy was shown to be necessary for much

The surface energy

Superconductivity

68

of the magnetic behaviour usually found in superconductors.


therefore, not surprising that superconductors in

which

this

It is,

energy

is

negative display quite different characteristics. They are accordingly


called superconductors of the second kind.

For a bulk specimen of such a superconductor the volume

free

energy in the superconducting phase remains lower than that of the

normal one in external fields up to the thermodynamic value


c
defined by equation II.4. The negative surface energy, however,
makes it energetically favourable for interphase boundaries to appear
at field lower than
c and for superconducting regions to persist to

In the limit k

1/V2, H =H

cl

= Hc2 Goodman
.

69
(1962a) has inter-

polated between the latter value and those given by VI.21 to get a

k A numerical solution has


cl /Hc for all
been obtained by Harden and Arp (1963).
The magnetization curve predicted by the Ginzburg-LandauAbrikosov (G-L- A) model can be compared with experiment, as it is

graphical representation of

possible to determine the value of k for a specimen by independent


measurement. Gor'kov (1959) has derived an expression for k valid

when

the electronic

mean free path

is

much smaller than

the intrinsic

H
H Goodman (1961) has shown that this can already
,

fields

higher than

be deduced from the London model by the single addition of a


negative surface energy term.

The details of the behaviour of superconductors of the second


kind can be deduced from the G-L theory, which is equally valid for
k > \/\/2 as for k <
section

still

/\/2. In particular, the analysis of the preceding

holds; that

is,

the

normal phase has a

stability limit at

H given by equation VI. 18, which shows that for k > 1/V2,
H > H Abrikosov (1957) has used the G-L equations to analyze
field

c2

c2

in

some

detail the

magnetic behaviour of superconductors of the


in the magnetization

second kind, and finds the features indicated

curve shown in Figure 19 for a cylindrical sample parallel to an

Hcl

There is a complete Meissner effect only up to


= ci < c at which point the magnetization changes with
e
infinite slope. For values of k not much larger than 1 /V2, Abrikosov
predicts in fact a discontinuity. At somewhat higher field, the magnetization approaches zero linearly, with a slope
external field

Hc

H c2

Applied Field H e

Fig. 19

coherence length This was shown by


the convenient form
.

-4tt^=

1-18/(2^-1),

(VI.20)

and vanishes

c2

= V(2)kHc

The magnetization curve should be

ko
(VI. 18)

Abrikosov canvalues of k; for

fully reversible.

not solve the equations determining


1

(1962a) to have

+7-5xloy /2 p.

(VT.22)

entirely at

BL =

k>

Goodman

cl

for all

is

the parameter for the pure substance,

heat constant, in erg

ohm-cm. For

V(2)kHc i/Hc =

In k+ 0-08

(VI.21)

deg

-2
,

y the Sommerfeld specific


and p the residual resistivity in

tin this predicts quite closely the resistivity at

which the

surface energy becomes negative (Chambers, 1956).

Using

he obtains

cm -3

model

this equation,

Goodman

satisfactorily explains the

(1962a) has shown that the G-L-A


magnetic behaviour of substances

70

Superconductivity

The surface energy

Ta-Nb alloys investigated by Calverley and Rose-Innes


and his own U-Mo alloys (Goodman et ai, 1960).
Further-

such as
(1960)

more, recent magnetization measurements on Pb-Tl single


crystals
(Bon Mardion et ai, 1962), indicate a considerable degree

of reverDetailed quantitative verification of the G-L-A


magnetization curves, as is possible only near T was
provided by Kinsel et al.
c
(1962), who used In-Bi specimens to compare values
of k calculated
from equations VI. 18, VI.20, VI.22, and from
Harden and Arp's
values of ci /Hc The different values of k for a given
specimen agree
to within a few per cent. Similar agreement can
also be deduced from
the results of Stout and Guttman (1952)
on In-Tl alloys. The G-L-A
model is thus well substantiated.
sibility.

The negative

surface energy need not be due to a short mean


free
it is possible for the coherence
length to be shorter
than the penetration depth, even in a pure superconductor;
this is
path. In principle,

most

likely in superconductors with a high


Tc (cf. equationIV.17).
Indeed, Stromberg and Swenson (1962) have found
that the magnetization curve of very highly purified niobium
is that of a superconductor of the second kind, with a value of
cl and
c2 corresponding
to k ~ 1 1 This conclusion is consistent also
with the results of Autler
(1962) as well as of Goedemoed et al. (1963).
Kinsel et al. (1963) have found with their
In-Bi alloys that the
effective value of k as defined by equation VI.
18 rises gradually as
the temperature decreases below T increasing
by about 25 per cent
c,
as T approaches 0K. This agrees with the
calculations of Gor'kov
(1960). The experiments further show that
at any temperature

nor the size

effects to

be discussed

in the

71

next chapter can increase the

of a superconductor without limit. Both Clogston (1962)


and Chandrasekhar (1962) have pointed out independently that in
sufficiently high fields it is no longer correct to assume that the free
energy of the normal phase is independent of field. With a finite
paramagnetic susceptibility Xp (which was ignored in deriving equa2
tion II.4), this free energy is, in fact, lowered by an amount \X.H
critical field

Thus, in sufficiently high

fields, this

alone could already bring about

from the superconducting to the normal phase. The limit


on the critical field imposed by this mechanism is estimated to be two
or three hundred K gauss, and this is consistent with the results of
Berlincourt and Hake (1962).
a transition

The mixed state or Shubnikov phase


The magnetization curve of type II superconductors clearly shows
that for
< e < cl the material is neither in the usual supercX
6.6.

H H

conducting nor
region the

in the

normal phase. Abrikosov (1957) has called this

mixed state, and De Gennes ([14]) has suggested naming it


honouring the scientist who first suggested the

the Shubnikov phase,

= 1/V2

continues to be the critical value for the


onset of superThus a specimen with k ~ 0-65 at its transition
temperature is there a superconductor of the first kind,
but becomes

conductivity.

one

of the second kind at that temperature at which k


reaches the

critical

value.

The temperature dependent increase in k leads to a


corresponding
decrease of the surface energy. Specimens for which
k goes through
the value 1/V2 at some temperature are those for
which at that temperature the surface energy changes from being
positive to being
negative, as has been observed for indium alloys
by Kinsel
et al.

There

is

964).

reason to believe that neither a negative surface


energy

fundamental nature of type

II

superconductivity (Shubnikov et

al.,

1937).
It is

evident from the importance of the negative surface energy that

mixed

specimen must contain as large an area of


is compatible with a minimum of normal
volume. This could be brought about by a division of the material
into a large number of very thin normal and superconducting sheets
in the

state the

interphase surfaces as

or laminae

G-L-A

(Goodman,

1961, 1964; Gorter, 1964). According to the

mixed state consists of a regular array


of normal filaments of negligible thickness which are arranged parallel
theory, however, the

and are surrounded by superconducting material.


At the normal filaments the superconducting order parameter
vanishes, and then rises from these linearly with distance. It reaches
to the external field

its

maximum

value as quickly as possible, that

the order of .
filaments,

means

and

falls off

over a distance of
at the

over a distance of the order of A

that the field decreases to zero only

at distances at least of the order of A. This


6

is

The magnetic field has a maximum value


if

>

normal
.

This

the filaments are spaced

mixed

state structure

can

72

Superconductivity

The surface energy

be shown to have a lower energy than any laminar arrangement


([14], p. 111.81).

One can

that the order parameter

thus think of the mixed state as

material were pierced by a

if

the superconducting

number of infinitesimally

thin filamentary

holes, regularly spaced parallel to the external field

containing magnetic

flux.

From the

and thus each

discussion in Chapter II

it

there-

fore follows that the total flux associated with each normal thread

quantized in units of

is

This flux does not penetrate far into the


superconducting material because of superconducting currents circu<f>

lating in planes perpendicular to the filament. This creates a vortex


line of superconducting pairs along each normal thread, in striking
analogy to the vortices existing in liquid Helium II (Rayfield and

Rcif, 1964).

The

and the currents associated with an isolated vortex line


extend over a distance of about A. The interaction between two
flux

vortex lines can thus be appreciable only at distances less than

A.

This means that when the formation of vortex lines becomes energeti= ci they can essentially immediately achieve
cally favourable at

H H

a density corresponding to a separation of about A without creating


much interaction energy ([14], pp. III. 74ff.). This causes the abrupt
decrease of the magnetization at
cl predicted by Abrikosov and

verified experimentally. It is not certain, however, whether the


magnetization actually decreases discontinuously at this field or
whether it merely drops with an infinite slope. The former would

correspond to a first order transition with a latent heat, the


only to an infinity in the specific heat.

With the external

field increasing

beyond

cX

latter

more and more

vortex lines are formed until their spacing approaches

Mh

finite

except along the

normal filaments of

negligible

volume. Thus the material can still be considered as entirely superconducting. Abrikosov (1957) showed in fact that in the mixed state

one can characterize the material by a mean square order parameter

and that near Hc2 this varied linearly with the magnetization.
The correctness of this and therefore the validity of the vortex
structure has been substantiated by measurements of the specific heat
(Morin, et al., 1962; Goodman, 1962b; Hake, 1964; Hake and
Brammer, 1964) and of the thermal conductivity (Dubeck et al.,
y*2 ,

1962, 1964).

De Gennes and his collaborators

(cf. [14],

Vol. II) have studied the

properties of an isolated vortex line, as well as the interaction between


lines. This leads to possible collective vibrational modes (De
Gennes and Matricon, 1 964), as well as to a surface barrier inhibiting
the motion of lines into or out of the superconducting material (Bean
and Livingston, 1963; [14], p. 111.85), De Gennes and Matricon

such

(1964) have also suggested the possibility of investigating the vortex

of the mixed state by slow neutron diffraction. Preliminary results have recently been reported (Cribier et al., 1 964).
line structure

In an ideal type

II

superconductor, homogeneous and devoid of

lattice imperfection, the

vortex lines would be pushed out of the

material by the Lorentz force

if

the specimen carried any current

In any actual material,


motion of the lines is inhibited by defects and inhomogeneities
which form potential barriers by which lines the are pinned. Anderson
at right angles to the field (Gorter, 1962a, b).

the

(1963).

c2

964) and Matricon (1 964) have shown that a triangular


somewhat lower energy throughout the mixed state. This
changes the coefficient in equation VI.20 from 118 to 116.
al. (1

everywhere

remain uniform, so that bundles of lines move together.


This vortex or flux creep has been further discussed by Friedel et al.

Kleiner et

is

lines tends to

array has a

centre of the vortices, which are

(1963) has investigated the thermally activated 'creep' of lines at low


current densities, and has shown that on a local scale the density of

(Abrikosov, 1 957). According to Abrikosov, the vortex


lines form a square array at all fields except very near
cU but
as //nears

73

A fundamental feature of the vortex structure of the mixed state is

With

increasing current densities to creep changes into a

viscous flow of the lines, giving rise to resistive

phenomena (Anderson
this has been done

and Kim, 1964). Extensive experimental work on


by Kim era/. (1963, 1964).

Tinkham

(1963, 1964) has

like that of the

mixed

shown that

a quantized vortex structure

state occurs even in

pure films

if

they are very

74

Superconductivity

and placed in a perpendicular external field. This is in agreement


with magnetization measurements on such films by Chang et al. ( 963)
and penetration depth and critical field data of Mercereau and Crane
(1963). Guyon et al (1963) have investigated the dependence of the
critical field on thickness. For thin films so narrow as to contain
only
a single row of vortices Parks and Mochel (1964) calculated that the
free energy should have a minimum at values of the perpendicular
external field at which the vortex diameter just equals the film width.
thin

At Tc

should result in a corresponding minimum of the film


resistance. They have observed such minima and take this as direct
evidence for the existence of quantized vortices.
6.7.

this

Surface Superconductivity

As mentioned in Chapter V,

the

boundary condition applicable to

G-L order parameter W is that its derivative vanish. Saint-James


and De Gennes (1963) have shown that in an external field parallel to
the

CHAPTER

The Low Frequency Magnetic


Behaviour of Small Specimens
7.1. Increase in critical field

When one of the dimensions of a superconducting specimen becomes


comparable to the penetration depth, its critical magnetic field becomes much higher than that of a bulk sample of the same material
at the same temperature. This follows already from the basic GorterCasimir thermodynamic description, according to which the free
energy difference per unit volume between the superconducting and
normal phases is

Hi
Gn (0)-GM =

the surface this leads to the persistence of an outer superconducting

sheath up to a

c3

1-695

c2

The thickness of this sheath is of the order of .

In
Its existence, explicitly

an external

netization

field

M(He

by many experiments (see, for example, Hempstead and Kim,


Tomasch and Joseph, 1963), explains what had often been

a puzzling discrepancy between magnetic and resistive transitions.


The surface sheath exists also in Type I superconductors, but can be

only

= 2-40kHc

if

c2

so that

>

c.

As

Hc3 > H

c2

for k

= \/2kHc

>

it

follows

that

Under this condition,


way of obtaining k for Type I
materials (Strongin et al. 1964; Rosenblum and Cardona, 1964).
The existence of the surface sheath in Type I superconductors means
c3

a measurement of

c3

/Hc

is in

0-42.

fact a

supercooling experiments are carried out on cylindrical


in a longitudinal field, as is usually the case, the ideal lower
limit for super cooling is
c2 rather than
c2 (cf. section 6.4). Thus the

that

if

samples

experimental value of S, should be set equal to 1-695V(2)k and the


values of k thus calculated are therefore correspondingly reduced.

a superconductor acquires an effective

mag-

and becomes normal when

verified

detected

(VII.l)

field

H
1963;

VII

lie

M{H )dHe = -*
e

The

(VII.2)

077

under the magnetization curve, and it was


that for any ellipsoidal specimens VU.2
=
was satisfied when
Actually this is true only when one
e
c
neglects the penetration of the external field into the sample, which
lowers the effective magnetization and the susceptibility of the
sample, as shown by equations IV.2 and IV.3. The susceptibility
determines the initial slope of the magnetization curve; a lower x
means that the curve has to go to a higher critical field to satisfy
equation VTI.2. Clearly, assuming this curve to remain linear with
slope x right up to a critical field
s
integral is the area

pointed out in Chapter

ni

"I
Hi

Xo

(VII.3)

X
75

76

Low frequency magnetic behaviour of small specimens

Superconductivity

and

CVII.4)

11

Hauser and Helfand, 1962), and have been used by Lutes (1957) in the
interpretation of his measurements of the critical field enhancement
in tin whiskers.

v^3

=
7T
using the

for a <^ Aq,

It is possible to relate

(VII.5)

the thin film critical field to the basic super-

London equation to evaluate a =

IV.3 (Ginzburg, 1 945


[11 p. 172). Similar expressions can be derived for spherical and
cylindrical samples. The resulting equations agree well with the fre$ in

of the bulk material. The penetration


conducting parameters anti
depth appearing in VII.8 should be given by 1V.20, in which (/) is

determined by IV. 13 with the film thickness taken as the effective

mean

free

path (Tinkham, 1958). In the limit a

quently observed enhancement of the

critical field in small specimens


uses for the penetration depth A the appropriate Pippard
value as calculated from IV.20. This is a good example of how ex-

<

this yields

(Alloy*

(VII.8 ')

when one

pressions derived from the


fied value

The

London model can be used with

the modi-

of A (Tinkham, 1958).

field

enhancement calculated from the Ginzburg-Landau

theory leads to nearly identical results. The essential difference is that


because of the additional terms V.6 in the G-L free energy of the
superconductor, the penetration depth increases in the presence of an
external field (see equation V.16), so that the critical field for small

samples becomes even higher. For thin films of thickness 2a the


critical field is

(Douglass and Blumberg, 1962). The use of the thin film susceptibility
as derived by Schrieffer (1957) with non-local electrodynamics leads
to

20-40 per cent higher values of the numerical constant (Ferrell and

Glick, 1962;Toxen, 1962).

7.2.

High

The

size effect

field

threads and superconducting magnets

on the critical field is particularly striking in experiments using extremely thin evaporated films. In their experiment on
the Knight shift in tin, Androes and Knight (1961) used films of thickGinsberg and Tinkham
ness a 100 A and found
c (0) 25 kgauss.
(1960) saw no effect on the superconducting properties of their
10-20 A lead film in a field of 8 kgauss.
The equivalent of small superconducting specimens can exist also
in bulk material. In an inhomogeneous specimen there will be local
variation of the surface energy due to varying strain or to varying

sir

(VII.6)

la

where /(*) is the same function of k which appears


and is very small for small values of k.
For very thin films, a < A G-L find that

in

equation V. 1 6,

electronic

mean

free path. If locally the surface energy

lower than the value elsewhere,


this

which for very small k reduces

ti c

field

to

even when the surrounding material has become normal


1 955). Under these conditions one can thus have a situation

(Pippard,

quite analogous to that of small specimens: small superconducting

(VII.8)

the same expression which for thicker films gives the supercooling field
c2 Expressions similar to VII.6 and VII.8 have also
been derived for spheres and wires (Silin, 1951; Ginzburg, 1958a;
is

is sufficiently

may be energetically favourable for

region to remain superconducting in the presence of an external

regions exist in a matrix of normal material (Gorter, 1935; Mendels-

sohn, 1935;

This

it

Shaw and Mapother,

1960). If their dimensions are

small compared to the penetration depth, the critical


regions will be correspondingly raised, and

it is

field

of these

known (Faber and

Pippard, 1955b; Cochran et al, 1958) that such regions can persist in

78

Low frequency magnetic behaviour of small specimens

Superconductivity

high

fields.

many

In

instances these regions are threads which can

form continuous superconducting paths from one end of the

men

to the other, resulting in a resistive transition

speci-

much broader and

extending to much higher fields than the magnetic one (Doidge, 1 956).
are, of course, likely to touch each other in many places,

The threads

what Mendelssohn (1935)

resulting in

called a superconducting

sponge. The multiple connectivity of such a structure generally leads


to highly irreversible magnetic transition with almost total flux trapping. Bean (1962) has used a simplified model with which to calculate
the magnetization curve of such a sponge. He has confirmed some
features of this

model with an

artificial

79

which has been verified by Zavaritskii (1951, 1952). Note that as the
penetration depth is inversely proportional to 0, A(//) for thin films
even in fairly small fields (Douglass, 1961c).
is much larger than A
Douglass (1961a) has pointed out that because of the proportionality

of the energy gap to

/r,

by Gorkov (1959, 1960),


field dependence of the
Thus one can write

as derived

equations VII.9 and VII. 10 represent the

energy gap

in sufficiently thin films.

e\Hs)

for 2a

< V(5)A

(VII. 11)

filamentary superconductor

made by forcing mercury into the pores of Vycor glass (Bean et a/.,
1962). The possible relevance of this to superconducting magnet wire
will be discussed in

7.3. Variation

Chapter XIII.

of the order parameter and the energy gap with magnetic

field

From equations V.G-L1 and V.G-L2 one can also calculate the variation of the order parameter

*fi inside the thin films. For thicknesses 2a


very small compared to the width of the transition layer Xq/k, or in
the equivalent Pippard terms for 2a < g if; can be considered con,

stant,

and one can take k

H
"6

For very thin

K 0.

This leads to (Ginzburg, 1958a)

\"c/ OA 6J/ L

30 \H
3oUc7\V

X o/)
A

(VII.9)

films VII.8 applies, so that

For thicker

films, VII.9

G-L2 must be
<l<i

For such films, therefore, i/j(Hs ) = 0, which means that the transition
into the normal state is of second order, without a latent heat and with
a discontinuity only in the specific heat, and not in the entropy. There
can be no supercooling, and therefore, no hysteresis. For thicker films
and bulk samples the transition in an external field, as discussed in
Chapter II, is always of first order. The critical thickness below which
there

is

a second order transition

2a

is

V(5)A

energy gap at

H =H
e

do not apply, and G-Ll and


The resulting variation of the

VII. 10

asa function of film thickness has been

calcu-

Douglass (1961a). It is displayed by the curve in Figure 20.


The points are gap values which Douglass (1961b) obtained from tunnelling experiments (see Section 10.6). Similar results have been found
lated by

by Giaever and Megerle (1961), also by means of the tunnel effect, as


and Tinkham (1961) with thermal conductivity
measurements. With e s the empirical variation of the energy
well as by Morris

H H

gap with
,

and

solved numerically.

field closely

Ginzburg-Landau-Gorkov prebelow Tc In such high fields the

agrees with the

dictions even at temperatures well

80

Superconductivity

order parameter

then small enough to

is

make tenable

the basic

G-L

assumptions as well as Gorkov's identification of the energy gap with


W, Tinkham (1962) has proposed ways in which the G-L equations
can be extended to give agreement also with low field results over a
wide range of temperature. The limitations of these equations in this
region have been discussed by Meservey and Douglass (1964).
Bardeen (1962) has calculated the critical field and critical current
for thin films

on the

basis of the

BCS theory. At higher temperatures

CHAPTER

The Isotope
8.1.

VIII

Effect

Discovery and theoretical considerations

The various phenomenological treatments based on

the empirical

his results generally confirm the predictions of the

Ginzburg-Landau
theory, including the vanishing of the energy gap and a resulting

characteristics of a superconductor provide an astonishingly complete macroscopic description of the superconducting phase. How-

second-order transition at the critical field in sufficiently thin films.


At much lower temperatures, however, below about TJ3, Bardeen

ever, they do not give any clear indications as to the microscopic nature

any thickness the energy gap remains finite and the


transition a first-order one. However, Maki (1963) as well as Nambu

One of the first such clues arose through the simultaneous and independent discovery, in 1950, by Maxwell, and by Reynolds et al., that
the critical temperature of mercury isotopes depends on the isotopic

finds that for

and Tuan (1963) predict


second order at
verify this

down

all

to

that the phase transition should be of the


temperatures. Merservey and Douglass (1964)

= 0- 14.

of the phenomenon.

mass by the relation

Tc M a =

Ave. mass no.:

199.5

r-

constant,

(vm.i)

200.7 (nat)
202.0
203.3
J

4.20

4.00

Fig. 21
81

82

Superconductivity

The isotope

where Mis the isotopic mass and a m . This is illustrated in Figure


21,
showing the variation of threshold field near T for different isotopes.
c

The

effect has since also

been established in a number of other elements. The following table contains the most reliable experimental
values of the exponent a, together with quoted probable errors.
Element

Cd

0-51

Hg

0-504

Mo

0-33

Os
Pb

Rh
Ru

0-461 0.025

0-501
0-4

001

Zn

was recog-

8.2. Precise threshold field

Shaw
Hake

The

measurements

variation of critical temperature with isotopic

lished

by measuring the critical

field

mass was estab-

as a function of temperature,

etal., 1961
et al.,

1958

Gcballee/o/., 1961

Finnemoreand Mapother, 1962

0-505 0019
0-46 002
0-462 001
0-50 005
0-62 01

Tl

subtle nature of the pertinent electron-lattice interaction

Maxwell and Strongin, 1964

<01
<005

Sn

independently suggested just such a mechanism without knowing of


the experimental work. However, it took several more years until the

Olsen, 1963

Reynolds et al., 1951


Matthias et al., 1963
Hein and Gibson, 1964

0-21

clearly pointed out the direction in which a microscopic explanation


of the phenomenon had to be sought. In fact, Frohlich (1950) had

nized and a valid microscopic theory began to be developed.

Reference

010

83

effect

0-5

Maxwell, 1952a
al., 1952

Serin et

Lock

et al.,

1951

Maxwell, 1952b
Alekseevskii, 1953

Geballeand Matthias, 1964

In all the non-transition metals, with the exception of molybdenum,


the results are consistent with a = 1/2. However, small mass
differences and the possibility of impurity and strain effects limit the
experi-

mental

reliability,

as

is

made evident by the variations between differ-

ent measurements

on the same element. Thus one cannot rule out


deviations from the ideal value of a = which may be as high as
20
per cent in some cases. In view of recent theoretical work to
be discussed in Chapter XI,

it is

significant that the trend of the published

= \ is toward lower values. The situation in the


ruthenium and osmium, however, appears to be
different. This will be further discussed in Section
1 1.5.
The inference to be drawn from the dependence of T on the isotopic
c
mass is startling. A relation between the onset of superconductivity,
which is quite certainly an electronic process, and the isotopic
mass,
which affects only the phonon spectrum of the lattice, must mean that
deviations from a
transition metals

superconductivity
the electrons

is

very largely due to a strong interaction between


lattice. Thus the discovery of the isotope effect

and the

and then extrapolating

this to

zero

field.

Magnetic measurements of

course make use of the perfect diamagnetism of a superconductor, and

can be made in one of two ways either the change in flux through the
sample at the transition induces an e.m.f. in a pick-up coil which is
connected to a suitable galvanometer, or the changing susceptibility
of the sample is reflected in the change of the mutual inductance of
coaxial coils of which the sample forms part of the core. Either of
:

methods can be applied with great accuracy in spite of simple


and has the further advantage of measuring a bulk
property virtually unaffected by the possible presence of small regions
with different superconducting characteristics. By providing a
these

apparatus,

84

Superconductivity

The isotope

misleading short-circuiting path, such minor flaws can lead to very


when Tc is measured by observing the variation of

A polynomial which fits the data for all tin isotopes as found by

erratic results

et al. (1951) to within one-half

electrical resistance.

The

careful determination of critical field curves which arose as


almost a by-product of the work on the isotope effect established a

//

85

effect

of a per cent

Lock

is

= l-10720/ 2 -0-0944/ 4 + 0-3325/ 6 -01660/ 8

(VIII.3)

All measurements to date have indicated that to within the available


precision all isotopes of a given element follow the

polynomial.

One

also finds that

same

critical field

has the same mass dependence

4-LEAD

Tc This means that the superconducting condensation energy


Hq/Btt varies proportionally to the isotopic mass, and also that, as
shown by equations 11.15 or 11.16, the value of y is independent of
as

Fig. 23

number of interesting
of the reduced

isotopic mass.
characteristics. Figure

critical field

22 shows the variation


as function of t 2 m T 2 /T 2 for a
C

m Hc /H

number of tin isotopes measured by Serin et al. (1952). It is evident,


as was indicated earlier, that equation I.2a is only an
approximation]
and that a

better representation for

is

a polynomial

W) = - S of,
i

n=2

(vni.2)

It

has also been found that in going from one element to another,
show small but definite variations.

the reduced threshold field curves

For

all

elements there are deviations from a

strictly

parabolic varia-

by a similar small amount in one direction, but in the


case of lead and mercury by an amount in the opposite direction.
Figure 23 shows these deviations as a function of reduced temperature. It is important to emphasize the smallness of these deviations.
tion, generally

86

Superconductivity

so as not to allow them to obscure the basic


similarity of the superconducting behaviour of all elements in terms
of reduced co-ordinates.
This not only sanctions the continuing
discussion of superconductivity in general terms with only
occasional references to specific
elements, but also allows one to look for a
microscopic explanation
of superconductivity, which in first approximation
need not concern
itself with the distinctive characteristics
of individual elements, but
only takes account of general and common
features. The deviations
of the measured threshold fields from a simple
parabolic variation
must be, according to the thermodynamic
treatment developed in

Chapter n, correlated with the empirical deviations


of the specific
heat from the corresponding change as the
cube of the temperature
Serin (1955) showed this strikingly by plotting
both these deviations
on the same graph, using the best available data for tin.
This is shown
in Figure 24. Mapother (1959) has
since established the correlation
between the experimental non-parabolic
threshold fields and the
exponential variation of the specific heat.

CHAPTER

IX

Thermal Conductivity
9.1.

Low

temperature thermal conductivity

normal metals, heat


and by the quantized
In

is

carried both by the conduction electrons

lattice vibrations, the

thermal conductivity consists of the


Km

phonons. The

sum of these two

k en -r k

total

contributions
(IX.1)

where e and g denote the electrons and the lattice, respectively. The
is limited by two scattering mechanisms the
phonons and the lattice imperfections, and one can write at T< :
electronic conductivity

\\k cn

The

first

= aT 2 + Po /LT.

(1X.2)

of the terms on the right gives the

due

resistivity

and predominates

electron scattering by phonons,

to the

at higher tem-

peratures; the second that due to scattering by imperfections, which

becomes important below the temperature at which k en has a

maximum:

TL* =

(IX.3)

Po/2aL.

In these equations p is the residual electrical resistivity, L the Lorentz


-8
(2-44 xlO
watt-ohm/deg2 ) and a is a constant of the

number

material which

is

inversely proportional to

@2

Note that for a given

material the addition of impurities increases p and thus raises Tmax


In pure metals and dilute alloys, k en
k g ; it is only in metals con.

>

taining as

much

as several per cent impurities that the

two contri-

butions are of the same order of magnitude.

The

two-fluid

model allows one

to predict qualitatively

what hap-

pens to the thermal conductivity of a metal when it becomes superconducting (see Mendelssohn, 1955; Klemens, 1956). The condensed
'superconducting' electrons cannot carry thermal energy nor can

With decreasing temperature their number


and that of the 'normal' electrons correspondingly

they scatter phonons.


increases,
7

87

88

Superconductivity

decreases, which will result in a rapid decrease of the electronic heat


conduction. At the same time the conduction by phonons will be

enhanced, as these are no longer scattered as much by electrons.


In pure specimens, the decrease in k will usually exceed any gain
es

in

kgs and

the total conductivity in the superconducting phase will


then be much smaller than in the normal phase. This is illustrated,
,

for example, by the results of

Hulm

(1950)

on pure

Hg shown

in

Thermal conductivity

89

For moderately impure specimens the superconducting conductivity will then not be very different from the
corresponding normal one. This is shown, for example, by the results
of Hulm (1950) on a Hg-In alloy, also displayed in Figure 25. The
results of Lindenfeld (1 96 ) on lead alloys shown in Figure 26 indicates
what happens with increasing inpurity content: as the phonon contribution to the normal conductivity becomes more appreciable, the
the thermal circuit.

0.30

K (watt/cm-deg)
Pb + 6%Bi

Pb+3*h
-Pb+6%ln

0.20-

0.10-

Fig. 25

Figure 25. There

exist, however, pure materials in which the normal


not very high but which are very free of grain boundaries and other lattice defects. In the superconducting phase of such
substances at very low reduced temperatures the phonons are then

conductivity

is

hardly scattered by anything except the specimen boundaries, resulting in a large value of kgs This has been observed, for example,
by

gain in

k^ increasingly outweighs

tivity in the

in the

the decrease in

superconducting phase becomes

k es and the conduc-

much

larger than that

normal one.

9.2. Electronic

conduction

Calverley et

al. (1961) in tantalum and niobium.


Suppressing the electronic conduction in the normal phase by
adding impurities decreases the effect of condensing electrons out of

on thermal conductivity by the superconducting transition


indeed due to the disappearance of electrons from the conduction

If the effect
is

process, then

one should be able to write IX.2 for a superconductor as


!/*

x(ir)aT 2 +y(^)p Q ILT,

(IX.4)

90

Superconductivity

Thermal conductivity

where x(if) and y(ir) are functions only of the order parameter or
which indicates the fraction of condensed electrons. Equation 11.25
shows that IT is a function only of / = T/Tc so that one can write
,

instead

written in this

form

(IX.4)

to agree with the

nomen-

clature introduced by Hulm (1950).

He pointed out that if one chooses


a sample in which the electronic heat conduction is predominantly
limited by one or the other of the two scattering mechanisms, the
measured

ratio

k e Jk en then equals the appropriate ratio function g(t)

or /(r). To a first approximation, at least, these functions should be


universal functions for all superconductors and be related to the

microscopic nature of the phenomenon.


For a specimen for which Tmax < Tc as
,

pure
tion

is

BCS

theory.

The gradual change from a phononsame

scattered to an impurity-scattered electronic conduction in the

material of increasing impurity

is

recent results of Guenault (1960)

Uk = aT 2 /g(t) + PoLT/f(t).
The equation has been

the basis of the

the case for reasonably

Hg and Pb, and for extremely pure Sn and In, the heat conducjust below Tc is by electrons limited by phonon scattering. For

91

particularly well illustrated

on a

series

by the

of monocrystalline tin

specimens.

When thermal conductivity measurements on superconductors are


extended to small values of t, as was first done by Heer and Daunt
(1949) and later by Goodman (1953), /(/) is found to decrease very
rapidly.

Goodman

pointed out that this could be represented by an

equation of the form

At)

= aexp(-*/0,

(IX.7)

and suggested that this implied the existence of an energy gap between
the ground state and the lowest excited state available to the assembly
of superconducting electrons.
This conclusion can be inferred from thermal conductivity results
in the following manner. Simple transport theory shows that

such samples

KJKn *
All pertinent measurements

show

the

same

qualitative features: g(t)

breaks away sharply from unity with a discontinuous slope,


and decreases as a power of / which is about 2 for Sn and In (Jones
and Toxen, 1960: Guenault, 1960), and 4 to 5 for Pb and Hg (Watson
at

and Graham, 1963; see also Klemens, 1956). Calculations by Kadanoff and Martin (1961), by Kresin (1959) and by Tewordt (1962,
1963a) appear to explain the experimental results for Sn and In,
but not for Hg and Pb.
For specimens for which Tmax ^ Tc , the electronic conduction in
the superconducting phase

is

at all temperatures limited

by impurity

*/*,

Ce

(IX.8)

/ is

the

mean

free path,

specific heat of the electrons.

free paths (which may differ in magnitude) vary only slowly


with temperature, then the temperature variation of k e Jk e must be
due entirely to that of the specific heats. In other words

mean

f(t)

Cen

is

known

* kjk M

CJC

(IX.9)

to vary linearly with temperature, so that IX.7 implies

a'Tc texpi-b/t).

(IX.10)

(IX.6)

shown

/(0 approaches unity smoothly with a continuous


decreases

(1/3) lv

v tne average velocity, and Cc the


Assuming that v the Fermi velocity in
the normal metal, remains the same for the uncondensed 'normal'
electrons in the superconducting phase, and that in both phases the

where

Ces =

* /CO.

Several investigations (see Klemens, 1956) have

it

that

scattering, so that for these

lower temperatures

ke

(IX.5)

sit).

that at

slope,

more slowly than

git).

and
The

/=

that at
results

are in reasonable agreement with expressions for /(/) derived


Bardeen etal. (BRT, 1959) and by Geilikman and Kresin (1959)

by
on

That such a temperature variation of the specific heat corresponds to


an energy gap in the electronic spectrum can be shown as follows: If
a gap of width 2e lies below the lowest available excited state, the

number of thermally

excited electrons will be proportional

expi-2e/2k B T), where k B

is

the Boltzmann constant,

to

and the factor

2 arises because every excitation creates two independent particles,

92

Superconductivity

an electron and a

hole.

Thus the

Thermal conductivity

energy of the superconducting


phase is equal to the condensation energy per particle
multiplied by
the exponential factor, which remains unchanged,
through two differentiations with respect to temperature, to
appear in the
specific

heat.

The parameter b
1.0

in IX. 10

is

93

free

thus seen to equal 2e/2k T


B c

.6

Figure 27, together with a theoretical curve calculated with a gap

equal to 3-50 k B Tc

The agreement is somewhat deceptive, since there


good evidence that the gap width for aluminium is only 3 40 k B Tc
From an observed anisotropy in the temperature dependence of k
at very low temperatures Zavaritskii (1959, 1960a, b) has been able
to infer a corresponding anisotropy in the width of the energy gap in
the spectrum of the superconducting electrons in the case of cadmium,
tin, gallium, and zinc. To the last he could apply theoretical expressions due to Khalatnikov (1959), from which he deduced a gap anisotropy of about 30 per cent. A similar result holds for cadmium. The
measurements of Zavaritskii also show that the gap anisotropy can
have different forms in the case of gallium the value of the gap can
be approximated by an ellipsoid compressed along the axis of rotation; for zinc and cadmium this ellipsoid is stretched out along the
.

is

Aluminium

K en
.4-

(.Zavaritskii)

axis of rotation.
{Aluminium(Satterthwaite)

ITheory (BRT)

.4

.6

.8

1.0

9.3. Lattice conduction

Fig. 27

According to the microscopic theory to be discussed in Chapter


XI,
the energy gap is a function of temperature.
The parameter b can
therefore be written as

<T)
kB

where

(0)

Tc

*(0)

kB Tc

"
x

e(D
(0)

the gap value at 0K.

The detailed dependence of


on b has been calculated by BRT, and the function
e(DMO),
calculated from the BCS theory, has been tabulated
by Miihlschlegei
(1959). Measurements oikJk can thus be used
to infer the value of
e
<0)lkB Tc
is

In cases where the energy gap is a function of the magnetic field,


measurements of k c Jk cn can be used to infer this field dependence.
This technique has been used by Morris and Tinkham (1961) for thin
films (see Section 7.3), and by Dubeck et al. (1962, 1964) for type II
superconductors in the mixed state (see Section 6.6).

kjk en

The appropriate temperature dependence of kjk


en has been
observed in a number of metals. The results
for aluminium by
Satterthwaite (1960) and by Zavaritskii
(1958a) are shown in

Far below Tc the fraction of 'normal' electrons becomes so small as


to make k cs <^ k gs At the very lowest temperatures, the phonons are
primarily scattered by crystal boundaries in a manner which is the
same in the superconducting as in the normal phase. The charac3
teristic T dependence in this limit (Casimir, 1938) has been well
established experimentally (Mendelssohn and Renton, 955 Graham,
.

1958).

In the normal state there occurs at these temperatures

still

appre-

ciable heat conduction by electrons, limited only by impurity scattering

Thus

and varying

linearly with temperature (see equation IX.2).

in this range

kjk s = aT-\
where a

is

(IX.11)

a constant of the material which can have values as high

as several hundred.

For example, a

suitable lead wire can have

94

k n /k s

Superconductivity

10 at 0-1 K.

A number of authors (see Mendelssohn,

1955)
suggested using such wires in ultra-low temperature experiments
as
thermal switches which would be 'open', i.e. non-conducting, in
the

CHAPTER X

superconducting phase, and 'closed' when the superconductivity


is
quenched by means of a suitable magnetic field. Such heat switches
are now widely used (see, for instance, Reese and Steyert, 1962).

The Energy Gap

At somewhat higher temperatures,

at which the phonons begin to


be scattered by the 'normal' electrons even in the superconducting
phase, there is necessarily a concurrent rise of the electronic
conduction. Experimentally it is very difficult to separate the
conduction
mechanisms. Where this has been possible (Conolly and Mendels-

sohn, 1962; Lindenfeld and Rohrer, 1963) the results have


been
consistent with the pertinent calculations by BRT and
by Geilikman
and Kresin (1958, 1959).

10.1. Introduction

Ever since the initial discovery of superconductivity it had been known


but barely noted that the striking electromagnetic behaviour of a
superconductor at low frequencies is not accompanied by any corresponding changes in its optical properties: there is no visible change
at

Tc

to

its

although the

reflectivity of a

metal at any frequency

conductivity at that frequency.

Thus

frequencies the resistance of a superconductor

The thermal conductivity in the intermediate state


A number of experiments, in particular those of Mendelssohn and co-

9.4.

workers (Mendelssohn and Pontius, 1937; Mendelssohn and Olsen,


1950; Mendelssohn and Shiftman, 1 959), have shown that the thermal
conductivity of a superconductor in the intermediate state generally
does not change linearly from its value in the one phase to its
value

when at a given temperature the external field is varied.


Instead there appears an extra thermal resistance, which in some
cases
in the other

can be very large, and which is attributed to the scattering of


the
predominant heat carriers (electrons or phonons) at the boundaries
between the superconducting and normal laminae which make up the
intermediate state. For materials in which
nates this has been analyzed by Cornish

phonon conduction domiand Olsen (1953) and by


and Wyder (1963) have devel-

Laredo and Pippard (1955). Strassler


oped a treatment for very pure specimens

in which the conduction is


mostly by electrons. Experiments on the thermal conductivity in the
intermediate state thus yield strong confirmation that the laminar
structure, observed by various techniques at the surface of
a specimen,

actually persists throughout a bulk sample.

is

related

at the very high optical


is

a constant, inde-

pendent of temperature, and equal to that of the normal metal. At


about the time of the discovery of the isotope effect steadily improving high frequency techniques had shown that atOK the normal
resistance persisted

down to frequencies of the order of 10 13 c/sec, but


10

remained zero up to frequencies of the order 10 c/sec. In 1952


already Shoenberg ([1], p. 202) concluded from this that at some frequency between these two limits '.. quantum processes set in which
could raise electrons from the condensed to the uncondensed state

that

it

and thus cause energy absorption'.


As shown in the previous chapter, Goodman (1953) very shortly
after this inferred from his thermal conductivity results the existence
of an energy gap in the single electron energy spectrum. A similar
conclusion had been deduced a few years earlier by Daunt and
Mendelssohn (1946) from the absence of any Thomson heat in the
superconducting state. This indicated to them that the superconducting electrons remain effectively at 0K up to T= Tc by being in
low-lying energy states separated from all excited states by an energy
gap of the order of k B Tc
In the years which followed, the existence of such a gap was firmly
established by a large number of experiments, and this, together with
the electron-phonon interaction indicated by the isotope effect, pro.

vided the keystones of a microscopic theory. This chapter will


95

96

The energy gap

Superconductivity

describe a few experiments which indicate the


energy gap most clearly
directly. The subject has been reviewed
by Biondi et al. (1958)
recently by Douglass and Falicov
(1964).

and
and

electron excitations

97

would be expected to occur across the narrower


would be reflected in an upward curva-

portions of the gap, and this


ture of

Ces when

plotted semi-logarithmically against

compares a number of measurements which show


10.2. The specific heat
After the resurgence of interest in specific
heat measurements as a
result of the suggestive results of precise
threshold field measurements,
of Goodman's thermal conductivity results,
and of the first clear
experimental verification of a deviation from a 3
T law by Brown et al
(1953) on niobium, there have been in recent
years a number of

measurements which clearly indicate the exponential


variation of
corresponding to an energy gap. The first of these
were the

this

BCS

1/r.

Figure 28

curvature with

3CWT)3

results of

Corak

et al. (1954) on vanadium and by


Corak and Satterthwaite
on tin and since then the exponential variation of C
es has been
established in a number of elements. The
appropriate column in
(1 954)

Table III lists the energy gap values of these elements


deduced from
the specific heat measurements. Note that in
units of k B Tc these gaps
are of very similar size for widely varying
superconductors. This again
bears out the basic similarity of all
superconductors in terms of reduced co-ordinates.
It is

perhaps useful to consider

briefly the difficulty

of obtaining

good values for Ces What is measured, of course, in


both the superconducting and in the normal phase, is the total
specific heat. It
.

is then
necessary to separate the electronic from the
lattice contribution in
the normal phase in order to be able to
subtract the latter from the
total specific heat in the superconducting
phase. Unfortunately,

at

low temperatures,

with large Debye

Cga

even

small compared to C only for metals


temperatures. These are just the hard, high-melting
is

point metals which are difficult to obtain with


high purity, without
which superconducting measurements are misleading.
The softer and
lower melting point metals, on the other hand,
have a very unfavourable ratio of electronic to lattice specific heat.
Measurements by Goodman (1957, 1958), Zavaritskii
(1958b), and
Phillips (1959) on aluminium have shown
at very low temperatures
(/ < 0-2) a deviation of C
es from a simple exponential law (Boorse,
1959). Cooper (1959) has pointed out that this
can be a consequence
of anisotropy in the energy gap. At the
lowest temperatures most

Fig. 28

the exponential law expected

theory of Anderson

(1

from the

BCS

theory. According to a

959) (see Section 12.2) the gap anisotropy of an

element diminishes with the addition of impurities. Indeed Geiser and


Goodman (1963) have found in aluminium specimens of different
purity that the deviation of

with increasing impurity.

Ccs from

an exponential form decreases

98

Superconductivity

The energy gap

10.3. Electromagnetic absorption in the far infrared

The magnitude of the energy gap 2e(0) can be characterized by

a fre-

i/
? such that hv ? = 2e(0). It is at this frequency that one would
expect the change from the characteristically superconducting re-

quency

sponse to low frequency radiation, to the normal resistance main-

99

The measurements of the transmission of such radiation through


superconducting films will be discussed in a later section. Richards
and Tinkham (1960), Richards (1 961), and Ginsberg and Leslie (1962)
have observed directly the absorption edge at the gap frequency in
bulk superconductors. Radiation from a quartz mercury arc infrared

Table

III

Energy gap (2<Q)lk B Tc)

Element

Aluminium

316

2-9

3-37 b , 3-43 c
3-3 d

3-5

...

...

3-3

...

...

3-5

3-9

3-63 a , 3-45 b

3-5

Cadmium
Gallium
Indium

4-1

Lanthanum

2-85

Lead

414

...

3-9

3-7

...

4-0

4-33 a , 4-26 b

...

3-9

418 d
4-6
2-8

Mercury
Niobium

...

...

...

4-4

3-7

...

3-84 e ,3-6*,

3-7

3-59 8

Rhenium
Ruthenium
Tantalum

3-3

...

31

< 30

3-6

3-6

3-60, 3-5',

3-65 u
2-8

3-2

Thallium

Thorium

3-5

...

3-6

Tin

3-3

3-5

3-6

3-46 a , 3-47 b

3-6

3-65 d

J-VL
10

15

Vanadium
20 25 30 35 40 45 50

FREQUENCY (cm -1 )
Fig. 29

tained at high frequencies. Unfortunately, the frequencies corresponding to gap widths inferred from the specific heat measurements
are 10 u -10 12 c-sec~ \ which is an experimentally awkward range
at
the upper limit of klystron-excited frequencies, yet very
low for
mercury arc ones. Only recently have Tinkham and collaborators
developed the techniques needed to detect the very low radiation
intensities available in this far infrared region.

Zinc

3-4

...

...

...

...

3-6

3-4'

3-6

2-5

...

3-4

A from infrared absorption (lead: Ginsberg and Leslie,

1962;

lanthanum: Leslie et al., 1964; all others Richards and Tinkham, 1960).
B from infrared transmission (Ginsberg and Tinkham, 1960).
C from microwave absorption (aluminium: Biondi and
Garfunkel, 1959; tin: Biondi et ai., 1957).

heat data to exponential (Goodman, 959).


Dby
E from tunneling ("Giaevcr and Megerle, 1961; Zavaritskii,
fitting specific

1961;

d
Douglass and Merservey, 1964;
Sutton. 1962; 'Giaever, 1962; "Sherrill and

Douglass, 1962;

Townsend and

Edwards, 1962; "Dietrich, 1962).

Fcalculated from XI.32 (Goodman,

1959).

100

The energy gap

Superconductivity

monochromator was fed by means of a light pipe into a cavity made


of the superconducting material under investigation. The cavity contained a carbon resistance bolometer, and was shaped so that the
incident radiation

would make many

reflections before striking this

For frequencies lower than v the superconducting walls of


g
do not absorb, and much radiation reaches the bolometer.
absorption by the walls sets in, and the signal from the bolo-

detector.

the cavity

At

g,

meter decreases sharply. Figure 29 shows normalized curves of the


fractional change in the power absorbed by the bolometer,
in arbitrary units, plotted against frequency for all the metals investigated

by Tinkham and Richards. The gap values obtained are listed in


Table III. The absorption edges for Pb and Hg show a certain struc-

which has also been found in the same elements in infrared


transmission measurements (Ginsberg et al., 1959). Ginsberg and
ture,

Leslie (1962) have

shown

that this structure persists even in a lead


alloy containing 10 atomic per cent of thallium, so that it is
probably

not due to gap anisotropy. The effect may be due to states of collective
excitations lying in the gap (Tsuneto, 1960)

into account in the

BCS

theory.

which have not been taken

However, calculations of Maki and

Tsuneto (1962) lead one to expect that the energy of collective excitations should be drastically shifted by impurity scattering.
Richards (1961) has reported measurements on single crystals of
pure tin and of tin containing 01 atomic per cent indium. His results

show

that the position of the absorption edge varies with


crystal
orientation, which clearly indicates the anisotropy of the gap.

Furtheranisotropy decreases with increasing impurity, which


strongly supports Anderson's suggestion (1959) that the anisotropy

more

this

becomes smoothed out in impure samples. The absorption edges


observed by Richards have a structure which, unlike that seen in
Pb
and in Hg, occurs for frequencies greater than v These postcursor
g
peaks do not seem to change with impurity, and have not yet found

101

from a simple two-fluid picture, according to which at any finite


temperature a fraction of the electrons remains ' normal'. H. London
pointed out that in the presence of an alternating electric field these

this

electrons absorb energy as they

would

in a

normal metal, and that

needed to sustain an alternating current even in a superconductor because of the inertia of the superconducting electrons.
Into a normal metal an alternating field penetrates to a skin depth
such a

field is

which leads to anomalous results if the mean free path l> 8, as is


the case at high frequencies and low temperatures (see p. 42). In the
8,

superconducting phase, the theory of the anomalous skin effect still


applies in principle, but has to be modified both because for high frequencies the superconducting penetration depth A is much smaller
than the skin depth 8 (except very near Tc ) and decreases very rapidly
with decreasing temperature, and because the number of 'normal'
Tc Both of these lead to a reduc-

electrons also drops sharply below

tion of the resistance in the superconducting phase as

that in the normal one

below

Tc

compared

to

the ratio of the resistances decreases rapidly

changes more gradually at lower temperatures where both

A and the order parameter are

0K where

there are

and

fairly constant,

no more 'normal'

finally vanishes at

electrons.

Unpublished calculations of the variation of RJR with temperaand with frequency have been made by Serber and by Holstein
on the basis of the Reuter-Sondheimer equations, the London theory,
ture

and the two-fluid model. Typical results are the solid curve labelled
rc and the dashed one labelled 2-37 k B Tc in Figure 30. With
10
c/sec there is general experimental agreefrequencies up to 8 x 10
calculations,
as shown, for example, by the recent
ment with these
results of Khaikin (1958) on cadmium and of Kaplan et al. (1959) on
tin. Their temperature dependence for a given frequency can be
represented by an empirical function, suggested by Pippard (1948):

0-65Ar B

an explanation.

#0 =

even at lower frequencies (H. London, 1940). One can understand

(l-/2 )0-/4
4/3

2
.

(X.1)

is as v
at low frequencies, tending toward a constant value at higher frequencies.
However, surface impedance measurements at frequencies con10
siderably higher than 8 x 10 c/sec show appreciable deviations from
the predictions of the simple two-fluid model. Figure 30 shows the

The frequency dependence


10.4. Microwave absorption
Although the resistivity of a superconductor vanishes at 0K for frequencies up to v there is a finite resistance at higher
temperatures
g

4
/

102
ratio

Superconductivity

RJR n

measured

The energy gap

as a function of reduced temperature


for
at three microwave frequencies
by

aluminium as
Biondie/A/ (1957) The
of kB TJh. For 0-65, the results
agree

frequences are given in units


well with the temperature variation
calculated without regard to an
energy gap. For 2-37, however, such

calorimetrically the

103

amount of energy absorbed by an aluminium

wave guide, over a range of frequencies ranging from 0-65k B Tc


10
10
(1-5 x 10
c/sec) to 3-91^7^ (10 x 10
c/sec) at temperatures down

calculations would give the


evident that for / > 0-7, the measured
ratio
considerably exceeds the predicted
one. The same is true for
hv - 304k B T except that in this
c
case the deviation already begins
at

dashed curve, and

it is

O.^r^k
0.6

07

'

'

'

0.8

'

0.9

1.0

Fio. 30

Clearly an additional absorption


quencies,

and of course

mechanism occurs

sufficient to bridge the

c,

are not

gap at / = 0, but become effective at that temperature at which the gap has shrunk
to a width of 2-37
B T A series
of measurements of the resistance ratio
as a function
both of frequency

and of temperature thus serves to map out


the temperature variation
aP f 3ny g,Ven ^Perconductor. Biondi and

no?
(1959)

1.0

for these fre-

due to the boosting of condensed electrons across the energy gap. If this
gap had a constant width at all
t< 1 the appearance of this extra absorption
would depend only on
the frequency. Its temperature
dependence, however, clearly shows
that the energy gap varies with
temperature, tending toward zero as
/-M. As a result, photons of energy 2-37k T
for example,

Garfunkel
u
have
obtained values of the resistance
ratio by measuring

1.5

Energy

this is

2.0
(in

2.5

units of

3.0

3.5

4.0

kTc )

Fig. 31

The accuracy of the measurements was such that the absorption of 10 " 9 watt could be detected. Their results give a temperato 0-35K.

ture variation of the


tions of the

gap which

is

in close

agreement with the predic-

BCS theory.

Mattis and Bardeen (1958) and Abrikosov et ul. (1958) have developed a theory of the anomalous skin effect in superconductors on
the basis of the BCS theory. Miller (1960) used the work of the former
to calculate the surface impedance for many different frequencies

temperatures.
8

The

close agreement between his results

and
and the

104

105

Superconductivity

The energy gap

measurements of Biondi and Garfunkel is shown in Figure 31, in


which points calculated by Miller are superimposed on smooth

of the energy gap. In particular, a detailed analysis by Hebel (1959)


has shown that the empirical results are compatible with the manner
of piling up predicted by the microscopic theory. Hebel's results and
some empirical values are given in Figure 32, in which the ratio of the

curves representing the empirical values. The theoretical treatments


are equally successful in the lower frequency range in which there are

no gap

relaxation rate in the superconducting phase to that in the normal one


variation of what can
is plotted against temperature. The temperature

effects.

Nuclear spin relaxation

10.5.

When

the nuclear spins of a substance are aligned by the application


of an external field, they again relax to their equilibrium distribution

be considered as the attenuation of the nuclear alignment is markedly


different from the corresponding change in the attenuation of an

predominantly by interaction with the conduction electrons. In this


interaction, a nucleus flips its spin one way as the electron spin flips

way so as to conserve the total spin. The electron can do this


is available an empty final state of correct energy and
spin direction, and the nuclear relaxation rate in a normal metal
depends therefore both on the number of conduction electrons (itself
proportional to the product of the density of states and the energy
derivative of the Fermi function) and on the density of states in the
the other

only

if

vicinity

To

there

of the Fermi surface.

predict the temperature variation of this relaxation process in

the superconducting phase one

is tempted to use again a simple twomodel, according to which the number of 'normal' electrons
available decreases rapidly below T To this should, therefore, correc
spond a decrease in the relaxation rate as compared to that in the

fluid

'0

normal phase. But the energy gap severely modifies the density of
gap there are, by
definition, no available states at all, and the missing states are 'piled
up on either side. The presence of the energy derivative of the Fermi

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

states available to the interacting electrons. In the

Fig. 32

'

function in the relaxation rate expression makes this rate essentially


proportional to the square of the density of states evaluated over a
range k B T on either side of the Fermi energy. At temperatures near

ultrasonic elastic

< k B T, so that the pile-up of


an appreciable increase of the rate over
that in the normal phase. At lower temperatures, the gap becomes
wider than k B T, and the relaxation rate rapidly diminishes, approach-

ductors

7;.,

the gap

states

is still

on either side

very narrow and


results in

ing zero as T-*-0.

The measurements of Hebel and Slichter (1959), of Redfield (1959)..


and of Masuda and Redfield (1960a) fully confirm this consequence

wave in a superconductor. As this difference is one


of the most striking consequences of the BCS theory, its discussion
and the general description of ultrasonic attenuation in superconis postponed until a later chapter.
In his calculations, Hebel avoids singularities on either side of the
gap by introducing a parameter r which represents a smearing of the
density of states over an energy interval small compared to the width
of the gap. It is possible to interpret this in terms of an anisotropy of

the gap, since the relaxation process samples the gap over all directions simultaneously. With this interpretation, the data on aluminium

106

of

Superconductivity

The energy gap

Masuda and

Redfield (1960a, 1962) indicate an anisotropy of the


order of 1/10 of the gap width, and recent measurement
by the same

authors (Masuda and Redfield, 1960b; Masuda,


1962b) indicate that
this anisotropy decreases in impure
aluminium. Anisotropy of magnitude similar to that in aluminium has been found
by Masuda
(1962a) in cadmium.
10.6. The tunnel effect
The most recent and the most direct measurement of the
energy gap
has been provided by the work of Giaever
(1960a), who essentially

metal,

all

107

at 0K. In the last of these, electrons

fill

all

available states

up to the Fermi level EF in the superconductor, there is a gap of halfwidth e(0), and states up to EF e(0) are filled. With such conditions
there can be no tunneling either way, as on neither side of the barrier
are there any available states.
A potential difference applied between the two metals will shift the
energy levels of one with respect to the other. It is evident from
Figure 33 that tunneling will abruptly become possible when the
;

applied voltage equals e(0).

The subsequent

variation of tunneling

on the details of the


on either side of the gap.
of current with voltage due to the

current with applied voltage of course depends


density of states curve of the superconductor

At

first,

there

is

a very rapid

rise

(0)

Voltage
Fig. 34

large density of piled-up stages; for voltages

much exceeding e(0), the

tunnelling samples the density of states well beyond the gap,

superconductor

normal metal

variation of I vs.

insulator

tion of
Fig. 33

and the

V approaches the purely ohmic character of a junc-

two normal metals. This

is

summarized

in

Figure 34, which

gives with the solid line the current- voltage characteristic of the

superconducting-normal junction at 0K. The dotted line indicates

measured the width of the gap with a voltmeter. He


accomplished this
by observing the tunneling of electrons between
a superconducting
film and a normal one across a thin insulating
barrier.

Quantum-

mechanically, an electron on one side of such a


barrier has a finite
probability of tunneling through it if there is
an allowed state of equal
or smaller energy available for it on the other
side. Figure 33 shows
the density of states function in energy
space for a sandwich consisting,

from left to right, of a superconductor, an insulator,


and a normal

<

T< T

the modification being due to the fact


c
on both sides of the junction some electrons
are excited across the gap or the Fermi level, respectively. The dashed
line shows the behaviour at T > Tc i.e. for a junction of normal metals.
Nicol et al. (1960) and Giaever (1960b) have extended such experiments to cases where both metals of the junction are superconductors,
but with very different critical temperatures, such as Al (Tc = 1-2K)
and Pb (Tc = 7-2K). The gaps of the two will be correspondingly

the behaviour at

that at finite temperatures

'08
different,
is

shown

The energy gap

Superconductivity

and

for such a junction the density of states function at

in Figure 35.

A tunneling current will

begin to flow

the potential difference between the


this case,

0K
when

two metals is (0) Pb +e(0) AI In


however, the modification due to finite temperature is more
.

than with an s-n junction. Imagining the density of states


curve of Figure 35 with a few excited electrons beyond both gaps, and
a few available states remaining below both, one recognizes that now
significant

superconductor

tion of the energy gap with magnetic field was discussed in Chapter
VI I. Recent tunneling studies have verified other aspects of the energy
gap, in particular its relationship to the phonon spectrum of the
lattice. This will be summarized in Chapter XI.
Simultaneous tunneling of two electrons has been observed by
Taylor and Burstein (1962), in agreement with the calculations of

superconducting

superconductor

109

compared to the penetration depths, and because of their size their


critical fields are very high. This and its use in investigating the varia-

6,-e 2 e,+e2

insulator

Voltage

Fig. 35
Fig. 36

the current / at

first increases with increasing potential V, then


decreases for e(0) Pb -e(0)
A1 < K<(0) Pb + (0) A1 and then increases
again. Figure 36 shows the current-voltage characteristics
in this
case; the limits of the negative resistance region are very sharp.
Thus
,

the current-potential characteristics yield the energy gap values at


a
given temperature for both metals.

Schrieffer and Wilkins (1962). This is not to be confused with the


tunneling of Cooper pairs, as predicted by Josephson (1962), which
will be discussed in Section 1 1.7. The results of Taylor and Burstein

also indicate the possibility of tunneling assisted by the simultaneous


absorption of a phonon. Theoretical aspects of this have been dis-

The energy gap values obtained by this method for several superconductors are listed in Table III, and can probably be considered
as

cussed by Kleinman (1963) and Fibich (1964).

the most reliable of all experimental determinations.


Measurements
as a function of temperature closely support the thermal variation
of the energy gap predicted by the BSC theory. The films used
are thin

10.7.

Far infrared transmission through thin films

Tinkham, Glover, and Ginsberg (Glover


and Tinkham, 1957; Ginsberg and Tinkham, 1960) have measured
In a series of experiments,

110

Superconductivity

The energy gap

the transmission through thin superconducting films


of electromagnetic radiation in the far-infrared range of
wavelengths between 0-1

and 6 mm. Their results lend themselves to an ingenious


analysis
leading to a number of very fundamental conclusions
about the inter-

independent of frequency

in the

1 1

range under investigation. The super-

conducting conductivity can be written as the complex quantity


(X.2)

relation of the energy gap, the response to high


frequency radiation,
It

then follows from general electromagnetic theory that

= T^Hi-rt^)

+[( I

-^> ,/2 ^]~}~

(X- 3 >

Microwave work on bulk superconductors, such as the measurements


of Biondi and Garfunkel (1959), have shown that at T< Tc and

ho < k B Tc

the surface resistance vanishes. It follows

from

this that

the real, lossy part of the conductivity must also vanish in this range,

0, so that the low frequency measurements of Ts /T can be


used to evaluate the corresponding values of a2 /a. For a number of

or a,

samples of tin and lead with widely varying normal conductivity, all
the data of Glover and Tinkham fit a universal curve represented by
il n

As a n

is

(\la)(kB TJtiw),

0-27.

(X.4)

independent of frequency, X.4 implies that

a2
This

is just

cc

(X.5)

1/cu.

the frequency dependence which follows from the

London

equation

curlJ + -_-2

Fig. 37

and the existence of perfect conductivity and of the Meissner


effect in
the limit of zero frequency (see Tinkham
[10], pp. 168-176).
In Figure 37, the curve labelled TJT is one which can
be drawn
n
through the empirical values of the ratio of the transmissivity
in the
superconducting phase, Ts , to the normal value, T,
all suitably
normalized for film resistance and substrate refraction, and
conductivity.

The

transmissivity of a substance

is

related to

curlE

real

number, an , which to a good approximation

= -H
c

leads to

a2

c2
4ttA

(X.7)

2
,

its

One can approximate the conductivity of the film in the

normal state by a

(X.6)

0,

since this with Maxwell's equation

plotted

against frequency.

H=

is

An

imaginary conductivity which

is

inversely proportional to the

frequency thus corresponds to the consequences of X.6 the Meissner


:

CHRIST'S
I

COLLEGE

mm nw

112

Superconductivity

The energy gap

and a finite penetration depth A. However, the magnitude of A


from the experimental transmission results with the aid
of X.7 exceeds the London value \ L = mc 2 jATT 2ne 2 by at least a factor

Substituting X.7 into the first of these two relations shows that the
imaginary conductivity o 2 must be accompanied by a real conduc-

effect

calculated

often. Furthermore, there is nothing in the

why a2 /cr

tivity

which takes the form of a delta-function

London theory to explain

equation like X.4.

On the other hand the


as

l< A,

A2

Pippard treatment predicts

that (see equation IV. 18a)

(c /8A ) 8(co

Similarly, in terms of the empirical value

at the origin:

- 0).

(X.9)

X.4 for o 2 /o one would

have

TT\

k B Tr

o~

2a

a, (to)

for different superconductors should satisfy a universal

for these films, in which

1 1

(lo/OAl,

where (equation IV.9)

S(o)-O).

(X.10)

Such an infinite real conductivity at zero frequency of course does not

afiv /k B

Tc

introduce losses.

Hence
ne2

CT

On

XI X X

to

For a normal metal

mvQ
6

and hence the Pippard theory

leads to

10
12
8
-h<y/k BTc

Fig. 38

kB T

On

htxi

Turning now to the high frequency far infra-red transmissivity data,


and subsequent decrease of Ts /T indicates that at a frequency roughly corresponding to the peak, a real, lossy component
ct, of the superconducting conductivity must appear. In the absence
of such a component Ts /T would continue to rise. The appearance of
the peak

for all superconductors. This

is

strikingly verified

by the

results of

Glover and Tinkham.

The real and imaginary parts of any linear response function, such
as the electrical conductivity, are related by a pair of integral transforms known as the Kramers-Kronig (K-K) relations. In terms of the
conductivity these take the form:

Tinkham et al. do not determine the gap quite unambiguously (see Forrester, 1958). However, accepting the existence of

+ 00
a 2 (co

i<)
t0

)dco\

(0 2

o 2 (to)

is,

of course, highly suggestive of an energy gap. Taken by themselves,


the data of

+ 00
co x

a real component of conductivity at or near some critical frequency

-.

co

to

(X.8)

a gap from other experiments allows a fully consistent interpretation


of the transmission results from which the magnitude of the gap as
well as other interesting quantities can be derived.

Superconductivity

14

The energy gap

The calculations of Miller (1960) of the variation of a


shown in Figure 37, the ordinate being scaled in units of

\a n are

it

A removed from under the o-,(o>) curve by


must reappear somewhere else, and it can do so only
the form of a delta function of strength A. This being

follows that any area

the energy gap


at the origin in

2e(0)

1 1

the case, one can then again apply the

K-K

relations to

show

that

associated with such a delta function


is the width of the gap at 0K. An energy gap implies
a normal metal, the imaginary part of the conductivity
vanishes for frequencies beyond the gap. Using X.3 one can then

where 2c(0)

a,(oi)

A8(oj-0)

(X.12)

that, as for

cr,/<x from the measured values of TJT to a first approxin


mation, and then apply an iterative procedure using the K-K relations

must be a contribution to the imaginary conductivity of magnitude

calculate

as well as the

sum

rule about to be

mentioned to obtain final values


of ajcr,,. Figure 38 gives the result thus obtained by Ginsberg and
Tinkham for lead, showing the precursor peak also found for mercury. One ignores this in deriving energy gap values from the limit
>
<y
i/CT/.- 0- The resulting gap widths are listed in Table III.
10.8. The Ferrell-Glover sum rule
The intimate connection between the experimentally verified decrease

of aj/a near

co

g , corresponding to the existence of a gap,

frequency London-type imaginary conductivity a \

and the low

<x 1/eo,

the real part of the conductivity vanishes.

The appropriate K-K relaa is an even


t

function,

a 2 (o>)

TTOi

9f(a>i)dh>|.

CX.11)

must lead to the appearance of a

ff2 ccl/o,

which was

seen to correspond to the Meissner effect and infinite conductivity.


One sees further that in terms of the parameter a of X.4, one can
write

A\on

(7Tl2)(k B

Tclh)(Ma).

(X.14)

Determining A/a from their transmission data and using this relation, Ginsberg and Tinkham obtain values for a of 0-23 for lead, 0-26
for tin,

and 0- 19 for thallium. These, as well as Glover and Tinkham's

value of 0-27 for both tin and lead, are in remarkable agreement both
with the Faber-Pippard data (0-15 for tin and indium) and with the

BCS

prediction for

all

metals (0-18).

The agreement

is

particularly

one considers the simplifications of the theory on the


one hand, and the wide variety and considerable difficulty of the
experiments on the other.
From X.13 and X.7 it is evident that
convincing

if

At these very high frequencies all electrons are free in both the normal
and the superconducting phases, and one would thus expect cr 2 (co)
and, therefore, the integral in X.l to have the same value in both
phases. In other words, there exists the sum rule that this integral

Thus the Ferrell-Glover sum-rule

remains unchanged under the superconducting transition.

by Gor'kov.

From this

delta function X.12. In turn this

leads to a London-type imaginary conductivity

A2

(X.13)

IA/tho.

The argument has now come through a full circle. An energy gap
corresponds to a disappearance of a^w) in the superconducting
phase over some frequency range in which this conductivity is finite
in the normal metal. This, according to the Ferrel-Glover sum rule,

corre-

sponding to infinite conductivity and the Meissner effect at zero


frequency, was first pointed out by Ferrell and Glover (1958) and
further elaborated by Tinkham and Ferrell (1959). The first of these
papers pointed out that at extremely high frequencies, such that hut
far exceeds any of the binding energies of an electron in the metal,
tion for the imaginary conductivity then becomes, since

aiiai)

c I8A.

(X.15)

leads to an inverse proportionality


between the square of the penetration depth and the energy gap. Such
a relation is implicit in the Pippard model and the Bardeen theory,
and appears explicitly in the Ginzburg-Landau treatment as extended

Microscopic theory of superconductivity

117

under suitable conditions. The absence of statistical fluctushows that the superconducting state is a highly correlated one
involving a very large number of electrons. Thus it is necessary to find
inherent in the basic properties common to all metals an interaction
transition

ations

CHAPTER

XI

Microscopic Theory of Superconductivity


11.1. Introduction

empirical description of superconductivity, perhaps the most striking


feature to be noticed

is

how much

quantitative information can be

given about superconductivity in general without speaking about the

of any one of the

indicates that the explanation for superconductivity should be in-

herent in a general, idealized model of a metal which ignores the complicated features characterizing any individual metallic element. It

should, therefore, be possible to find in the simple model of the ideal

metal the possibility of an interaction mechanism leading to the superstate,

in such a

way that the energy

and to derive from

this at least qualitatively the

properties of an ideal superconductor.

One would judge from

this that an explanation for superconducshould be fairly easy, until he realizes the extreme smallness of
the energy involved. A superconductor can be made normal by the
tivity

which at absolute zero is of the


order of a few hundred gauss. The energy difference between the
superconducting and the normal phase at absolute zero, which is
given by Hq/Stt, thus is of the order of 10~ 8 e.v. per atom. How
very small this is can best be judged by remembering that the Fermi
energy of the conduction electrons in a normal metal is of the order
of 10-20 e.v. The simple model of Bloch and Sommerfeld gives a
reasonably accurate description of the basic characteristics of a metal
although it completely ignores, among other things, the correlation
energy of the conduction electrons due to their Coulomb interaction.
This energy is of the order of 1 e.v.
As a further difficulty in arriving at a microscopic theory of superapplication of a magnetic field

conductivity one must add the extreme sharpness of the phase


116

conducting elements clearly indicated that in these the interaction in


question must be one between the electrons and the vibrating crystal
lattice,

and indeed Frohlich (1950) had suggested such a mechanism

independently of the simultaneous experimental results.

many superconducting elements.

The astonishing degree of similarity in the superconducting behaviour


of metals with widely varying crystallographic and atomic properties

conducting

number of electrons

of the system relative to the normal metal is lowered by a very small


amount. The discovery of the isotope effect in a number of super-

In reviewing the contents of the preceding chapters, which give an

specific properties

correlating a large

11.2.

The

electron -phonon interaction

Frohlich and, a little later, Bardeen (1 950) pointed out that an electron
moving through a crystal lattice has a self energy by being 'clothed'
with virtual phonons.

through the

means is that an electron moving


and the lattice in turn acts on
of the electrostatic forces between them. The

What

this

lattice distorts the lattice,

the electron by virtue

is quantized in terms of phonons,


and so one can think of the interaction between lattice and electron
as the constant emission and reabsorption of phonons by the latter.
These are called 'virtual' phonons because as a consequence of the

oscillatory distortion of the lattice

uncertainty principle their very short lifetime renders

it

unnecessary

Thus one can think of the electron


being accompanied or 'clothed', even

to conserve energy in the process.

moving through the

lattice as

by a cloud of virtual phonons. This contributes to the electron


an amount of self-energy which, as was pointed out by Frohlich and
by Bardeen, is proportional to the square of the average phonon
at 0K,

energy. In turn this

is

inversely proportional to the lattice mass, so

would have the


mass dependence indicated by the isotope effect. Unfortuhowever, the size turns out to be three to four orders of

that a condensation energy equal to this self-energy


correct
nately,

magnitude too large.


It was only seven years later that Bardeen, Cooper, and Schrieffer
(BCS, 1957) succeeded in showing that the basic interaction responsible for superconductivity appears to be that of a pair of electrons
by means of an interchange of virtual phonons. In the simple terms

118

Microscopic theory of superconductivity

Superconductivity

used above

this

means

that the lattice

is

distorted

by a moving

elec-

19

between a pair of electrons just above the Fermi surface, these elecform a bound state. The electrons for which this can occur
as a result of the phonon interaction lie in a thin shell of width ^ hu) q
where hco q is of the order of the average phonon energy of the metal.
If one looks at the matrix elements for all possible interactions which

phonon. A second electron some


when it is reached by the propagating
fluctuation in the lattice charge distribution. In other words, as shown
in Figure 39, an electron of wave vector k emits a virtual phonon q

trons can

k'. This scatters k into k q and


The process being a virtual one, energy need not be
conserved, and in fact the nature of the resulting electron-electron
interaction depends on the relative magnitudes of the electronic
energy change and the phonon energy ficoq If this latter exceeds the

take a pair of electrons from any two k values in this shell to any two
others, he finds that because of the Fermi statistics of the electron
these matrix elements alternate in sign and, being all of roughly equal

tron, this distortion giving rise to a

distance away

which

is

is

in turn affected

absorbed by an electron

k' into k' + q.

magnitude, give a negligible total interaction energy, that is, a


vanishingly small total lowering of the energy relative to the normal
situation of unpaired electrons. One can, however, restrict oneself to
matrix elements of a single sign by associating all possible k values in
pairs, kj and k 2 , and requiring that either both or neither member of
a pair be occupied. As the lowest energy is obtained by having the
largest number of possible transitions, each represented by a matrix
all of the same sign, one wants to choose these pairs in such
a way that from any one set of values (kj, k^, transitions are possible
into all other pairs (k|", k. As momentum must be conserved, this

element

means that one must require that

k + k2
1

that

is,

that

all

bound

k,'

pairs should

+ k^ = K

(XI.1)

have the same total

momentum K.

(See, for example, Cooper, 1960.)

Fig. 39

To
the

former, the interaction


lattice

is

is

the charge fluctuation of the

attractive

then such as to surround one of the electrons by a positive

screening charge greater than the electronic one, so that the second

and is attracted by a net positive charge.


The fundamental postulate of the BCS theory is that superconductivity occurs when such an attractive interaction between two
electrons by means of phonon exchange dominates the usual repulsive
screened Coulomb interaction.
electron sees

11.3.

The Cooper

pairs

Shortly before the formulation of the

been able to show that

if

there

is

BCS theory, Cooper (1956) had

a net attraction, however weak,

find the possible value of kj

same time

lie in

a narrow

and k 2 which

1 and at
Fermi surface k F

satisfy XI.

shell straddling the

one can construct the d iagram shown in Figure 40, d rawing concentric
of radii k F - 8 and k F + 8 from two points separated by K. It is
clear that all possible values of k, and k 2 satisfying XI. 1 are restricted
to the two shaded regions. This shows that the volume of phase space

circles

what has become known as Cooper pairs has a very


for K = 0. Thus the largest number of possible
yielding the most appreciable lowering of energy is

available for

sharp

maximum

transitions

obtained by pairing all possible states such that their total momentum
vanishes. It is also possible to show that exchange terms tend to reduce
the interaction energy for pairs of parallel spin, so that it is energetically most favourable to restrict the pairs to those of opposite
spin.

One can, therefore, summarize the

basic hypothesis of the

BCS

120

Superconductivity

theory as follows: At

0K

correlated one in which in


in

thin shell near the

Microscopic theory of superconductivity

the superconducting ground state

momentum space

Fermi surface are

the

is

a highly

normal electron states

to the fullest extent possible

occupied by pairs of opposite spin and momentum. The most direct


verification of the existence of these pairs arises from the flux quantization measurements mentioned in Chapter III.

The energy of this state is lower than that of the normal metal by a
finite amount which is the condensation energy of the superconducting

0K must equal Hi/Sir per unit volume. Furtherhas the all-important property that it takes a finite
quantity of energy to excite even a single normal ', unpaired electron.

state

and which

more,

at

this state

'

For not only does this require the very small amount of energy needed

121

mentioned earlier, namely that the correlation energy in question is so


very much smaller than almost any other contribution to the total
electronic energy. BCS therefore take the bold step of assuming that
all

interactions except the crucial

one are the same

for the supercon-

ducting as for the normal ground state at 0K. Taking as the zero of
energy the normal ground state energy and including in this all

normal state correlations and even the self energy of the electrons due
to virtual phonon emission and reabsorption, BCS proceed to calculate the superconducting ground state energy as being due uniquely
to the correlation between Cooper pairs of electrons of opposite spin
and momentum by phonon and screened Coulomb interaction.
The interaction leading to the transition of a pair of electrons from
the state (k t -k | ) to (k' t -k' I ) is characterized by a matrix
,

element,

-]^

fc/

= 2(-k'i,k'tl#int|-k!M),

(XI.2)

all terms comremoved.


have
been
superconducting
phases
and
mon
Vkk is the difference between one term describing the interaction
between the two electrons by means of a phonon, and a second one
giving their screened Coulomb interaction. The basic similarity of the

where i/ int

is

the truncated Hamiltonian from which

to the normal

Fig. 40

up a bound pair, but more importantly the occupation of a


k state by an unpaired electron removes from the system a large
number of pairs which could have interacted so as to occupy k and
k. Hence the total energy difference between having all paired
electrons and having a single excited electron is finite and equal to a
to break

single

large multiple of the single pair correlation energy. In terms of the

theBCS theory correctly yields an


has already been shown that such an energy gap not
only leads to the observed variation of the specific heat, the thermal
single electron spectrum, therefore,

energy gap.

It

conductivity,

teristic

BCS

of individual substances.

Vkk

therefore

make

the further

and constant for all


electrons in a narrow shell, straddling the Fermi surface, of thickness
(in units of energy) less than the average energy of the lattice, and
that Vkk vanishes elsewhere. Measuring electron energy from the
Fermi surface, and calling e k the energy of an electron in state k, one
simplifying assumption that

is

isotropic

can state

this formally

by the equations:

and the absorption of high frequency electromagnetic

radiation, but also that

it is

Vkk .= V forhfcl.M &

correlated with the existence of perfect

diamagnetism and perfect conductivity in the low frequency


11.4. The ground state energy
The recognition of the basic electron
sible

superconducting characteristics of widely different metals implies that


the responsible interaction cannot crucially depend on details charac-

(XI.4)
limit.

mechanism responfor superconductivity does not remove the major difficulty


interaction

Vkk =

and

The

basic

BCS

elsewhere.

criterion for superconductivity

condition

V<

0.

is

equivalent to the

Microscopic theory of superconductivity

Superconductivity

122
It is well to

note clearly at this point that this simplification of the

interaction parameter

F necessarily leads to what can be called a law

of corresponding states for


identical predictions for the

all

superconductors, that

exist,

and

to see whether this

respect to h k

virtually

is,

in terms of reduced co-ordinates. Any empirical deviation from such


complete similarity is, therefore, no invalidation of the basic premise
of the BCS theory, but merely an indication of the oversimplification

inherent in XI.4. (See footnote, page

30.)

Let h k be the probability that states k and k are occupied by a


pair of electrons, and (\h k ) the corresponding probability that the

possible XI.5' can be minimized with

This leads to
[h k (l-h k)}

magnitudes of all characteristic quantities

is

123

'

= vw

(XI.6)

2e k

l-2h.

By defining

KStMl-Ml

e(0)=

1/2

(XI.7)

kf

equation XI.6 simplifies to

h,==*-!)

(XI.8)

states are empty. W(Q), the ground state energy of the superconducting

state at

0K

as

compared

to the energy of the

normal metal,

is

then

where

Ek = [4+e 2 (0)]

given by

1/2

(XI.9)

relation
Substituting XI.8 back into XI.7 one obtains a non-linear

^(0)=

S^ArSW^l-MMl-^}"
k

2
.

(XI.5)

for(0):

Ky

kk"

e{0)

g(0)

"2Z[ I + eW

The summation is over all those k-values

for

which

Vkk

9* 0

k hk

- v2

{h k (i

kK

-MMi -h)) m

(Xi.50

to an
This can be treated most readily by changing the summation
k
to e.
integration
from
integration and transforming the variable of
surface
Fermi
the
of
either
side
states
on
Assuming symmetry of
=
introducing the density of single electron states of one
(c

0),

and

spin in the normal state at e

The first term gives the difference of kinetic energy between the superconducting and normal phases at 0K. The factor 2 arises because for
every electron in state k of energy e k there is with an isotropic Fermi
surface another electron of the same energy in k. This first term can
be either positive or negative, and is smaller than the second term

The limit of integration is

which

to XI.4,

from a
For such a transition to be

gives the correlation energy for all possible transitions

pair state (k,

k)

to another (k',

k').

(XI. 10)

so that

using XI.4 one can simplify to

wm - Sk 2e

'

k must initially be occupied and k' empty. The simultaneous


probability of this is given by h k {\ h k >). The final state must have k
empty and k' occupied, and this has probability h k -{\ h k). The
square root of the product of these probabilities multiplied by the
matrix element for the transition and summed over all possible values
of k and k' gives the total correlation energy.
W(0) must of course be negative for the superconducting phase to

= 0: M0), XI. 10 becomes


JlCOq

M0)

The

[e

the

+ 2 (0)] ,/2

phonon energy above which, according

V=0.

solution of XI. 11

possible,

e(0)

is

= /mysinMl/MO) V\.

(XI.12)

one finds
Putting this back into XI.9 and XI.7 and finally into XI.5',
given by
is
state
superconducting
of
the
that the ground state energy
W(0)

= -

2MQ)(W 2
exp[2/M0)F]-l

(XI. 13)

124

Superconductivity

from any theory which postulates an interaction between electrons


and phonons and allows this interaction to be cut off at some average

phonon energy hw q k B 6, beyond which the interaction becomes


repulsive. A term like this had been contained in the earlier
attempts
of Frohlich and of Bardeen, and, as mentioned before, is much too
large. The success of the BCS theory lies in the appearance
of the
exponential denominator which reduces W(0) by many orders of
magnitude. Although a precise calculation of the average interaction
parameter V for a specific metal continues to be among the most
important questions still to be solved, various estimates (Pines, 1958

Morel, 1959; Morel and Anderson, 1962) indicate that the values
Vx 0-3, derived from a knowledge of
are reasonable.

of N(0)

Q,

Thus the denominator has a value of about e 7


The isotope effect follows from the numerator of XI. 1 3, as it would
from any theory involving electron-phonon interaction with a cut.

off frequency related to the

6 and hence to the isotopic mass.

Debye

Equation XI. 13 shows that

H
~=
2

discussed by Ehashberg(1961),Bardeen[9],Schrieffer(1961),Betbeder-

Matibet and Nozieres (1961), and Bardasis and Schrieffer (1961).


quasi-particles is found to be very small even up

The damping of the


to energies well

the

beyond the Fermi energy. This

BCS assumption embodied

off of the

in

XI .4,

is in

contradiction to

as the justification of the cut-

Coulomb interaction at hu) q is essentially that quasi-particles


damped as not to be available for pair

of larger energy are so strongly

It is thus necessary to modify the BCS cut-off by taking into


account the existence of the repulsive interaction for e k > hwq This
does not appreciably affect the gap at the Fermi surface (e k = 0), but
will result in its variation with e k , as will be further discussed in

formation.

Section 11.7.

With a compound tunnelling arrangement in which electrons are


and then have the possibility of tunnelling through a second junction into normal metal,
Ginsberg (1962) was recently able to place an upper limit on the lifeinjected into a layer of superconducting lead

time of the quasi-particles in a superconductor. According to his pre7


liminary result this upper bound is 2-2 x 10~ sec, which is only about
times as large as the average time calculated by Schrieffer and
Ginsberg (1962) for quasi-particle recombination into pairs by means
of phonon emission. This has also been calculated by Rothwarf and
five

^(0)

cc {hu>
q)

For a group of isotopes, one

finds

(k D Q)

Tc

cc

Tc Mr* 12

cc

Mfol

(XI. 14)

Cohen

so that
(XI.15)

Any

appreciable deviation of the isotope effect exponent from the


value 0-5 could indicate that the simplifying BCS assumption of
a cut-

Coulomb and phonon interaction at hw has to be modiq


(Tolmachev, 1958; Swihart, 1959, 1962). Bardeen (1959) has
pointed out that the cut-off may be determined by the lifetime of the
'normal' electrons which can be excited across the gap. These elecoff for both
fied

trons are not the bare, non-interacting electrons of the simple BlochSommerfeld model. Instead they are so-called quasi-particles
'clothed' by their interactions with each other and with the lattice
\1], pp. 184-95). As a result the wave functions describing them
are not eigenfunctions of the system, so that the particles
have a finite
lifetime. The effect of this on the pair interaction
has been further

(see

125

Microscopic theory of superconductivity

The numerator of this quantity follows from dimensional reasoning

(1963).

Swihart (1962) as well as Morel and Anderson (1962) have studied


the isotope effect for different forms of the energy dependence of the

They find that the exponent of the isoXI.


5
less than the ideal value of one half by
equation
is
mass
in
topic
amounts of 10-30 per cent which increase with decreasing TJ6.
However, the isotope effects in ruthenium (Geballe et ol., 1961,
Finnemore and Mapother, 1962), osmium (Hein and Gibson, 1964)
and perhaps also in molybdenum (Matthias et a I., 1963) appear to be

electron-electron interaction.
1

too small to be explained by these calculations.


This raises questions about the origin of the attractive interaction
responsible for the formation of

Cooper

pairs in these as well as

perhaps in other metals. Matthias (see for example, 1960) has repeatedly suggested that in all transition metals there exists an attractive
magnetic interaction responsible for superconductivity. However,
both Kondo (1963) and Garland (1963a) have tried to explain the

126

Comparing XI.16 with XI.7 one

on

Rondo assumes a

the other hand, believes that the electrons of high effective

mass

As

1/N(0)

11.6.

added to the usual

As

attractive interaction

by exchange of

virtual

phonons.

Garland (1962b) calculated the magnitude of the isotope effect for


superconducting elements and obtains results which agree closely
available experimental results, including in particular the

all

reduced

2c(0)

d-band tend not to follow the motion of the s-electrons. This


results in 'anti-shielding' the interactions between ^-electrons, leading
to an attractive screened Coulomb interaction between them being

with

gap has the

2^^/sinh [1 /N(0)

(XI. 1 7)

V\.

larger interband interaction; Garland,

V& 3-4, this can be approximated by

in the

all

sees that this energy

value 2e(0), which according to XI. 12 equals

and d bands of the electronic spectrum, and not because of a magnetic


interaction.

127

Microscopic theory of superconductivity

Superconductivity

apparently anomalous superconducting behaviour of the transition


metals as a consequence of the overlap at the Fermi energy of the .v

effect in transition metals.

This also

results, at least quali-

from Rondo's calculations. Garland was also able to explain


the anomalous pressure effect in transition metals (Bucher and Olsen,

2 (0)

The superconductor

4Aco9 exp [-

(XI. 1 8)

/tf(0) V\.

at finite temperatures

the temperature of the superconductor is raised above 0R, an


increasing number of electrons find themselves thermally excited into

These excitations behave like those of a


normal metal; they are readily scattered and can gain or lose further
energy in arbitrarily small quantities. In what follows they are simply
single quasi-particle states.

called

normal electrons. At the same time there continues to exist the


all electrons still correlated into Cooper pairs, and

tatively,

configuration of

1964).

displaying superconducting properties, being very difficult to scatter


or to excite. One is thus led again to a two-fluid point of view.

As at 0R, one can write down an analytic expression for the ground
The energy gap at 0K
From XI.5' one can see that the contribution of a

state energy

11.5.

(k,

-k)

to this total condensation energy

single pair state

is

Letting

Wk =

2* k h k -2Vj:

W{J) containing a kinetic energy term and an interaction

term. In both, the presence of the normal electrons must be accounted


for, which is done by introducing a suitable probability factor fk

{0-MM

,/2
.

fk =

probability of occupation of

(XI.16)
1

-2fk =

first

term represents the kinetic energy of both electrons in the


and the second term the total interaction energy due to all

pair state k,

Wn] K E
.

the lowest excited state of the superconductor

must correspond to breaking up a single pair by transferring an electron from a


state k to another, leaving an unpaired electron in - k. The condensation energy is then reduced by
k The first term of this can be made
arbitrarily small, and is analogous to the excitation energy in a normal
metal, for which there is a quasi-continuous energy spectrum above
the ground state. The second term of
k however, is finite for all
values of k, which is why in the superconducting phase the lowest
excited state is separated from the ground state by an energy gap.

-k

k nor

occupied by a

2 M[fk + V-2fk)h k
k

(XI. 19)

electron.

This leads to a kinetic energy term

possible transitions into or out of the state.

At 0R

by a single

is

probability that neither

normal

The

-k

k or of

normal electron, then

],

where the summation is over the same range as at 0R, and h k retains
the same definition, though no longer the same value. The second
term
(k,

in the brackets clearly gives the probability that the pair state

k) not be occupied by normal electrons but by a correlated pair.

The

correlation energy at a finite temperature

is

kk'

x(l-2/*)(l-2/*0.

(XI.20)

128

Microscopic theory of superconductivity

Superconductivity

The last two terms ensure

that the correlated pair states not be occu-

pied by normal electrons.

It is

obvious that the presence of these

terms decreases the pairing energy.


The thermal properties of the superconductors can

where

Wm-TS =

the magnitude of the energy gap one must

expression for

fk

with respect to

fk

which one obtains by minimizing the

[W(T)\ K B MW(Tj\ C0 -TS,


.

(XI.21)

fk =

to the

Wv

The critical temperature Tc is reached when


up so that e(Tc) = 0. Hence

W
1

free

this

Ml-hM* = V S[Mi-M]"
l-2h

As long

(i-2A0

(XI.24)

and again obtains

One

now

(XI.28)

k B Tc

= M4/K^exp[-

1/W(0) V\.

(XI.29)

tion with pressure in aluminium.

11.7.
hi

is

The exponential dependence of the transition temperature has been


verified by Olsen et at, (1964) by means of measurements of its varia-

<T)= F[M1-/'a<)] ,/2 (1-2/,0,

Ek

pair states are broken

f de tan .

all

<XL27)

k B Tc < hw q the solution of this can be written as

(XI.23)

This time one defines

where

as

2e 4

(XI.26)

7^^(-^H'

(XI.22)

and XI.22 into XI.21, and minim ising


energy with respect to h k one now obtains

J [e2+

Substituting XI. 19, XI.20

C xp(Ek /k B T)+])-

Jiuia

state of highest possible order

TS = -2k B T-Z {A.lnA. + (l-/*)m(l-A)}.

an

free energy

XI.26, XI. 18, and XI.24 yield for e(T) sl non-linear relation which,
changing as before from a summation over k to an integration over
e, becomes

J is the temperature and S the entropy. This last is due entirely

normal electrons; the electrons which are still paired are in a


and do not contribute at all. Thus the
entropy is given by the usual expression for particles obeying FermiDirac statistics

129
find

first

This yields

now

be found
quite readily by writing down the free energy of the system and
requiring this to be at a minimum. The free energy is

G=

To evaluate

defined as

-Hi

Ek m [e k + e 2 (!T)] ,/2

(XI.25)

energy gap at

XL 18

and XI.29

yields for the width of the

0K
2e(0)

3-52 kB Tc .

(XI.30)

sees that, as at 0K, 2e(T) represents the contribution

of a
and that to break up
one such pair at any finite temperature removes from the superconducting energy at least this amount. In other words, the superconducting state continues to contain an energy gap 2e(T) separating the
lowest energy configuration at any given temperature from that with
one less correlated pair.
single pair state to the total correlation energy,

Experimental verification of predicted thermal properties

Combining equations

This

is in remarkable quantitative agreement with empirical values


obtained from the wide variety of measurements mentioned in Chap-

ter

X. Table HI shows that for the most widely different elements

the energy gap does not appear to deviate from this idealized value

by more than about 20 per cent. The theoretical temperature variation


of the gap width is displayed in Figure41
firmed by a

number of experiments.

this

has also been well con-

Superconductivity

130

Microscopic theory of superconductivity

Muhlschlegel (1959) has tabulated values of the energy gap, the


entropy, the critical magnetic field, the penetration depth, and the
specific heat, all in reduced coordinates, as functions of the reduced
temperature. All these are in close agreement with experimental
results.

These agreements clearly vindicate the basic BCS approach accord,

ing to which the similarities between superconductors outweigh their


differences, so that an

approximate law of corresponding states should

1.0

the dependence of the interaction parameter

Fon k and k',

131

so as to

be able to calculate directional effects. Even more challenging are the


previously mentioned attempts to relax, even in an isotropic model,
the assumption XT.4 that Fis strictly constant for e k < hojq and is then
cut off abruptly. A better knowledge of the variation of Fwith e k in

would allow the more precise calculation of the corresponding


dependence of the energy gap on e k The actual form of this variation
undoubtedly more nearly resembles the solid line in Figure 42 rather
than the dotted line which corresponds to the simple BCS assumption.
Usually one is interested in excitation energies of the order of k B Tc
and the BCS assumption is then fully applicable as long as k B Tc < hu) q
turn

OB

em

o.6

02
0.2

0.6

0.4

0.8

t.O

Fig. 41

is called the weak coupling limit. As hw q &k B , where


Debye temperature, this requires that

which
hold.f This similarity principle had of course emerged from

much pre-

vious experimental evidence. However, differences between metals


and anisotropics in a given metal do exist, and the experimental evi-

dence for gap variations from one metal to another, as well as for gap
anisotropics, clearly indicates the need to refine the details of the BCS
calculations. For one thing it is of course desirable to take into account
t Deviations from such a law can occur even with the BCS assumption
of constant Kif in solving equations XI.27 and XI.28 one takes into account
higher order terms in k B Te lhwq (Muhlschlegel, 1959). The resulting correction factors appearing in equations XI. 30, XI.35, and XI.36 are, however,
too small to explain the empirical deviations from similarity discussed in
this section. Thouless (I960) has shown that in the BCS formulation the
energy gap at 0K is only 4-0 k B Tc even in the non-physical limit

the

Tc <

is

e.

For a number of superconducting elements, in particular for Pb and


Hg, this condition does not hold.
Swihart (1962, 1963) as well as Morel and Anderson (1962) have
investigated the consequences of an energy dependence of the interaction Kmore realistic than that assumed by BCS. In particular they
take into account that, as was mentioned earlier, lifetime effects are
too small to justify cutting off the Coulomb repulsion at hm q Therefore these authors include in the interaction a repulsive part (V> 0)
The resulting variation of the energy gap at
for energies e > hw
.

0K

as a function of e k has been

shown by Morel and Anderson

have the form represented schematically

in

to

Figure 42. Swihart found

Microscopic theory of superconductivity

Superconductivity

132

that a rise of this gap function

on moving from

leads to the correct specific heat

jump

the Fermi surface

for lead at

Tc The
.

relation

between calorimetric and magnetic properties indicates that such a


gap variation is also consistent with the observed critical field curve

and probably also with that for mercury.


For even higher quasi particle energies the energy gap continues
to change sign periodically at multiples offiw This is consistent with
q
the observations of Rowell et al. (1962), who found maxima in the
tunneling conductance with that periodicity. Excitation of high
energy quasi particles involve multi-phonon interactions.
for lead

A precise calculation of the energy


content

itself

gap variation with

ek

in

which the energy gap

is

of the tunnel are superconducting. This current can be considered

/sla
.

n
SaTh V S+

3.6

kT c

calculation

V^g

Al|

26(0)

1 InRe

3.2

Ru
IS, -

fTl
Zr

(1964).

2.4

In considering an energy gap which changes sign as a function of


it

Pb

4.0 _

was carried
out for lead by Culler et al. (1962), using an on-line computer facility.
The relation between the phonon spectrum and the tunneling
characteristics has been fully discussed by Scalapino and Anderson

ek ,

et al.

The tunneling discussed thus far in this section and in Section 10.6
more quasi-particles. Josephson
(1962) has predicted an additional tunneling current when both sides

necessary to recognize the different frequency distributions for the

and transverse phonons. Such a

Rowell

involves the passage of one or

it

longitudinal

ments with lead, tin, and aluminium. Indications of this structure


had been seen earlier by Giaever et al. (1962) in lead and by Adler
and Rogers (1963) in indium.

must take into account the details of the phonon spectrum,


as determined, for example, by neutron diffraction. In particular it
Instead

taken to vary with

133

(1963) have closely verified the expected structure by tunneling experi-

cannot

with assigning to the phonons an average energy hcu

e(0) is

ek

must be remembered that

in

10r4

an experiment involving thermal

Ek

denned by XI. 9. This involves only the


However, the
details of the variation of the variation of the gap with e k can be
verified by tunneling experiments, in which the conductance dl/dV is
is

the energy

square of the gap, and

(Bardeen, 1961a,

962a; Cohen et al.

however pointed out that

Fig. 43

therefore always positive.

is

directly proportional to the density of states in the superconductor


1

10"

Tc /)

or electromagnetic absorption across the gap, the quantity actually

observed

10' 2

10' 3

have
one cannot use the standard

1962).Schrieffer<?/a/. (1963)

for tunneling

as being due to the direct passage of coherent


side of the insulating barrier to the other.

Cooper

pairs

from one

As has been elaborated by

Anderson (1963) and by Josephson (1964), the

relative

phase of the

superconducting wave functions on either side of the barrier has


physical meaning because it is a quantity conjugate to the number of

not quasi-particle eigenstates of the


individual metals making up the tunnel. Instead the appropriate

and because this number is not


As a result the energy of the
system depends on this phase difference, and in turn this gives rise
to a flow of pairs across the barrier in the absence of an applied

density of states to use

potential difference.

expression for the quasi-particle density of states. This

when an

electron tunnels

the initial

and

from one

is

m)

because

side of the barrier to the other,

final states are

is

Ek

Hvv^}

electrons

on each of

constant, that

is,

the

two

sides

not fully determined.

The Josephson current is very difficult to detect because it is quenched by a magnetic field of a few tenths of a gauss. It was first observed

Superconductivity

Microscopic theory of superconductivity

by Anderson and Rowell (1963) and by Shapiro (1963). The temperature dependence has been studied by Fiske (1964), following calculations by Ambegeokar and BaratofF (1963). Ferrell and Prange
(1963) have discussed the self-limitation of the Josephson current by
the magnetic field it generates itself, and De Gennes (1963) has derived
an expression for the current from the Ginzburg-Landau-Gor'kov

except for the upward deviation at the lowest temperature which was

34

mentioned

earlier (see Figure 28).

Further numerical predictions of the

=
yTc

equations.

A more

realistic cut-off

can also yield theoretical justification for

The following

table

TJ@. Such a correlation was


whose plot of energy gap values
against TJ0 for 17 different superconductors is shown in Figure 43.
Goodman used gap values deduced from empirical values of y, H
and Tc by combining XI. 1 3, XT. 8, and XI.20, and remembering that

is

taken from

BCS theory

Goodman

[7] (p.

= ffl^jAftO). This yields


(XI.32)

11.8.

The

One can

Element

yTc

Lead
Mercury
Niobium
Tin
Aluminium
Tantalum

3-65

listed in

Table

215

most experimental values agree with

obtain the electronic specific heat in the superconducting

phase by twice differentiating with respect to temperature the free


energy expression XI.21. At sufficiently low reduced temperatures,
for which 2e(7")

2-60
2-60
2-58
2-57
2-25

Zinc
Thallium

III.

specific heat

> k D Tc

this yields

closely

318
307

Vanadium

V3

Appropriate values of 2e(0)/( B rc ) are

how

CJTJ

k B Tc

(XI.35)

212) and shows

(1958),

include

2 43.

the apparent correlation of e(0) with

suggested by

135

Hi

this.

The theory

also yields that

0170,

(XI.36)

(see equation 11.15) that the predicted


polynomial expansion of the threshold field is

from which one can calculate


coefficient

of

2
t

in the

yTc

a2

107.

(XI.37)

where K and Kj, are first and third order modified Bessel functions of
the second kind. This simplifies in the temperature regions indicated

This agrees exactly with the experimental value for tin (VIII.3) and
closely with that for several other elements. In terms of the deviation

to the following exponential expressions:

of the threshold

yT

8-5 exp(-l -44TJT),

2-5

< TJT <

field curve from a strictly parabolic variation as displayed in Figure 23, any value of a 2 greater than unity corresponds
to a curve below the abscissa; only mercury and lead are seen to have

6,

(XT.34)

26 exp (- -62TJT),

< TJT <

1 1

deviations corresponding to values of a 2 smaller than unity.

The BCS calculations are based on an isotropic model, in which the


Kdoes not depend on the direction of A: and k
Pokrovskii (1961) and Pokrovskii and Ryvkin (1962) have investigated the effects of anisotropy on thermal and magnetic properties.
interaction parameter

Experimental data at

this

regions where they are in

time exist only in the

first

of these two

good agreement with the BCS

values,

10

'

Microscopic theory of superconductivity

Superconductivity

136

find that in anisotropic superconductors the specific heat ratio

They

137

transitions interfere either constructively or destructively depending

in Xl.35 should be smaller than 2-43, the quantity in XI. 36 larger than

on the type of scattering phenomenon

therefore the coefficient a 2 larger than 107. In the second


of the papers cited these results are compared with extensive experi-

interference in the case of electromagnetic interaction, such as the

mental data.

action which determines the nuclear relaxation rate.

The thermal conductivity in superconductors has been calculated


on the basis of the BCS theory for several of the pertinent mechanisms.
Bardeen et al. (BRT, 1959) and Geilikman (1958) have derived the

mental results expected

01 70, and

ratio of the electronic conductivity in the superconducting phase to

by impurity
have been well confirmed
experimentally, as was discussed in Chapter IX. The derivation of
BRTforthecaseofelectronicconductionlimitedbyphononscattering
(equation IX.5) does not, however, lead to the empirical behaviour.
Calculations by Kadanoff and Martin (1961) and by Kresin (1959)
are in better agreement, but further theoretical work is needed for this

that in the normal one

when

this is primarily limited

scattering (equation IX.6). Their results

involved. There

is

constructive

absorption of electromagnetic radiation, and the hyperfine inter-

The

experi-

two cases are therefore qualitatively


those which follow from a two-fluid model consideration of the total
number of electrons available as well as from the density of available
states. It has already been mentioned how this explains the observed
rise in the nuclear relaxation rate just below the critical temperature
in these

(Figure 32).

On

the other hand, the contributions of the

two

fere destructively in the case of the absorption of

transitions inter-

phonons, such as

occurs in the attenuation of ultrasonic waves. This destructive interference so decreases the probability of absorption that the effect of

the increase in density of states on either side of the gap

is

completely

conduction mechanism, in which quasi-particle life times may again


be important (see [7], pp. 272 ff.). According to calculations of
Tewordt (1962, 1963), however, these appear to have little effect on

wiped out, and the absorption just below Tc drops very sharply. For
low frequency phonons, ha> <^ 2e(0), the ratio of attenuation coeffi-

conduction mechanism.
BRT as well as Geilikman and Kresin (1958, 1959) have derived
the lattice conductivity limited by electron scattering. Experimentally

with an

this

cient in the superconducting


infinite slope,

-?
a

very difficult to separate out this part of the heat transport.


Where this has been possible (Connolly and Mendelssohn, 1962;
it is

Lindenfeld and Rohrer, 1963) the results have been in general


agreement with the theoretical predictions.

This function

is

and

shown

is

and normal phase

o.J<x

2/{l+exp[2e(T)lk B T]}.

in

Tc

drops below

given by

(XI.38)

Figure 44, which includes experimental

points on both tin and indium by

Morse and Bohm

(1957). It should

be contrasted with the theoretical prediction for nuclear relaxation


11.9.

One

Coherence properties and ultrasonic attenuation


BCS theory arises as a

of the most striking predictions of the

consequence of the pairing concept, and experimental verification of this point is thus of particular importance. In a normal metal
the scattering of an electron from state k t to state k' t is entirely
independent from the scattering of an electron from - k | to -k'
or of any other transition. The coherence of the paired electrons in
the kf and -k j states in the superconducting phase, however,
direct

makes
(see

two transitions interdependent. The details of the theory


pp. 212-24) show that the contribution of the two possible

these

[7],

rate,

shown

in

Figure 32.

Measurements of the ultrasonic attenuation


in different crystal directions

in single crystals of tin

has yielded very convincing demonstra-

When an electron absorbs a


momentum can both be conserved only if the

tion of the anisotropy of the energy gap.

phonon, energy and

component of the electronic velocity parallel to the direction of sound


propagation is equal to the phonon velocity, which is the velocity of
sound S. Since, however, the Fermi velocity of the electrons, v is
,

several orders of magnitude larger than S, this

electrons which

is

possible only for

move almost at right angles to the direction of sound

138

Superconductivity

Microscopic theory of superconductivity

propagation. Thus a measurement of the attenuation of sound propagated in a particular crystalline direction involves only those electrons

To

whose velocity

state in the presence of an external field,

a thin disk at right angles to this direction. The value of the energy gap appearing in equation XI. 38 i > thus
one averaged over this particular disk. Such measurements have been
1.0

directions

lie in

11.10. Electromagnetic properties

describe the many-particle

interaction as a perturbation,

wave function of the superconducting

BCS treat the electromagnetic

and obtain an expansion

in terms of the
spectrum of excited states in the absence of the field. This wave function is then substituted into an equation of the form III. 20 to calculate
the current density. Mattis and Bardeen (1958) and also Abrikosov
et al. (1958) have expanded this to treat fields of arbitrary frequency.

The

result of the former has been used by Miller (1960) to calculate


values of cri/o n and of a 2 lan over a wide range of temperatures and

TIN
0.8

139

x INDIUM

frequencies. His calculations are in excellent agreement with all the

experimental results using weak

0.6

in

Chapter X,

to

its

if the

energy gap

is

fields at

high frequencies, described

taken as a parameter to be adjusted

empirical value.

The treatment of a magnetic field as a perturbation in the BCS


formulation makes it very difficult to extend it to high field values
(H m c). This can be done more readily from a representation of the

0C n

BCS

0.4

ideas in terms of Green's functions which has been developed

by Gor'kov (1958). A simplified version of this method has been presented by Anderson (1960). The electromagnetic equations occurring
in this formulation were shown by Gorkov (1959, 1960) to be equivalent to the Ginzburg-Landau expressions in the region near Tc and
under circumstances where A > . As was pointed out in Chapters V
and VII, Gor'kov showed that the energy gap is proportional to the
G-L order parameter, so that the dependence of the latter on temperature, magnetic field, and co-ordinates, also applies to the former.

02

The successful application of these ideas to a number of experimental


results

has been mentioned

in

Chapter VII.

An apparent shortcoming of the original BCS

performed on variously oriented tin single crystals both by Morse et al.


(1959) and by Bezuglyi et al. (1959). Their results are in good agreement and are summarized in the following table:

Wave

vector q

2e(0)/k B Tc

parallel to [001]

3-2 01

parallel to

4-3

[1 10]

perpendicular to [001] and 18 from [100]

0-2

3-5

01

treatment is its lack


of gauge invariance. It was suggested by Bardeen (1957) and worked
out by various authors that this can be remedied by taking into
account the existence of collective excitations. A discussion of this
with full references is given in [7] (pp. 252 ff.).

There exists as yet no fully satisfactory explanation that the Knight


superconductors does not vanish in any of the elements in
which is has thus far been studied mercury (Reif, 1 957), tin (Androes
and Knight, 1961), vanadium (Noer and Knight, 1964) and aluminium
shift in

'40

Superconductivity

(Hammond and

The Knight shift is defined as the fractional difference in the magnetic resonance frequency of a
nucleus
in a free ion and the same nucleus in a metallic medium. It is
due to
the field at the nucleus created by the free electrons, and is usually
Kelly, 1964).

CHAPTER

XII

taken to be proportional to the electronic spin susceptibility. A literal


interpretation of the Cooper pairs of opposite spin would lead
one
to expect that in a superconductor this susceptibility and hence
the

12.1. Introduction

should vanish at 0K. A number of authors (see


[7],
pp. 261-263; Anderson, 1960; Suhl, 1962; Cooper, 1962) have
suggested why this may actually not be the case, and although none
of these explanations appears fully adequate, they have shown that

Ever since the discovery of superconductivity there have been many


searches for new superconducting materials. Roberts (1961) has
recently listed more than 450 alloys and compounds with critical
temperatures ranging from 016 up to 18-2K. In the appearance of

Knight

shift

the

Knight shift

the

BCS

offers

no fundamental disagreement with the idea of

theory.

(1962, 1964) deduce

from the temperature variation of the Knight


that in the superconducting state the dominant
to the d-electron spin does vanish, as the simple

vanadium

contribution due

superconductivity

among

Compounds

these substances there exist certain regu-

which were discovered by Matthias (1957) and to which


reference was made in Chapter I. One might consider as an ultimate
goal of any complete microscopic theory the ability to derive these
Matthias rules from first principles. This would be equivalent to being
larities

It is, furthermore, possible that the Knight shift in


some of these
elements is not primarily due to spin paramagnetism. Clogston et al.

shift in

Superconducting Alloys and

theory would predict. This, however, leaves a

finite Knight shift


due to orbital paramagnetism which involves electrons too far from
the Fermi surface to be involved in pairing. Thus this contribution

to the Knight shift in vanadium

not affected by the superconducting


transition, and perhaps the orbital part is the dominant one
in tin
and mercury.
is

some precision the actual critical temperature


of any superconductor. At the moment our understanding of superconducting and of normal metals is still very far from such
able to calculate with

achievements.

One of

the

many ways

of increasing this understanding

is

sys-

tematic study of superconducting alloy systems in which solvent or


solute are used as controlled parameters. This has been done in a

number of experiments.
12.2. Dilute solid solutions with

Serin, Lynton,

non-magnetic impurities

and collaborators (Lynton

et

al.,

1957; Chanin et

al.,

1959) have investigated the superconducting properties of dilute


alloys of various solutes into tin, indium, and aluminium, up to the
limit of solid solubility. For low impurity concentrations, of the order

of a few tenths of an atomic per cent,


reciprocal electronic

mean

Tc decreases linearly

with the

free path, independently of the nature of

When plotted against the reduced co-ordinate //, where


the coherence length of the pure solvent, the fractional change

the solute.

is

in

Tc is the same for elements as different as Sn and Al (Serin,

1960).

shown in the initial portions of both curves in Figure 45. The


existence and the magnitude of this seemingly general effect lend
This

is

141

Superconductivity

Superconducting alloys and compounds

strong support to Anderson's model of impure superconductors

anisotropic energy gap indeed leads to a lowering of Tc of the observed


magnitude. Hohenberg (1963) has calculated the dependence of Tc

142

(Anderson, 1959). He suggested that the energy gap anisotropy is


smoothed out by impurity scattering and disappears when the electronic mean free path is comparable to
fiv

so-

MO)

the energy gap,

and the density of

on the concentration of

states

impurities.

This general mean free path effect on Tc has also been found in
tantalum by Budnick (1960). It has been verified by using a number
of different ways of scattering the electrons by mechanical deforma-

143

and cold work in aluminium (Joiner and Serin, 1961), by size


effects in indium (Lynton and McLachlan, 1962), by quenching
(De Sorbo, 1959), electron irradiation (Compton, 1959), neutron
bombardment (Blanc et al., 1960), and by using isoelectronic ternary
compounds (Wipf and Coles, 1959) in tin.
Figure 45 shows that for P /l > 1 the effect on Tc deviates from the
initial linear decrease in a way which depends on whether the solute
tion

+.02

ELECTRONEGATIVE

+.01

Aoll

is

-.01

electropositive (valence smaller than that of solvent) or electro-

al. (1961) have extended such


measurements to higher concentrations. They found that for both
types of impurities Tc ultimately rises to values above that of the

negative (valence larger). Chiou et

-02

and were able to repissent the variation of Tc with impurity


all cases by an empirical relation containing two

solvent,

-04

combination.

-05-

parameters an average phonon frequency m q (which is proportional


to the Debye temperature ), the density of normal electron states at

concentration in

parameters adjusted according to the particular solvent-solute

According to the BCS theory (equation XI.29),

Fig. 45

This should then result in a lowering of

Tc

by an amount approxi-

resonance in aluminium (Masuda and Redfield, 1960a, 1962) and ultrasonic and infrared absorption in tin (Morse et al., 1959; Bezuglyi,e/ al.,
1959; Richards, 1961) have

shown

by about 10 per cent from

be lowered by about

confirm

this very well.

Kadanoff

per cent

its

that the

gap

in these elements

average value, so that Tc should


P The measurements of Tc

when /

Recently Caroli et

al.

(1962),

Markowitz and

and Tsuneto (1962) have shown in terms of the


microscopic theory that Anderson's idea of the smoothing of an
(1963),

is proportional to the Sommerfeld y),


and the BCS interaction parameter V. Specific heat measurements on
tin alloys have recently enabled Gayley et al. (1962) to find the effects
of the addition of indium, bismuth, and indium antimonide on the
values of y and of for tin. One can use equation XT.29 to calculate
the corresponding change in Tc This seems to account for most of
the difference in the behaviour of electropositive and electronegative
solutes, at least in the case of indium and bismuth, but not for the

the Fermi surface, N(0) (which

mately equal to the square of the fractional anisotropy. Nuclear

varies

Tc depends on three

increase in
increase

is

Tc at high solute concentrations. One concludes that this


mainly due to

effects

of alloying on the interaction

energy V.

Any attempt to calculate Kin the presence of impurities has to take

144

Superconducting alloys and compounds

Superconductivity

wave vectors k are no longer


good quantum numbers. Hence the question arises of the criterion
for pairing of the electrons. Abrahams and Weiss (1959) and
Anderson (1959) have pointed out that in impure superconductors
Cooper pairs are formed of two electrons the wave functions of which
are identical except for the reversal of the time co-ordinate, and which
have the same energy. The former authors have used this to deduce
into account that with scattering the

145

Also the absorption edge of infrared radiation at the


gap frequency can be very sharp (Ginsberg and Leslie, 1962).
Detailed magnetization curves (Lynton and Serin, 1958) however,

et al., 1956).

show

that the transitions for such alloys are nevertheless not fully

46 for 311 per cent In-Sn cylinders


In decreasing field the magnetic
moment does not attain its full diamagnetic value until vanishes.
This indicates that flux is initially trapped, but then leaks out as
reversible, as

shown

in Figure

transverse to an external

field.

suggested by Faber and Pippard (1955b).


12.3. Compounds with magnetic impurities
Matthias and collaborators have traced the occurrence of superconductivity in a large number of compounds containing paramagnetic and ferromagnetic impurities (Matthias, 1960). Their

can be summarized as follows:


Ferromagnetic transition elements with 3d electrons (Cr, Mn, Fe,
Co, and Ni) put into fourth column superconductors (Ti, Zr) raise
results

Tc more than
Ad

do corresponding amounts of transition elements with

electrons (Re, Rh, Ru, etc.) (Matthias and Corenzwit, 1955;

Matthias et al., 1959b). At the same time magnetic measurements on


Ti-Fe and Ti-Co alloys indicated the absence of localized moments.
The effect of the 4d electrons can be attributed to the increase in
Applied Field

He

the

number of valence

electrons per

atom toward
The

particularly favourable for superconductivity.

Fig. 46

3d electrons

is

is

difficult to esti-

mate. Anderson (1959, 1960) has discussed the general implications


of the use of time-reversed wave-functions. Detailed microscopic
calculations of the impurity effects on the superconducting parameters

have been attempted by Caroli


Kadanoff(1963).

et al.

(1962) and by

Markowitz and

annealed solid
conductors according to several

criteria.

Transitions occur within a

few millidegrees, and very little flux remains in suitably oriented


cylindrical samples after an external field has been removed (Budnick

same reason adding Fe

(3d electrons) to a Ti 6 V 4 compound lowers Tc less than does an


equal amount of Ru (4d electrons): in both cases Tc is decreased
.

because the number of valence electrons per atom rises beyond five,
but with Fe the apparent magnetic interaction counteracts this in
part. It must be pointed out, however, that ferromagnetic transition
fifth column superconductors (Nb,
V) lower Tc in approximate agreement with the expected effect due to
the change in valence electrons per atom (MUller, 1959). There does
not appear to be any added effect due to the magnetic nature of the
impurities. Why such effects should appear with fourth column
metals but not with fifth column ones is far from clear, as in neither

elements with 3d electrons put into

to note that the work

on carefully homogenized and


solutions has shown these to be 'well-behaved' super-

It is interesting

a number

extra rise with

attributed to a magnetic electron-electron inter-

action favouring superconductivity. For the


several impurity effects, the magnitude of which

five,

146

Superconducting alloys and compounds

Superconductivity

958b, 1959a).

case are there any localized magnetic moments associated with the 3d
solute atom.

et al.,

Quite recently Cape (1963) has measured the electrical and magnetic properties of very carefully prepared alloys of Ti containing

effective

0.2 to 4 at

Mn. Depending on

the

method of preparation

these

moments

is

147

decrease, for each per cent

is correlated with the spin rather than with the


magnetic moment of the solute. This is shown in Figure 47
in which - ATC for each per cent, the spin, and the effective moment

of rare earth impurity,

/x

eff are

plotted for the different rare earths.

Tc

A higher effective moment,

hep phase, which however is not

consistent with the usual suppression of

centage of neodynium, which has the same spin but a smaller moment.

exist only in the

superconducting. This

The magnitude of this

perhaps for the same reason as in


the case of the 3d impurities in fourth column metals erbium, with
spin 3/2 and large moment lowers Te less than does an equal per-

specimens are either in a single hexagonal close packed (hep) phase,


or contain an admixture of a second, body centred cubic (bec) phase.
Localized

in fact, appears to tend to raise

superconductivity by impurities retaining localized

moments (see
The non-magnetic bec phase, on the other hand, has a transition temperature which is raised above that for pure Ti by an amount
commensurate with the increase in the number of valence electrons.
Hake et al. (1962) had earlier deduced from their measurements of
transport properties that the hep phase of Ti-Cr, Ti-Fe, and Ti-Co
also carried localized magnetic moments. In addition there is calorimetric evidence (Cape and Hake, 1963) that in Ti-Fe samples only a
small fraction of the volume is superconducting. These results throw
considerable doubt on Matthias' speculation that iron-group impurities which do not carry a localized magnetic moment enhance
superconductivity by means of a magnetic interaction between
below).

electrons.

While 3d impurities in fifth column metals (for example Nb) do not


show any evidence for a localized moment, they do when put into
sixth column metals (for example, Mo), and in fact Matthias et al.
(1960) found that the change in behaviour occurs in

Nb-Mo solutions

Mo. One would therefore


some special effects on Tc to appear in 3d compounds with
sixth column metals. Until the recent discovery of the superconductivity of Mo, no such metal was known to be superconducting. For
that reason, this effect was studied on superconducting Mo
8 Re 2
and indeed small amounts of 3d impurities lower Tc far more than one
would expect from valence effects. It is in fact this which made the disat a concentration of about 60 per cent

expect

Fig. 47

All these compounds containing 4/electrons show ferromagnetic


behaviour at somewhat higher concentrations of the rare earth
solutes, with the Curie temperature rising with increasing number of
4/ electrons. Such dilute ferromagnetism has not been observed for

compounds with 3d electrons which


interaction

is

indicates that the s-f magnetic

rather long range, while the d-d one

is

a short-range

covery of the superconductivity of Mo so difficult: a few parts per


million of iron are enough to depress Tc below the measurable range
(Geballe et al, 1962). A less abrupt decrease in T is obtained when
c

rare earth elements with 4/electrons are put into lanthanum (Matthias

inter-action effective only through nearest neighbours,

which

is

impossible in dilute solutions (Matthias, 1960).


Interesting analogies in the variation of the Curie temperature and
the superconducting critical temperature are found by investigating
the magnetic characteristics of so-called Laves compounds AB2
,

where B is germanium or a noble metal (Ru, Os,

Ir,

Pt)

and A is either

148

Superconducting alloys and compounds

Superconductivity

a rare earth with 4/ electrons (A') or one of the group Y, Sc, Lu, or
La {A"), none of which contain 4/ electrons (Suhl et al., 1959). A'B2
is

always ferromagnetic,

A"B 2 always

superconducting. Comparing

149

an amount proportional to the


between the initial and final electron

electron-spin scattering interaction by

reciprocal energy difference


state.

In the normal state this difference can be arbitrarily small, in

the Curie temperatures of the former with the critical temperatures of

the superconducting case this difference cannot be smaller than the

on spin and on the number

energy gap. As a result the free energy of the normal phase is lowered
more than that of the superconducting one, and the onset of super-

the latter one finds a similar dependence

of valence electrons per atom. This

but one of a number of interesting correspondences which Matthias has found between superconis

conductivity therefore occurs at a lower temperature. Suhl and

and ferromagnetism. There are, for example, several groups


of isomorphous compounds which are either superconducting or
ferromagnetic (see, for example, Matthias et al., 1 958a Compton and

Matthias ignore the small changes

Matthias, 1959; Matthias et al., 1962). Also, the appearance of locali-

(1960)

ductivity

zed

moments when

a ferromagnetic impurity

netic transition element

electrons in a

manner

is

put into a non-mag-

seems to depend on the number of valence

similar to the criterion for the appearance of

superconductivity (Matthias, 1962). Matthias hasfrequentlysuggested

may

that an electron configuration favourable to superconductivity


also be favourable to ferromagnetism.

The possible coexistence of superconductivity and ferromagnetism


in the

same substance has been investigated in lanthanum-rare earth


compounds (Matthias et al., 1958b) and in Laves compound

binary

mixtures (A'i^ x A%)B

(Matthias et

al.,

1958c; Suhl et a/.,1959).

Both magnetic (Bozorth et al., 1960) and calorimetric measurements


(Phillips and Matthias, 1960) have shown that ferromagnetism and
superconductivity occur in the same sample, but the evidence
entirely conclusive in ruling out the possibility that these

is

not

two phe-

nomena merely exist side by side in different portions of the specimen.


Anderson and Suhl (1959) have shown that the actual coexistence of
ferromagnetism and superconductivity on a microscopic scale is
energetically possible if the ferromagnetic alignment occurs in the
form of extremely small domains probably of the order of 50 A. They

V,

in the interaction

and as a result their prediction (dTJdc -* 0)

impurity concentrations

show

is

matrix element

for very small magnetic

probably wrong. Abrikosov and Gor'kov

that magnetic impurity effects

linearly with impurity concentrations.

on

V initially

lowers

Tc

Atmuch higher concentrations

Suhl and Matthias find that S7y3c->co, which is supported by


experiment (Hein et al., 1959). Similar calculations have been carried
out by Baltensperger (1959). Suhl (1962) has recently reviewed this

and

similar work.
Abrikosov and Gor'kov (1960) as well as De Gennes and Sarma
(1963) show that magnetic impurities will lower the energy gap more
rapidly than the transition temperature. There should thus be a
range of concentration for which the alloy has a finite critical tem-

perature at which

its

DC resistance disappears without the existence

of an energy gap. Indeed Reif and Woolf ( 1 962) have found this para-

They measured the electrical resistance as


number of lead and indium
film containing magnetic impurities. The gap decreased twice as
rapidly as the transition temperature, and an indium film containing
1 at
Fe, for example, had no resistance below 3K but a perfectly
ohmic tunneling conductance.
doxical behaviour to exist.

well as the tunneling characteristics of a

Phillips (1963) has pointed out that such gapless superconductivity

Suhl and Matthias (1959) have treated the general problem of the
lowering of Tc due to the presence of magnetic impurities by extending
an argument of Herring (1958), according to which the polarization

does not violate any fundamental principle. The spin-flipping scattering of the conduction electrons by the magnetic ions gives the former
a very short lifetime. This broadens the electron states, particularly
those nearest the gap, so much as to spread the states into the gap.
At a certain impurity concentration states will have spread throughout

due to the coupling of the conduction electrons with the spins of the
paramagnetic impurity ions lowers the free energy in both the normal
and in the superconducting phases. The free energy is lowered by each

making it disappear. The density of states,


however, will still have maxima at what used to be the edge of the gap,
and as long as this exists the material will have zero DC resistance.

call this 'cryptoferromagnetic' alignment.

the width of the gap,

Superconductivity

Superconducting alloys and compounds

have been worked out by Skalski et al. (1963). In


terms of the Ferrell-Glover rule and the frequency dependence of the
conductivities shown in Figure 37, the situation can be described by
saying that at some concentration there is a finite real conductivity a

of the gradient of h k which, however, have yet to be justified from


more fundamental considerations. From these he concludes that a
normal layer sandwiched between two superconductors will itself be
-5
supreconducting if it is no thicker than about 10
cm. A superconducting layer between two normal metals will remain normal up to

50

The

details of this

For some further range of impurity concentration,


however, o-j will still be less than aN for cu/wg < 1 resulting in a reduced
delta function at the origin. At even higher concentrations, ct, x aN
for all frequencies. The sum rule is now satisfied without an infinite
DC conductivity, so that the metal is normal in every respect.
at all frequencies.

some similar critical thickness.


Cooper (196 ) has given a more intuitive argument for the possibil1

ity that the

may be
He emphasized

superconducting properties of thin metallic films

strongly affected by direct contact with other metals.


that in the BSC theory

12.4.

Superimposed metals

A recent series of experiments by Meissner (see Meissner,

1960 for full

one must clearly distinguish between the range


of the attractive interaction between electrons, and the distance over
which as a result of this interaction the electrons are correlated into
Cooper pairs. The range of the interaction is very short (10~ 8 cm) ; the
of the wave packets of the pairs, on the other hand,

references) has revived interest in the question whether thin layers of

'size'

superconducting material deposited on a normal metal would themselves become normal, and whether conversely sufficiently thin layers

order of the coherence length, that is, 10

of normally non-superconducting metal would become superconducting when in contact with a superconductor. Such superimposed metals
differ

from the sandwiches used

in the tunnelling

experiments by the

absence of an insulating layer.

Parmenter (1960a) has constructed a theory for such direct metallic


contacts in which he attempts to introduce directly into the

BCS

extend a considerable distance into a region in which the interaction


between electrons is not attractive. Thus when a thin layer of superconducting material is in contact with a layer of normal metal, the

zero-momentum

pairs

formed because of the attractive interaction in


As a result the ground

the superconductor extend into both layers.

a spatial variation of the parameter h k appearing

state energy of this thin bimetallic layer

of h k 12 and of

-h k ) 112

, broadly analogous to the extra energy term


V.6 in the Ginzburg-Landau theory. Near the boundary of a superconductor this leads to a significant variation of the energy gap over
-6
distances which are of the order of 10
cm, that is, two orders of

(1

magnitude smaller than the coherence length The configurational


surface energy resulting from this gap variation turns out in this
-5
theory to be about 10
cm (Parmenter, 1960b) which is an order of
magnitude smaller than the generally accepted value.
To investigate the behaviour of superimposed metallic layers under
this theory, Parmenter postulates a set of plausible boundary conditions involving the continuity of h k and of the normal component
.

of the

cm. This, as Cooper points


out, is analogous to the difference between the range of the nuclear
interaction and the much larger size of the resulting deuteron wave
packet. Because of this long coherence length the Cooper pairs can

formulation a dependence of the energy gap on position by postulating


in equation XI.5.
This adds to the kinetic energy portion of the ground state energy (the
first term in equation XI.5) terms involving the square of the gradient

is

-4

is characterized by some average over both metals of the parameter N(0) V, which in turn determines the energy gap of the layer and its transition temperature,

according to equations XI. 8 and XI.29. The form of this average of


course depends on the nature of the boundary between the two metals
1

the better the contact, the more effective is a superimposed layer in


changing the properties of the substrate. Regardless of how one
accounts for this, one would expect the average to depend also in some

manner on the
normal

energy gap width and the transition


the values they

would have

Similarly a combination of
to

have a
11

The thicker the


and the more the
temperature are decreased from

relative thickness of the

two

layers.

layer, the smaller the average interaction,

Te

if

only the superconductor were present.

two superconductors would be expected


somewhere between the Tc values of the two materials,

Superconductivity

52

varying from one extreme to another as the relative thickness of the

two layers is varied.


These qualitative conclusions presuppose that both layers of the

CHAPTER XIII

bimetallic film are sufficiently thin so that the coherent electron pairs

extend over the entire volume. One expects the critical thickness for
this to be of the order of the coherence length, although it is not clear

whether this should be the ideal value or the mean free path limited
value (f). If one of the two superimposed metals is much thicker than
whatever critical length is appropriate, then presumably the average

Superconducting Devices

interaction

is

determined by the ratio of the smaller thickness to the

critical length.

The experiments of Smith et al.


on or between

silver films

(1

96 1 ) with lead films of about 500 A

varying from 100 to 7000

A indicate

a de-

crease of the transition temperature of the lead with increasing silver

Wilhelm ( 935) and


of Meissner (1960). Similar results have been obtained by Hilsch and
Hilsch (1961 ) with combinations of copper and lead films. These agree
with the calculations of De Gennes and Guyon (1962) and the
more extensive treatment of Werthamer (1964). However, the work
of Rose-lnnes and Serin (1961) has shown that results can be strongly
influenced by varying evaporation procedures, even under conditions which quite preclude ordinary bulk intermetallic diffusion.
Interpenetration of metals seems to occur quite readily with superimposed layers, possibly by the mechanisms of surface and defect
diffusion. Because of this the experimental situation is at this time far
from clear.
thickness, supporting earlier work of Misener and

13.1.

The

Research devices

characteristics of superconductors

have for a long time already


been put to use in many low temperature experiments. It is very common to use niobium wires, for example, in electrical connections to
samples which one wishes to isolate thermally as well as possible. Such
wires are superconducting with a high critical field throughout the
entire liquid

helium temperature range, and combine low thermal

transport with perfect electrical conductivity.

heat switches at temperatures below

0TK

The use of lead wires as


has been mentioned in

Chapter DC.

More

specialized research devices using superconducting

com-

ponents of varying complexity have been suggested or used frequently,

and it is possible in this cursory survey to mention only a few of these.


A number of such devices have been developed to detect very small
potential differences, as occur, for instance, in studies of thermo-

powers. Pippard and Pullan (1952) improved earlier designs


by Grayson Smith and co-workers (Grayson Smith and Tarr, 1935;
Grayson Smith et al., 1936) by using a single turn of superconducting

electric

wire to construct a galvanometer capable of detecting e.m.f.s of

10" ,2 V. With a resistance as low as 10~ 7 ohm this required a current


5
sensitivity of only 10~ amp the time constant L/R was kept short by
;

the single turn design which reduced the effective inductance.

conducting magnetic shield made possible controlling

001

A super-

fields as

low as

gauss.

A different approach to the measurement of very small potentials


was suggested by Templeton (1955b) and by De Vroomen (De
Vroomen, 1955; De Vroomen and Van Baarle, 1957). These authors
designed 'chopper' amplifiers in which the small

d.c. signal is

con-

verted into an alternating one by passing through a superconducting

wire which

is

modulated into and out of the normal


153

state

by being

Superconductivity

Supei conducting devices

placed in an alternating magnetic field. The resulting oscillating


potential across the wire is then amplified in a conventional manner.
10~ ' ' V.
These devices can operate stably with a noise level at about

have been observed for many


and for strained or impure samples of the superconducting
elements. For niobium published values of the critical field at 0K
vary from about 1950 to 8200 gauss. Quite recently Kunzler et al.
(1 961 b) discovered Nb s Sn to have a critical field of about 200 kgauss,
and similar critical fields have since been found in other substances.
These seem to be either intermetallic compounds of the /3-wolfram
structure, or body centered cubic alloys. When suitably prepared

154

Templeton (1955a) has also designed a superconducting reversing


switch to suppress undesirable thermal voltages in measurements of
6
potential differences of the order of about 10~ V.
temperature experiments as well as superconducting
magnets require rather high direct currents at very low voltages.
To avoid the use of thick electrical leads which would bring too

Many low

heat into the helium dewar, Olsen (1958) has designed a


superconducting rectifier and amplifier which, together with a low

much

temperature transformer, allows one to feed in a low alternating


current through thin leads. The rectification occurs as the current
flows through a superconducting wire placed in an external, nearly
critical field, such that the field due to the current in one direction
is sufficient

to

make

the wire normal during about one-half of each

tively large values of the critical field

alloys

these materials remain superconducting while carrying current densities

as high as 5 x 10

amp/cm2

in fields

almost up to the

critical

value.

The

critical fields

thermodynamic

of these materials are

critical fields

much

too high to be the

as defined by equation

II.4.

How-

have discussed two reasons why superconductivity can persist in a given specimen to fields higher than
c One
possibility is that the material is sufficiently inhomogeneous, so as to

ever, earlier chapters

display the characteristics of a 'Mendelssohn sponge', as discussed

cycle.

D. H. Andrews

et al. (1946)

made use of the change

in resistivity

at the superconducting transition in designing a bolometer.

A different

superconducting radiation detector has been suggested by Burstein


et al. (1961) who pointed out that a tunnelling device (Chapter X)
suitably biased would respond to absorption of electromagnetic
radiation in the microwave and submillimetre range.

RF

detection

with a tunneling device has been achieved by Shapiro and Janus


(1964). For work at high frequencies superconducting metals may
also be used to construct resonant cavities of extremely high Q. This

has been discussed by Maxwell (1960), and preliminary experiments


have been reported by Fairbank et al. (1964) as well as by Ruefenacht
and Rinderer (1964). Thought is also being given to the use of
superconducting cavities in high energy proton linear accelerators
(Parkinson, 1962; Fairbank et al., 1964). Many of the devices listed
in this section as well as others

have recently been discussed by

in Section 7.2.

In such a specimen superconductivity persists in a

filamentary structure, the dimensions of which are

than the penetration depth.


conducting to a

1931,

H >H
s

c,

much

smaller

a result, the filaments remain superas given by equation VII.8.

On

the

mechanism for high field superconductivity was discussed in Chapter VI, where it was shown that superconductors of the second kind remain in a superconducting mixed
other hand, a quite different

up to U c2 > c (Abrikosov, 1957). Superconductors of the


second kind are materials which may be quite homogeneous and
which have a negative surface energy, generally because of their very
state

Goodman

(1961) was the first to


mechanism to explain the high
critical field, found by Kunzler and others, and there is convincing
evidence that this is indeed the case (see for example, Berlincourt and
Hake, 1963). The specific heat results of Morin et al. (1962) on V3 Ga

short electronic

mean

free path.

suggest the possible relevance of this

the second kind


13.2. Superconducting

field

As

are consistent with the behaviour expected for superconductors of

Parkinson (1964).

As early as

155

De Haas and Voogd found critical fields as high as

15 kgauss in some lead-bismuth alloy wire. Other instances of

(Goodman,

observed by Berlincourt and

magnets
rela-

limit for

1963b), and so are the critical fields


(1962, 1963) in the low current

Hake

a number of high field alloys and compounds. Hauser


Swartz (1962) have further shown that the

(1962) as well as

Superconductivity

Superconducting devices

magnetization curves of suitably prepared specimens of various


compounds are consistent with the identification of these materials

of 4-3 kgauss, and since then the interest in the subject has grown

156

as superconductors of the second kind.


role of defects

systematic study of the

on the magnetization curve has been

carried out by

Livingston (1963, 1964).

However, Gorter (1962a, b) has pointed out that a homogeneous


superconductor with uniform negative surface energy cannot in the
presence of a transverse magnetic field carry the high current densities
which are actually observed in most of the compounds and alloys
under discussion. This can best be understood in terms of the vortex
structure of the mixed state which is created by the external field (see
Section 6.6). When a current passes through the specimen at right
angles to the vortices, it interacts with the latter so as to push them
out of the specimen. This, as was mentioned in Section 6.6, can be
prevented only if the vortices are pinned down by local variation of
the surface energy, as would be present if the specimen were inhomogeneous. Indeed, there
capacity

is

is

157

much scientific and technical activity in a large


number of laboratories. Kunzler?/ al. (196 la) and others used Mo 3 Re
to wind solenoids producing up to 5 kgauss; much higher fields were
achieved soon thereafter as a result of work with Nb 3 Sn (Kunzler
et al. 1961b), Nb2 Zr (Kunzler, 1961; Berlincourt et al., 1961) and
NbTi (Coffey et. al., 1964). Solenoids wound of these materials
have produced fields up to 100 kgauss, and both suitable
explosively, with

superconducting wire as well as entire solenoid assemblies have

become commercially available. At the moment, the size of these


measured in inches, but large-scale superconducting coils
producing fields well in excess of 100 kgauss seem quite feasible.
Kropschot and Arp(1961) have recently reviewed the subject of superconducting magnets, and have discussed the considerable technical
and economic advantages of such devices. Much information can also
be found in [11] as well as in Berlincourt (1963).
is still

much evidence that the high current carrying

associated with the presence of dislocation in cold-worked

specimens (Hauser and Buehler, 1962). Annealed samples

may

still

13.3. Superconducting

Much

computer elements

research and development

work

is

currently being devoted to

have a very high critical field while carrying a low current density,
but turn normal when the latter is increased. Rose-Innes and Heaton
(1963) have used Ta-Nb wire to show very strikingly how sample
treatment can change the current carrying capacity without changing

attempts to use superconductors both as switching devices and as


memory storage elements in electronic computers. The basic idea for

the critical

further able to carry high current densities in high


through cold work they are made to contain a high density
of dislocations which pin down the current carrying regions. A nearly
uniform distribution of these dislocations explains why the critical

niobium wire wound on to a thicker (0009 in.) tantalum wire. A


sufficiently large current through the former, called the control
winding, can quench the superconductivity of the latter, called the
gate. The two materials are chosen because the convenient operating
temperature of 4-2K is only a little below the critical temperature
of Ta, but much lower than that of Nb, so that a control current
sufficient to 'open the gate' is still much less than the critical cur-

current increases as the cross-sectional area of the specimen (Lock,

rent of the control.

1961a; Hauser and Buehler, 1962).

large so as to

Thus

field.

the present picture of high field superconductors

basically they are materials characterized

energy.

is

that

by a negative surface

They are

fields if

The ability of some superconductors to carry high current


in

high

fields,

Yntema

densities

of course, suggests their use in the winding of magnets.

(1955) described a superconducting solenoid

wound with

niobium wire and producing up to 7 kgauss, but this received little


attention. In 1960 Autler wound a niobium solenoid creating a field

a superconducting switching element originated with Buck (1 956) who


invented the cryotron. This consists of a layer of thin (0003

The diameter of the gate

is

in.)

furthermore kept

maximize the amount of gate current, Ig which can be


,

controlled by the control current, Ic Calling


.

the critical field of

the tantalum gate at the operating temperature, and

D its diameter,

then

(/,)m

H ttD,
c

(xni.i)

Superconducting devices

Superconductivity

158

flows only

and

open,
*c

(XIII.2)

= number of turns/unit length

of control winding. Thus

(XIII.3)

This

is

the 'gain' of the cryotron, which must be kept at a value

(a)

59

cryotron A or B is open, (b) cryotrons A and B are


A nor B are open. More complicated logical circuits

Buck (1956) as well as in review


by Haynes (1960), and by Lock (1961b).

are discussed by
(1959),

where n

if:

(c) neither

articles

by Young

Basically all these cryotron circuits consist of a number of parallel


superconducting paths between which the current can be switched by
the insertion of a resistance into the non-desired branches. Under
steady-state conditions the power dissipation is zero as long as there
is

always at least one path which remains superconducting. The speed

greater than unity in order that the gate current of one cryotron can

be used to control another.


.+

INPUT
"ONE"

INPUT

"ZERO"

37
;-!

READ
"ZERO"

READ
"ONE"

Fig. 48
Fig. 49

A great variety of logical circuits can


this reciprocal control

cuits contain the basic

be built up by making use of


number of cryotrons. Most of these cirflip-flop or bistable element, shown in Figure
of a

through this element can flow in either one or the other


branch and, once established in one, will flow in it indefinitely since
it makes the other one resistive. The choice of branch can be dictated
48. Current

by placing a further cryotron gate in series with each branch, and conan outside signal, which can 'open the gate', making
the corresponding branch resistive and forcing the current into the
other path. This is shown in Figure 49, which also indicates that if
each branch also controls the gate of a read-out cryotron, the position
of the bi-stable element can be read. Figure 50 shows other basic
logical circuits using cryotrons; the current through the heavy line

trolling this by

with which the resistance can be inserted, that is, the speed with which
a given gate can be made normal, depends on the basic phase transition time

and is small enough

( as

10~

10

sec) not to be a limiting factor

at this time (see, for instance, Nethercot, 1961

Feucht and Woodford,

On the other hand, the switching time from one current path
to another is determined by the ratio L/R, where L is the inductance
1961).

of the superconducting loop

made up

of the current paths, and

R the

resistance introduced by an opened gate. The usefulness of wirewound cryotrons is severely limited by the fact that this time is no less
than 10~ 5 sec, even if the gate consists of a tantalum film evaporated

on

to an insulating cylinder. Because of this

development

effort

is

directed toward

all

making

current research

and

thin film cryotrons

Superconducting devices

Superconductivity

160

161

however, no need for a


is not in turn used
to drive another unit. One therefore often calls the memory elements
low gain cryotrons.
The Crowe cell basically consists of a thin film of superconducting

consisting of crossed or parallel gate

and control films separated by


and placed between additional superconducting
shielding films called ground planes. The resistance of the thin film
gates is comparable to that of a wire gate, but the ground planes confine magnetic flux to a very small region and thus result in L/R values
8
10
of the order of 10" -10~ sec. Cryogenic loops with a time constant
-9

superconductivity in another. There

insulating layers,

greater-than-unity gain, as the controlled current

of 2x 10
sec have been operated (Ittner, 1960b). An account of
many of the design considerations governing such thin film cryotrons

material (for example, lead) with a small hole, a few millimetres in


diameter, which has a narrow cross-bar running across it. This is
drive 'wire' in the form of a
shown schematically in Figure 51.

can be found

second narrow

in several

papers in

is,

above the cross-bar, separated only by


a thin insulating layer. As long as the entire configuration remains
superconducting, the magnetic flux threading the hole must retain its

[9].

strip lies just

INPUTS

a\or
IB

xr5=-fo

_<ztXp

SENSE
B

-^^Zpr*^

(neither

c\
t

nor B

WIRE.

cnto

QXlo

Fio. 51

Fig. 50

original value,
is

Suggestions for superconducting

memory

devices were advanced

which we

shall take to

passed through the drive wire,

and the remainder of the

film.

be zero. Therefore

from penetrating, and


which shows a
The cross-bar is very thin and

keep the

lating current will be such as to

fact that a current induced in a superconducting ring will persist in-

schematic cross section of the cell.


narrow and therefore has a low critical current.

possibility of a two-state

can circulate either way one has the

memory storing one bit of information with

no dissipation of power other than that required to maintain the low


temperature. Of the three suggestions it is that of Crowe on which in
recent years most attention has been concentrated and which will be
briefly described here. Before doing so it might be noted that persistent current memory devices have in common with switching
cryotrons that a current in one superconducting circuit quenches the

a current

will

simultaneously by Buckingham (1958), Crittenden (1958), and Crowe


(1958). Their devices are basically quite similar and make use of the
definitely. Since the current

if

induce currents in the cross-bar


The direction of this induced circu-

it

flux

results in a flux distribution indicated in Figure 52a,

When

the induced

current exceeds this critical value, the cross-bar becomes normal. The
flux now changes to the configuration shown in Figure 52b, as the

remainder of the film remains superconducting. If finally the drive


current is again removed, the superconductivity of the cross-bar is
restored, and now the flux threading the hole is trapped, as long as the
cross-bar remains superconducting, by a persistent current which
in the opposite direction

of the originally induced flow. Even

is

when

the drive wire current

is

now removed,

the flux distribution remains

that of Figure 52c.

The

163

Superconducting devices

Superconductivity

162

rendered more complicated than

The operation of the Crowe cell is


indicated in the preceding simplified account because the cross-bar
heats up through joule heat when it becomes normal, and the thermal
is

Crowe cell (Garwin, 1957) is indicated


which shows on equal time scales, but arbitrary vertical
scales, the drive current Id and the cross-bar current /c Pulse 1 is
too small to induce a critical value of Ic Pulse 2 results in Ic > Icril
the cross-bar becomes momentarily normal, and after the drive pulse
idealized operation of a

in Figure 53,

may be appreciable. Crowe (1957), Rhoderick (1959),


Von Ballmoos (1961), and several papers in [9] discuss the resulting
recovery time

complications.

STORE
"0"

Id

READ STORE

READ

"1"

_
V

v--

STORED STORED
"0"
"0
-

is

removed a

persistent current l
v

stored. Pulse 3

is
it

is

now

4,

READY

a 'readFig. 53

induces a current in a direction

opposite to that of the persistent current. With pulse

"I"

IV

K READO"

out' pulse which has no effect since

STORED

however, the

state

Crowe cells can be arranged into a two-dimensional matrix of


memory elements with the drive wire forming part both of an x- and

read-out.

a y-circuit, as indicated in Figure 51 Driving pulses Ix Iy are then so


chosen that either alone is not sufficient to activate the device, but

persistent current

is

reversed, storing the other possi bility of the two-

memory, and now read-out pulse 5 succeeds in driving the


cross-bar well beyond the critical value. Note that this is a destructive

The memory
very close to

duced

it

is

sensed by means of a wire below the cross-bar, also

but electrically insulated.

in the sense

wire because of

linking the cross-bar changes, that

is,

A current pulse will

be in-

proximity whenever the flux


whenever the cross-bar becomes

its

normal. Thus we note on Figure 53 that the sense wire response Is to


pulse 3 is nothing, which can be taken as ' Read 0', while its response
to 5

is

a pulse which can be taken as 'Read

'.

The reader is again referred to [9] for a number


memories built up of such matrices.
superconducting
papers
on
of
Rose-Innes (1959) has estimated the consumption of liquid helium
required to keep a memory like that cold, and finds this to be of the
order of two litres per hour for an array of one million cells. This is

that both together do.

as
well within the capacity of closed cycle helium refrigerators such
the one described by McMahon and Gifford (1960).

Bibliography
General References
University Press,
[I] shoenberg, d., Superconductivity, Cambridge
1952.
[2]

London,

f.,

[3]

Progress

in

Superfluids, Vol.

Low

I,

New York,

Temperature Physics, Vol.

Wiley, 1950.
C. J. Goiter, ed.

I,

New York, Interscience,

1955.
Superconductivity: Experimental Part, Handbuch der
Physik, Vol. XV, S. Fliigge, ed.; Berlin, Springer, 1956.
ibid.
[5] bardeen, J., Theory of Superconductivity,
Graham and A. C.
M.
Phys.,
G.
Low
Temp.
VII
Int.
Conf
Proc.
[6]
Hollis-Hallett, eds. ; Toronto, University Press, 1960.
[4] serin, b.,

[7]

bardeen,

j.,

and schrieffer,

Superconductivity, Progress in
Ill,

[8]

C.

J.

j.

r.,

Recent Developments

Low Temperature

in

Physics, Vol.

New York, Interscience, 1961.


Review of the Present Status of the Theory of Super-

Goiter, ed.;

bardeen,

j.,

conductivity, IBM Journal 6, 3 (1962).


Proc. Symp. Super cond. Techniques, Washington, May, 1960;
Solid State Elect r. 1, 255^08 (1960).
Physics, C. de
[10] tinkham, m., Superconductivity, Low Temperature
Witt, B. Dreyfus, and P. G. De Gennes, eds. London, Gordon
[9]

[II]

[12]

and Breach, 1962.


High Magnetic Fields, H. Kolm, B. Lax, F. Bitter, and R. Mills,
eds.; New York, Wiley, 1962.
Proc. VIII Int. Conf. Low Temp. Phys., R. Davics, ed.; London,

Butterworth, 1964.
Conf on the Science of Superconductivity, Colgate,
1963; published in Rev. Mod. Physics, 36, 1-331 (1964).
Lecture
[14] Metaux et Alliages Supraconducteurs, P. G. De Gennes,
Notes, Paris 1962-63.

[13] Proc. Int.

Individual References
e., and weiss, p. R. (1959), Cambridge Conference on
Superconductivity (unpublished).
abrtkosov, a. a. (1957), J.E.T.P. USSR 32, 1442; Soviet Phys.
J.E.T.P. 5, 1 174; /. Phys. Chem. Solids 2, 199.
abrikosov, a. a., and gorkov, l. p. (1960), J.E.T.P. USSR 39, 1781

Abrahams,

Soviet Phys. J.E.T.P. 12, 1243 (1961).


165

ABRIKOSOV,
J.E.T.P.

A. A.,

GORKOV,

USSR 35, 265

L. P.,

and KHALATNIKOV,

Soviet Phys. J.E.T.P. 8,

82

(1

I.

M. (1958),

959) see also


;

khalatnikov, i. m., and abrikosov, a. a. (1959), Advances


Physics, N. F. Mott, ed., London, Taylor and Francis Ltd.

in

adkins, c. J. (1963), Phil. Mag. 8, 1051 ; [13], p. 21 1.


adler, J. g., and Rogers, J. s. (1963), Phys. Rev. Lett. 10, 217.
alekseevskii, n. e. (1953), J.E.T.P. USSR 24, 240.
alers, g. a., and waldorf, d. l. (1961), Phys. Rev. Lett.
Journal 6, 89 (1962).

6,

677;

108, 1175.

rickayzen,

j.,

g.,

and tewordt,

l.

(1959), Phys. Rev. 113,

982.

IBM

p. b. (1957), Phys. Rev. 105, 104.


alers, p. b. (1959), Phys. Rev. 116, 1483.

ambegaokar,

v., and baratoff, a. (1963), Phys. Rev. Lett. 10, 486;


Erratum: ibid., 11, 104.
Anderson, p. w. (1959), /. Phys. Chem. Solids 11, 26.
ANDERSON, P. w. (I960), [6], p. 298.
Anderson, p. w. (1962), Phys. Rev. Lett. 9, 309.
Anderson, p. w. (1963), Ravello Lectures (unpublished).
Anderson, p. w., and KIM, y. b. (1964), [13], p. 39.
Anderson, p. w., and rowell, J. m. (1963), Phys. Rev. Lett. 10, 230.
Anderson, p. w., and suhl, h. (1959), Phys. Rev. 116, 898.
andres, k., olsen, j. l., and rohrer, h. (1962), IBM Journal 6, 84.
ANDREWS, D. H., MILTON, R. M., and DE SORBO, W. (1946), J. Opt. Soc.

36, 518.

androes, g. m., and knight, w. d. (1961), Phys. Rev. 121, 779.


autler, s. h. (1960), Rev. Sci. Inst. 31, 369.
autler, s. h., rosenblum, e. s., and GOOEN, K. h. (1962), Phys. Rev.
Lett. 9, 489; see also [13], p. 77.

bean, c.
bean, c.
9,93.
bean, c.
becker,

261 (1951).

(1962), Phys. Rev. Lett. 8, 250.

p.

DOYLE, m.

p.,

P.,

r.,

v..

and pincus,

a. u. (1962),

Phys. Rev. Lett.

and Livingston, J. d. (1964). Phys. Rev. Lett. 12, 14.


heller, c, and SAUTER, F. (1933), Z. Phys. 85, 772.

berlincourt, t. G. (1963), Brit. J. Appl. Phys. 14, 749.


berlincourt, t. o., and hake, r. r. (1962), Phys. Rev. Lett. 9, 293.
berlincourt, t. g., and hake, r. r. (1963), Phys. Rev. 131, 140.
berlincourt, t. g., hake, r. r., and Leslie, d. h. (1961), Phys. Rev.
Lett. 6, 671.

betbeder-matibet,

o.,

and nozieres,

p.

(1961), C. R. Acad. Sci. 252,

3943.

BEZUGLYl,

P. O.,

USSR 36,
BIONDI, M.

GALK1N,

1951

A.,

A. A.,

and KAROLYUK,

A. P. (1959), J.E.T.P.

Soviet Phys. J.E.T.P. 9, 1388.

FORRESTER, A. T., GARFUNKEL, M.


Mod. Phys. 30, 1 109.

P.,

and SATTERTHWAITE,

C.B. (1958), Rev.

biondi, m.

A.,

BIONDI, M.

A.,

and garfunkel, m. p. (1959), Phys. Rev. 116, 853, 862.


GARFUNKEL, M. P., and MCCOUBREY, A. O. (1957), Phys.

Rev. 108, 495.

BLANC,

J.,

(1960),

balashova, b. m., and sharvin, vu. v. (1941), J.E.T.P. USSR 17, 851
ballmoos, f. von (1961), Thesis, E.T.H., Zurich.
baltensperger, k. (1959), Helv. Phys. Acta 32, 197.
bardasis, a., and schrieffer, j. r. (1961), Phys. Rev. Lett. 7, 79.
bardeen, j. (1950), Phys. Rev. 80, 567; see also Rev. Mod. Phys. 23,

GOODMAN,
[6],

B.

B.,

KUHN,

G.,

LYNTON,

E. A.,

and WEIL,

L.

p. 393.

blumberg, R. H. (1962), J. Appl. Phys. 33, 1822.


BON-MARDION, O., GOODMAN, B. B., and LACAZE,
2,321.
boorse, h. a. (1959), Phys. Rev. Lett. 2, 391.
bozorth, r. m., davis, d. o., and Williams, a.

A. (1962), Phys. Lett.

j.

(I960), Phys. Rev.

119, 1570.

brenig, w. (1961), Phys. Rev. Lett. 7, 337.


a., zemansky, m. w., and boorse, h.

j.

(1952), Phys. Rev. 87, 192.

j.

(1954), Phys. Rev. 94, 554.

j.

(1957),

j.

(1958), Phys. Rev. Lett. 1, 399.

j.

(1959),

Nuovo CimentoS,

brown,
1766.

Cambridge Conference on Superconductivity (un-

published).

bardeen,
bardeen,

bardeen, j. (1962a), Phys. Rev. Lett. 9, 147.


BARDEEN, j. (1962b), Rev. Mod. Phys. 34, 667.
bardeen, j., cooper, l. n., and schrieefer, j. R. (.1957), Phys. Rev.
bardeen,

alers,

bardeen,
bardeen,
bardeen,
bardeen,
bardeen,

167

Bibliography

Superconductivity

166

J.

(1961a), Phys. Rev. Lett. 6, 57.

j.

(1961b), Phys. Rev. Lett. 7, 162.

a. (1953),

Phys. Rev. 92,

52.

bryant, c. a., and keesom, p. H. (I960), Phys. Rev. Lett. 4, 460; Phys.
Rev. 123,491 (1961).
bucher, e., gross, d., and olsen, j. l. (1961), Helv. Phys. Acta 34, 775.
bucher, e., and olsen, j. l. (1964), [12], p. 139.
buck, d. a. (1956), Proc. I.R.E. 44, 482.
12

Super conduc tivity

68

Buckingham,

M.

J.

(1958), Proc.

V Int.

Conf.

Bibliography

Low Temp.

Madi-

Phys.,

son, J. R. Dillinger, ed.: Madison, U. of Wisconsin Press, p. 229.


budnick, J. I. (1960), Phys. Rev. 119, 1578.
budnick, J. I., lynton, e. a., and SERIN, b. (1956), Phys. Rev. 103, 286.
BURSTEIN, E., LANGENBERCi, D. N., and TAYLOR, B. N. (1961), Phys. Rev.
Lett. 6, 92.

byers, n., and yang, c. n. (1961), Phys. Rev. Lett. 7, 46.


calverley, a., Mendelssohn, K., and rowell, p. M. ( 96 ), Cryogenics
1

2,26.

69

Phys. Lett. 9, 106.

calverley, a., and rose-innes, a. c. (1960), Proc. Roy. Soc. A255, 267.
cape, j. a. (1963), Phys. Rev. 132, 1486.
caroli, c, de gennes, p. g., and matricon,

j.

(1962). J. Phys. Rod.

23, 707.

chandrasekhar, b. s. (1962), Phys. Lett. 1, 7.


chang, g.k., jones, r. f.., and toxen, a.m. (1962), IBM Journals, 12.
chang, g. k., kinsel, t., and SERIN, b. (1963), Phys. Lett. 5,
chanin, g., lynton, e. a., and serin, b. (1959), Phys. Rev. 114, 719.
chiou, c, quinn, d., and seraphim, d. (1961), Bull. Am. Phys. Soc. 6,
122; seraphim, D.,cmou,c, and quinn, d., Acta Met. 9, 861 (1961).
I

clogston,
CLOGSTON,

a.

m. (1962), Phys. Rev. Lett. 9, 266.

A. M.,

GOSSARD,

A.

C, JACCARINO,

V.,

and YAFET,

Y. (1962),

Phys. Rev. Lett. 9, 262; see also [13], p. 170.

COCHRAN,

J. F.,

MAPOTHER,

D. E.,

and MOULD,

R.

E.

(1958), Phys. Rev.

103, 1657.

cody, G. d. (1958), Phys. Rev. Ill, 1078.


COFFEY, H. T., FOX, D. K., HULM, J. K., SPAN,

R. E.,

and REYNOLDS, W.

T.

Am.

Phys. Soc. 9, 454.


COHEN, M. H., FALICOV, L. M., and PHILLIPS, j. c. (1962), Phys. Rev. Lett.
8, 316; see also [12], p. 178.
cohen, m. l. (1964), [13]. p. 240; Phys. Rev. 134, A51 1.
collins, s. c. (1959) (unpublished). I am grateful to Prof. Collins for
a detailed description of this experiment.
(1964), Bull.

compton, v. b., and Matthias, b. t. (1959), Acta Cryst. 12, 651.


compton, w. dale (1959), private communication to B. Serin.
conolly, a., and Mendelssohn, k. (1962), Proc. Roy. Soc. A266, 429.
l. n. (1956),
l. n.

Phys. Rev. 104, 1189.

(1959), Phys. Rev. Lett. 3, 17.

Crittenden, jr., e. c. (1 958) Proc. Vint. Conf. Low Temp. Phys., Madison, J. R. Dillinger, ed.; Madison, U. of Wisconsin Press, p. 232.

crowe,
crowe,

j.

w.
w.

(1957),

IBM Journal 1,

295.

Conf. Low Temp. Phys., Madison,


J. R. Dillinger, ed.; Madison, U. of Wisconsin Press, p. 238.
CULLER, G. J., FRIED, B. D., HUFF, R. W., and SCHRrEFFER, J. R. (1962),
Phys. Rev. Lett. 8, 399.

casimir, h. b. g. (1938), Physiea 5, 595.


casimir, h. b. g. (1940), Physiea 7, 887.
chambers, r. g. (1952), Proc. Roy. Soc. A215, 481.
chambers, r. g. (1956), Proc. Canib. Phil. Soc. 52, 363.

cooper,
cooper,

COOPER, L. N. (1960), Am. J. Phys. 28, 91.


cooper, L. N. (1961), Phys. Rev. Lett. 6. 689: IBM Journals, 75 (1962).
cooper, l. n. (1962), Phys. Rev. Lett. 8, 367 see also [12], p. 126.
CORAK, W. S., GOODMAN, B. B., SATTKRTHWAITE, C. B., and WEXLER, A.
(1954), Phys. Rev. 96, 1442; see also Phys, Rev. 102, 656 (1956).
corak, w. s., and satterthvvaite, c. b. (1954), Phys. Rev. 99, 1660.
cornish, f. h. j., and olsen. j. l. (1953), Helv. Phys. Acta 26, 369.
CRIBIER, D., JACROT, B., MADHOV RAO, L., and FARNOUX, B. (1964),

daunt,
DAUNT,

j.

(1958), Proc.

Int.

and Mendelssohn,

j.

g.,

J.

G., MILLER, A. R.,

k. (1946), Proc. Roy. Soc.

PIPPARD, A.

B.,

A185, 225.

and SHOENBERG,

D. (1948),

Phys. Rev. 74, 842.


da vies, E. A. (1960), Proc. Roy. Soc. A255, 407.
deaver, jr., b. s., and fairbank, w. m. (1961), Phys. Rev. Lett. 7, 43.
de gennes, p. g. (1963), Phys. Lett. 5, 22.
de gennes, p. g., and guyon, e. (1962), Phys. Lett. 3, 168.

de gennes,
de gennes,

and matricon, j. (1964), [13], p. 45.


and sarma, g. (1963), /. Appl. Phys. 34,
and shoenberg, d. (1948), Proc. Phys. Soc.

p. g.,

p. g.,

1380.

desirant, m.,
60, 413.
de sorbo, w. (1959), J. Phys. Client. Solids 15, 7.
de sorbo, w. (I960), Phys. Rev. Lett. 4, 406; 6, 369.
de sorbo, w. (1961), IBM Conf. on Superconductivity (unpublished).
DIETRICH, I. (1964), [12], p. 173.
doidge, p. r. (1956), Phil. Trans. Roy. Soc. A248, 553.
doll, r., and nabauer, m. (1961), Phys. Rev. Lett. 7, 51.

Douglass, jr., d. h. (1961a), Phys. Rev. Lett. 6, 346.


Douglass, jr., d. h. (1961b), Phys. Rev. Lett. 7, 14.
Douglass, jr., d. h. (1961c), Phys. Rev. 124, 735.
Douglass, jr., d. h. (1962a), Bull. Am. Phys. Soc. 7, 197.
Douglass, jr., d. h. (1962b), Phys. Rev. Lett. 9, 155.
Douglass, jr., d. h., and blumberg, r. h. (\962). Phys. Rev. 127, 2038.
douglass, jr., d. h., and falicov, l. m. (1 964), Progress in Low Temp.
Phys., Vol. IV, C. J. Gorter, ed.; New York, Interscience.
douglass, jr.. d. H.. and meservey, r. b. (1964a), [12], p. 180.

1
.

Bibliography

Superconductivity

170

Douglass, jr., d. h., and meservey, r. ii. (1964b), Phys. Rev. 135, A 19.
drangeid, k. e., and sommerhalder, r. ( 962), Phys. Rev. Lett. 8, 467.
DRESSELHAUS, M. S., DOUGLASS, JR., D. H., and KYHL, R. L. (1964), [12],
1

L1NDENFELD,

L.,

P.,

LYNTON,

E. A.,

and ROHRER,

H. (1963),

P.,

LYNTON,

E. A.,

and ROHRER,

H. (1964),

Phys. Rev. Lett. 10, 98.

DUBECK,

LINDENFELD,

L.,

[13], p. 110.

ehrenfest,

p.

BUASHBERO,

(1933),

Comm.

Leiden Suppl. 75b.


USSR 38, 996; Soviet Phvs. J. E.T.P.

Q. M. (1960), J. E. T. P.

11, 696.

ERLBACH, E., GARWIN,


Journal*, 107.
faber, t.
faber, T.
faber, t.

faber,
FABER,
faber,

R.

L.,

and SARACIIIK.

e.

(1952), Proc. Roy. Soc.

E.

(1954), Proc. Roy. Soc. A223, 174.

e.

(1955), Proc. Roy. Soc. A23I, 353.

t. e. (1957), Proc.
T. t. (1958), Proc.

M.

P.

(1960), l.B.M.

A214, 392.

Roy. Soc. A241, 531.


Roy. Soc. A248, 460.

communication.
and pippard, a. b. (1955a), Proc. Roy. Soc. A231, 336.
e., and pippard, a. b. (1955b), [3], Chapter IX.
iairbank, w., pierce, J. m., and WILSON, p. b. (1964), [12], p. 324.
t. E. (1961), private

faber, T.
faber, t.

ferrell,
ferrell,
ferrell,
ferrell,

T. H.,

and Matthias,
and MATTHIAS,

T. H.,

MATTHIAS,

t. h.,

E.,

r. a. (1961),

Phys. Rev. Lett. 6, 541.

and glick, a. J. (1962), Dull. Am. Phys. Soc. 7, 63.


r. a., and GLOVER III, R. e. (1948), Phys. Rev. 109, 1398.
r. a., and PRANGE, R. e. (1963), Phys. Rev. Lett. 10, 479.
feucht, d. l., and WOODFORD, JR., J. B. (1961 j, J. Appl. Phys. 32, 1882.
r. a.,

mini, m. (1964), Bull. Am. Phys. Soc. 9, 454.


HNNtMOKF., D. K., and mapoi hik, D. E. (1962), Phvs. Rev.

B. T. (1964), [12], p. 159.

B. T.,

HULL,

W., and CORENZWIT,

JR., G.

E.

giaever, i. (1964), [12], p. 171.


GIAEVER, L, HART, JR., H. R.,andMGERLE, K. (1962),/%.?. Rev. 126,941
giaever, i., and megf.rle, k. (1961), Phys. Rev. 122, 1 101.
Ginsberg, D. M. (1962), Phys. Rev. Lett. 8, 204.
Ginsberg, d. m., and LESLIE, j. d. (1962), IBM Journal 6, 55.
GINSBERG, D. M., RICHARDS, P. L., and TINKHAM, M. (1959), Ph)'S. Rev.
Lett. 3, 337.

Ginsberg, d. m., and tinkham, m. (1960), Phys. Rev. 118, 990.

ginzburg,
ginzburg,

v. l. (1945), /.
v. l. (1955),

Lett. 9, 288.

FORRESTER,

A. T. (1958),

j.,

de gennes,

p. g.,

and MATRICON,

J.

(1963), Appl. Phvs.

irohlich,

305.

USSR29,

748; Soviet Phys. J.E.T.P.

3,

621

v. l. (\956a),J.E.T.P. USSR 30, 593; Soviet Phys. J.E.T.P.


Dokl. Akad. nauk USSR 110, 368; Soviet Phvs. Dokladv 1.

541.
v. l.

(1956b),7..r.P.

USSR 31,

541

Soviet Phys. J.E.T.P.

4, 594.

7,78.
ciiNZBURG,

l.

(l958a),J.E.T.P.

v. l. (1958b),

USSR 34,

Soviet Phvs. J.E.T.P.

Physica 24, S42.

ginzburg, v. l., and landau, l. d. (1950), J.E.T.P. USSR 20, 1064;


sec also ginzburg, v. l., Nuovo Cimento 2, 1234 (1955).
glover in, R. e., and tinkham, m. (1957), Phys. Rev. 108, 243.
GOEDEMOED, S. H., VAN DER GIESSEN, A., DE KLERK, D., and GORTER,

Lett. 2, 119.
H. (1950), Phys. Rev. 79, 845.

garfunkel, m. p., and serin, b. (1952), Phys. Rev. 85, 834.


OAKLAND, JR.. J. w. (1963a), Phys. Rev. Lett. 11, 111.
garland, jr., J. w. (1963b), Phys. Rev. Lett. 11, 114.
garland, jr., j. w.( 1964). [12], p. 143.
garwin, R. l. (1957). I.B.M. Journal 1, 304.
C.AYLEY, R. !., LYNTON, E. A., and SERIN, B. (1962), Phvs. Rev.

USSR 9.

ginzburg,

ginzburg. v.

Phys. Rev. 110, 776.

Phys.

J.E.T.P.

2,589(1956).

ginzburg,

FISKE, m. d. (1964), [13], p. 221.

43.

IBM Journal 6, 256.

(1962),

geilikman, b. t. (1958), J. E.T.P. USSR 34, 1042; Soviet Phys.


J.E.T.P.1,12\.
geilikman, b. t., and kresin, v. z. (1958), Dokl. Akad. nauk USSR
123, 259; Soviet Phys. Doklady 3, 1 161.
geilikman, b. t., and kresin, v. z. (1959), J. E.T.P. USSR 36, 959;
Soviet Phys. J.E.T.P. 9, 677.
geiser, r., and goodman, b. b. (1963), Phys. Lett. 5, 30.
giaever, i. (1960a), Phys. Rev. Lett. 5, 147.
giaever, i. (1960b), Phys. Rev. Lett. 5, 464.

friedf.l,

b. t.

(1961), Phys. Rev. Lett. 6, 275.

p. 328.

DUBECK,

geballe,
GEBALLE,
GEBALLE,

c.

j.

(1963), Phys. Lett. 3, 250.

goodman, b. b. (1953), Proc. Phys. Soc. A66, 217.


goodman, b. b. (1957), C. R. Acad. Sci. 224, 2899; see also
126,

Sci. 246. 3031 (1958).

goodman,

u. b.

(1958), C. R. Acad. Sci. 246. 3031

C. R. Acad.

Bibliography

Superconductivity

172

goodman,

(1961), Phys. Rev. Lett. 6, 597;

b. b.

IBM

Journal

6,

63.

goodman,
goodman,
goodman,
GOODMAN,
GOODMAN,

IBM Journal 6,

62.

B. B.

(1962a),

b. b.

(1962b), Phys. Lett. 1, 215.


(1962c), J. Phys. Rod. 23, 704.

b. b.

HILLAIRET,

J.,

VEYSS1E,

J. J.,

and WEIL,

L.

(1960),

[6],

p. 350.

gorkov,

L. P.

(\958), J.E.T.P.

l. p.

(1959), J.E.T.P.

USSR 34,

735; Soviet Phys. J.E.T.P.

7,

505.

2,

hebel, l. c. (1959), Phys. Rev. 116, 79.


hebel, L. c, and slichter, c. p. (1959), Phys. Rev. 1 13, 1504.
heer, c. v., and daunt, j. g. (1949), Phys. Rev. 76, 854.
HEIN, R. A., FALGE JR., R. L., MATTHIAS, B. T., and CORENZW1T, E.
(1959),
Phys. Rev. Lett. 2, 500.
hein, r. a., and gibson, j. w. (1963), Phys. Rev. 131, 105.
HEIN, R. A., GIBSON, J. W., MATTHIAS, B. T., GEBALLE, T. H., and CORENzwit, E. (1962), Phys. Rev. Lett. 8, 313.
HEIN, R. A., GIBSON, J. W., MAZELSKY, R., MILLER, R. C, and HULM,
K.
1

gorkov,

USSR 36,

1918; 37, 833, 1407; Soviet

Phys. J.E.T.P. 9, 1364; 10, 593, 998 (1960).


l. p. (1960), [6], p. 315.

gorkov,
gorter,
gorter,
gorter,
gorter,
GORTER,
gorter,
gorter,

Mus. Teyler

.1.

(1964), Phys. Rev. Lett. 12, 320.


HEIN, R. A., GIBSON, J. W., PABLO, M. R.,

c.

J.

(1933), Arch.

c.

J.

(1935), Physica 2, 449.

c.

j.

(1962a), Phys. Lett. 1, 69.

c.

j.

(1962b), Phys. Lett. 2, 26.

c.

j.

(1964), [13], p. 27.

Hempstead, c. f., and KIM, y. b. (1963), Phvs. Rev.


herring, c (1958), Physica 24, SI 84.

c.

J.,

c.

j.,

and casimir, h. b. g. (1934a), Physica 1, 306.


and casimir, h. b. g. (1934b), Phys. Z. 35, 963 Z.

hilsch, p., and hilsch, r. (1961), Naturwiss. 48, 549.


HIRSHFELD, A. T., LEUPOLD, H. A., and BOORSE, II. A. (1962), Phys. Rev.

7, 378.

grayson smith,

ii.,

mann,

Roy. Soc. Can. 30,

grayson smith,

h.,

K.

techn.

c, and wilhelm,

j.

o. (1936), Trans.

g. a. (1935), Trans. Roy. Soc. Can. 29.

guenault, a. m. (1960), [6], p. 409; Proc. Roy. Soc. A262, 420 (1961).
guyon, e., caroli, C, and martinet, a. (1964), J. Physique 25, 661.
haas, w. j. de, and voogd, j. (1931), Comm. Leiden 214c.
haenssler, f., and rinderer, l. (1960), [6], p. 375.
hake, r. R., and bremmer, w. g. (1964), Phys. Rev. 133, A 179; sec
r., [13], p.

R. R., LESLIE, D. H.,

124.

and BERLINCOURT,

T. G. (1962), Ph)'S.

Rev.

127, 170.

HAKE, R.

R.,

MAPOTHER,

D. E.,

and DECKER,

D. L. (1958), P/lVS. Rev.

1522.

HAMMOND,
harden,
hauser,
hauser,

and KELLY, G. M. (1964), [13], p.


and arp, v. (1963), Cryogenics 4,

R. H.,

j. l.,
j. J.
j. J.,

185.
105.

(1962), Phys. Rev. Lett. 9, 423.

and buehler,

e.

(1963),

Lett. 12, 145.

p.

(1963), J.E.T.P.

USSR 45,

1208; Soviet Phvs. J.E.T.P.

18,834(1964).

23.

also hake, r.

R. D.

127, 1501.

hohenberg,

13.

andTARR, F.

and BLAUGHER,

Phys. Rev. 129, 136.

Phys. 15, 539.


graham, g. m. (1958), Proc. Roy. Soc. A248, 522.
grassmann, p. (1936), Phys. Z. 37, 569.

HAKE,

73

196.

B. B. (1964), [13], p. 12.


B. B.,

HAUSER, J. J., and iiiu and, e. (1962), Phys. Rev. 127, 386.
HAYNES, M. K. (1960), [9], p. 399.
heaton, j. w., and rose-innes, a. c. (1963), Appl. Phys. Lett.

(1961), Phys. Rev. 125. 142.

1 18,

hulm,

j.

k. (1950), Proc. Roy. Soc.

ittner m, w.
ittner in, w.

A204, 98.
(1960a), Phys. Rev. 119, 1591.
b. (1960b), Solid State Jr., July/Aug.

b.

and sommerhalder, r. (1959), Helv. Phys. Acta 32, 313.


and sommerhalder, r. (I960), Helv. Phys. Acta 33, 1.
joiner, w., and serin, b. (1961), private communication; see also
jaggi,

r.,

jXggi,

r.,

serin (1960).
jones, R. e., and toxen, a. m. (1960), Phys. Rev. 120,
josephson, b. d. (1962), Phys. Lett. 1, 251.

josephson,

b. d.

kadanoff,

l. p.,

167.

(1964), [13], p. 216.

and martin, p. c. (1961), Phys. Rev. 124, 670.


kamerlingh onnes, h. (191 1), Leiden Comm. 122b, 124c.
kamerlingh onnes, h. (1913), Leiden Comm. Suppl. 34.
kamerlingh onnes, h., and tuyn, w. (1924), Leiden Comm. Suppl.
50a; see also tuyn, w., Leiden

KAPLAN,

R.,

NETHERCOT,

A. H.,

Comm.

198 (1929).

and BOORSE,

H. A. (1959), P/lVS.

116, 270.

keesom, w. h. (1924), 4C Congr. Phys. Solvay,

p. 288.

Rev

Bibliography

Superconductivity

174

keesom, w.

and kamerlingh onnes,

ii.,

h. (1924),

Comm.

l.chlcn

174b.

keller, J. b., and zumino, b. (1961), Phys. Rev. Lett. 7, 164.


khaikin, M.S. (1958), J.E.T.P. USSR34, 1389; Soviet Phys. J.E.T.P.

York
6,

(1952).

lazarev,

and sudovstov,

b. g.,

a.

i.

(1949), Dokl.

Adak. nauk

USSR

69, 345.

735.

khalatnikov,

I.

m. (1959), J.E.T.P.

USSR

36, 1818; Soviet Phys.

kikoin,

i.

k.,

and goobar,

also broer, l.

j. f..

s.

v. (1940), J. Phys. U.S.S.R. 3, 333; see

Physica 13, 473 (1947).

t.,

131, 2486.

HEMPSTEAD,
kleiner, w. h., roth,
B.,

and STRNAD, A. R. (1964), [13], p. 43.


m., and AUTi.ER, s. h. (1964), Phys. Rev.

m.,

e.

133,

A 1226.
kleinman, l. (1963), Phys. Rev. 132, 2484.
klemens, P. g. (1956), Chapter IV, Vol. XIV, Hanclbuch der Physik,

lindenfeld,
little,

w.

Flugge, ed.; Springer Vcrlag, Berlin.


kok, J. a. (1934), Physica 1, 11 03.
kondo, j. (1963), Prog. Theor. Phys. 29, 1.
kresin, v. z. (1959), J.E.T.P. USSR 36, 1947; Soviet Phys. J.E.T.P.

parks, r.
Livingston,

9,

1385.
1

(1961a), J. Appi. Phys. 32, 325.

KUNZLER,

J.

E.,

E.,

F. S. L.,

and WERNICK,

J. E.

(1961b),

Phys. Rev. Lett. 6, 89.


kuper, C. g. (1951), Phil. Mag. 42, 961.
landau, l. d. (1937), J.E.T.P. USSR 7, 371 ; Phys. Z. Sowjet. 11, 129.
landau, l. d. (1943), J.E.T.P. USSR 13, 377.
landau, l. d., and lifshitz, e. m. (1958), Statistical Physics, pp.

434 ff. ; London, Pergamon Press.


laredo, s. j., and pcppard, a. b. (1955), Proc. Cambridge

Phil. Soc.

51, 369.

laurmann,

e.,

and shoenberg,

d. (1949), Proc.

Roy. Soc. A198,

560.

laurmann,

e.,

and shoenberg,

d. (1947),

K.,

Nature 160, 747.

p. (1961),

Phys. Rev. Lett. 6, 613.

and parks, r. d. (1962), Phys. Rev. Lett. 9, 9; see also


d., and little, w. a., Phys. Rev. 133, A97 (1964).
J.

b. (1963),

Phys. Rev. 129, 1943; see also/. Appl. Phys.

LIVINGSTON,

lock,
lock,
lock,
lock,

j.

J. b. (1964), [13], p. 54.


m. (1951), Proc. Roy. Soc. A208, 391.
m. (1961a), Cryogenics 1, 243.

j.

m. (1961b), Cryogenics 2, 65.

J.

j.

m.,

Phil. Soc.

kropshot, R. h., and arp, v. d. (1961), Cryogenics 2,


kunzler, j. e. (1961a), Rev. Mod. Phys. 33, 501.
kunzler, j. e. (1961b), Conf. High Magn. Fields, Cambridge, Moss.
(to be published).
KUNZLER, J. E., BUEHLER, E., HSU, F. S. L., MATTHIAS, B. T., and WAHL, C.
HSU,

GINSBERG, D. M., FDMNEMORE, D.

a.,

S.

BUEHLER,

L.,

and beaudry, b. j. (1964), Phys. Rev. 134, A309.


and sharvin, yu. v. (1951), Dokl. Akad. nauk USSR

h.,

f.

34, 3028.

C. F.,
l.

lifshitz,

CAPELLETTI, R.

D.,

J.

79, 783.

lynton, e. A., and serin, b. (1962), Phys. Lett. 3, 30.


KINSEL, T., LYNTON, E. A., and SERIN, B. (1964), [13], p. 105.
KIM, Y. B., HEMPSTEAD, C. F., and STRNAD, A. R. (1963), Phys. Rev.
kinsel,

LESLIE,

spedding,

J.E.T.P. 9, 1296.

KIM, Y.

175

laue, m. von (1949), Theorie der Supraleitung, 2nd ed., Berlin,


Springer Verlag; Eng. transl. by L. Meyer and W. Band, New

London,
London,
London,
London,
London,

f.

f.,
f.,

pippard, a. b., and shoenberg, d. (1951), Proc. Camb.


47,811.
(1936), Physica 3, 450.

and London,
and London,

h. (1935a), Proc. Roy. Soc. A149, 71.


h. (1935b), Physica 2, 341.

H. (1935), Proc. Roy. Soc. A152, 650.


h. (1940), Proc. Roy. Soc. A176, 522.
lutes, o. s. (1957), Phys. Rev. 105, 1451.

lynton,
lynton,
lynton,

and mclachlan, d. (1962), Phys. Rev. 126, 40.


and serin, b. (1958), Phys. Rev. 112, 70.
a., serin, b., and zucker, m. (1959), /. Phys. Chem.

e. a.,

e. a.,
e.

Solids 3, 165.

MCLEAN, w. l. (1960), [6], p. 330.


mclean, w. l., and maxfield, b. w. (1 964), Bull. Am. Phys. Soc. 9, 454.
mcmahon, h. o., and gifford, w. e. (1960), [9], p. 273.
maki, k. (1963), Prog. Theor. Phys. 29, 603.
maki, k., and tsuneto, t. (1962), Prog. Theor. Phys. 28, 163.
mapother, d. e. (1959), Conf. on Superconductivity, Cambridge (unpublished).

mapother,

d.

e.

(1962),

IBM Journal 6,

77; see also Phys. Rev. 126,

2021.

markowitz, d., and kadanoff, l. p. (1963), Phys. Rev.


masuda, y. (1962a), IBM Journal 6, 24.
13

131, 563.

Superconductivity

176

masuda,
masuda,

Bibliography

Phys. Rev. 126, 127.


y., and redfield, a. g. (1960a), Bull. Am. Phys. Soc. 5, 176;
Phys. Rev. 125, 159(1962).
masuda, y., and redfield, a. g. (1960b), [6], p. 412; see also masuda,
y., Phys. Rev. 126, 1271 (1962).
matricon, J. (1964), Phys. Lett. 9, 289.
Matthias, b. t. (1957), Chapter V, Progress Low Temperature Physics,
Vol. II, C. J. Gorter, ed.; New York, Interscience.
Matthias, b. (1960), J. Appl. Phys. 31, 23S.
Matthias, b. t. (1961), Rev. Mod. Phys. 33, 499.
Matthias, B. t. (1962), IBM Journal 6, 250.
Matthias, B. t., and CORENZWIT, E. (1955), Phys. Rev. 100, 626.
MATTHIAS, B. T., CORENZWIT, E., and ZACHARIASEN, W. H. (1958a),
Phys. Rev. 112, 89.
MATTHIAS, B. T., COMPTON, V. B., SUHL, H., and CORENZWIT, E. (1959b),
Phys. Rev. 115, 1597.
MATTHIAS, B. T., GEBALLE, T. H., GELLER, S., and CORENZWIT, E. (1954),
Phys. Rev. 95, 1435.
MATTHIAS, B. T., GEBALLE, T. H., COMPTON, V. B., CORENZWIT, E., and
hull, JR., G. w. (1962), Phys. Rev. 128, 588.
y. (1962b),

MATTHIAS, B. T., GEBALLE, T. H., CORENZWIT, E., and HULL, JR., G. W.


(1963), Phys. Rev. 129, 1025.
MATTHIAS, B. T., PETER, M., WILLIAMS, H. J., CLOGSTON, A. M., CORENZWIT, E., and sherwood, r. c. (1960), Phys. Rev. Lett. 5, 542; clogston, a. m., et al., Phys. Rev. 125, 541 (1962).
Matthias, b. t., suhl, h., and corenzwit, e. (1958b), Phys. Rev.
Lett. 1, 92.

Matthias,
1,449.
Matthias,

suhl,

h.,

b. t.,

suhl,

h.,

and corenzwit,
and corenzwit,

e.

e.

(1958c), Phys. Rev. Lett.

(1959a), /. Phys.

Chem.

mattis, d. c, and bardeen, j. (1958), Phys. Rev. Ill, 412.


maxwell, e. (1950), Phys. Rev. 78, 477.
e.
e.
e.

k. (1962),

IBM Journal 6, 27.

and moore, J. r. (1934), Nature 133, 413.


k. and olsen, j. l. (1950), Proc. Phys. Soc. A63, 2.
k., and pontius, r. b. (1937), Phil. Mag. 24, 777.
k., and renton, c. a. (1955), Proc. Roy. Soc. A230,
k.,

157.

mendelssohn,

k.,

and shiffman,

c. a. (1959), Proc.

Roy. Soc. A255,

199.

and crane, l. t. (1963), Phys. Rev. Lett. 11, 107.


and Douglass, jr., d. h. (1964), Phys. Rev. 135, A24.
meshkovsky, a., and shalnikov, a. (1947), J.E.T.P. USSR 17, 851
mercereau,

j. e.,

meservey,

r.,

J. Phys.
miller, p.

b.

USSR

11,1.

(1960), Phys. Rev. 118, 928.

misener, a. d., and wilhelm, j. o. (1935), Trans. Roy. Soc. Can.


29,5.
morel, P. (1959), /. Phys. Chem. Solids 10, 277.
morel, p., and Anderson, p. w. (1962), Phys. Rev. 125, 1263.
MORIN, F. J., MAITA, J. P., WILLIAMS, H. J., SHERWOOD, R., WERNICK, J. H.,
and kunzler, j. e. (1962), Phys. Rev. Lett. 8, 275.
morris, d. e., and tinkham, m. (1961), Phys. Rev. Lett. 6, 600; see also
Phys. Rev. 134, Al 154 (1964).
morse, r. w., and bohm, h. v. (1957), Phys. Rev. 108, 1094.
morse, r. w., olsen, t., and gavenda, j. d. (1959), Phys. Rev. Lett.
15; 4, 193; see also morse, r. w., IBM Journal 6, 58 (1962).
muhlschlegel, b. (1959), Z. Phys. 155, 313.
muller, j. (1959), Helv. Phys. Acta 32, 141.

3,

nambu,

b. t.,

Solids 13, 156.

maxwell,
maxwell,
maxwell,
maxwell,

Mendelssohn,
Mendelssohn,
mendelssohn,
mendelssohn,
mendelssohn,

177

(1952a), Phys. Rev. 86, 235.


(1952b), Phys. Today 5, No. 12, p. 14.
(1960), Adv. Cryo. Eng. 6, 154.

e., and strongin, m. (1964), Rev. Mod. Phys. 36, 144.


meissner, h. (1960), Phys. Rev. 117, 672.
meissner, w., and ochsenfeld, r. (1933), Naturwiss. 21, 787.

Mendelssohn,
Mendelssohn,

k. (1935), Proc. Roy. Soc.


k. (1955),

[3],

Chapter X.

A152, 34.

y., and tuan, s. f. (1963), Phys. Rev. Lett. 11, 1 19; see also
Phys. Rev. 133, Al ( 1964).
NETHERCOT, a. h. (1961), Phys. Rev. Lett. 7, 226.
nicol, j., shapiro, s., and smith, p. h. (1960), Phys. Rev. Lett. 5,
461.

and knight, w. d. (1964), [13], p. 177.


(1958), Rev. Sci. Inst. 29, 537; see also purcell,
payne, e. g. (1960), Adv. Cryo. Eng. 6, 149.

noer, r.
olsen, j.

j.,

l.

j.

r.,

and

olsen, j. l. (1963), Cryogenics 2, 356.


OLSEN, J. L., BUCHER, E., LEVY, M., MULLER, J., CORENZWIT, E., and
GEBALLE, T. (1964), [13], p. 168.
olsen, j. l., and rohrer, h. (1957), Helv. Phys. Acta 30, 49.
olsen, j. l., and rohrer, h. (1960), Helv. Phys. Acta 33, 872; see also

andres,

k.,

olsen,

O'NEAL, H.

R.,

SENOZAN, N.

j. l.,

and rohrer, h., IBM Journal 6, 84


M., and PHILLIPS, N. E. (1964), [12],

(1962).
p. 403.

Bibliography

Superconductivity

178

onsager, l. (1961), Phys. Rev. Lett. 7, 50.


Parkinson, d. h. (1962), Brit. J. Appl. Phys. 13, 49.
Parkinson, d. h. (1964), Brit. J. Appl. Phys. 41, 68.
parks, r. d., and mochel, j. m. (1963), Phys. Rev. Lett. 11, 354; see
also [13], p. 284.
r. h. (1960a), Phys. Rev. 118, 1174.

parmenter,
PARMENTER,

R. H. (1960b), [6], p. 317.

peierls, r. (1936), Proc. Roy. Soc. A155, 613.

peter, m. (1958), Phys. Rev. 109, 1857.


Phillips, J. c. (1963), Phys. Rev. Lett. 10, 96.
Phillips, n. e. (1958), Phys. Rev. Lett. 1, 363.
Phillips, n.

e.

Phillips, n.

e.,

(1959), Phys. Rev. 114, 676.

and Matthias, b. t. (1960), Phys. Rev. 121, 105.


l., and jekula, s. t., Phys. Rev. Lett. 9, 254.

PICKLESIMER, m.

A. B. (1947a), Proc.
a. b. (1947b), Proc.
a. B. (1948),

Roy. Soc. A191, 385.


Roy. Soc. A191, 399.

Nature 162, 68.


Roy. Soc. A203, 210.

a. b. (1950), Proc.

a. b. (1951), Proc.

Camb.

Phil. Soc. 47, 617.

Roy. Soc. A216, 547.


A. b. (1 954), Chapter I, Advances in Electronics and Electron
Physics, L. Marton, ed. New York, Academic Press.
pippard, a. b. (1955), Phil. Trans. Roy. Soc. A248, 97.
pippard, a. b. (1960), [6], p. 320.
pippard, A. B.,andPULLAN,G.T. (1952), Proc. Camb. Phil. Soc. 48, 188.
a. B. (1953), Proc.

POKROVSKH,

v. l.

1961 ),J.E. T.P.

USSR40, 641

Soviet Phys. J.E.T.P.

13, 447.

pokrovskii, v. l., and ryvkin, m. s. (1962), J.E.T.P. USSR 43, 92;


Soviet Phys. J.E.T.P. 16, 67 (1963).
quinn in, d. j., and ittner m, w. b. (1962), J. Appl. Phys. 33, 748.
rayfield, g. s., and reif, f. (1963), Phys. Rev. Lett. 11, 305.
redfield, a. g. (1959), Phys. Rev. Lett. 3, 85; see also redfield, a. g.,
and anderson, a. g. (1959), Phys. Rev. 116, 583.
reese, w., and steyert, jr., w. a. (1962), Rev. Sci. Inst. 33, 43.
reif, f. (1957), Phys. Rev. 106, 208.
reif, f., and woolf, m. a. (1962), Phys. Rev. Lett. 9, 31 5.
reuter, g. e. h., and sondheimer, e. h. (1948), Proc. Roy. Soc. A195,
336.

Reynolds,
691.

c. a., serin, b.,

and nesbitt,

C. A., SERIN, B.,

WRIGHT, W.

II.,

179

and NESBITT,

L. B.

(1950),

Phys. Rev. 78, 487.

rhoderick, e. h. (1959), Brit. J. Appl. Phys. 10, 193.


Richards, p. l. (1960), [6], p. 333; see also Phys. Rev. 126,912(1962).
Richards, p. l. (1961), Phys. Rev. Lett. 7, 412.
Richards, p. L., and TTNKHAM, m. (1960), Phys. Rev. 119, 575.
Roberts, B. w. (1963), Superconducting Materials and Some of Their
Properties, General Electric Report, No. 63-RL-3252M.
rose-innes, a. c. (1959), Brit. J. Appl. Phys. 10, 452.
rose-innes, a. c, and serin, b. (1961), Phys. Rev. Lett. 7, 278.
rosenblum, b., and cardona, m. (1964), Phys. Lett. 9, 220.
rothwarf, a., and cohen, m. (1963), Phys. Rev. 130, 1401.
rowell, j. m. (1963), Phys. Rev. Lett. 11, 200.

rowell,

j.

m.,

anderson,

p.

w., and thomas, d.

e.

(1963), Phys. Rev.

Lett. 10, 334.

pines, d. (1958), Phys. Rev. 109, 280.

pippard,
pippard,
pippard,
pippard,
pippard,
pippard,
pippard,

REYNOLDS,

l. b.,

(1951), Phys. Rev. 84,

rowell,

j.

m.,

chynoweth, a. g., and phillips, t.

c. (1962),

Phys. Rev.

Lett. 9, 59.

ruefenacht, j., and rinderer, l. (1964), [12], p. 326.


saint james, d., and de gennes, p. g. (1963), Phys. Lett. 7, 306.
satterthwaite,
b. (1960), [6], p. 405; Phys. Rev. 125, 873

(1962).

scalapino, d.

and anderson,

p. w. (1964), Phys. Rev. 133, A921.


Phys. Rev. 101, 573.
a. l. (1958), Phys. Rev. 109, 1856.
a. l., and devlin, g. e. (1959), Phys. Rev. 113, 120.
A. L., MATTHIAS, B. T., LEWIS, H. W., and DEVUN, G. E.
(1954), Phys. Rev. 95, 1344.
schooley, j. f., hosler, w. r., and cohen, m. l. (1964), Phys. Rev. Lett.

schawlow,
schawlow,
schawlow,
SCHAWLOW,

j.,

a. l. (1956),

12, 474.
schrieffer,

j. r. (1957), Phys. Rev. 106, 47.


schrieffer, j. r. (1961), IBM Conf. on Superconductivity (unpublished).
schrieffer, j. r., and Ginsberg, d. m. (1962), Phys. Rev. Lett. 8,

207.

schrieffer,

j. r., scalapino, d. j., and wilkins, j. w. (1963), Phys.


Rev. Lett. 10, 336.
schrieffer, j. r., and wilkins, j. w. (1963), Phys. Rev. Lett. 10, 17.

serin, b. (1955),
serin, b. (1960),
serin,
162.

b.,

[3],

Chapter VII.

[6], p.

Reynolds,

391.

c. a.,

and lohman,

c. (1952), Phys. Rev. 86,

Superconductivity

180

shalnikov,

a.

I.,

and SHarvin, yu.

Bibliography

v. (1948), Izv.

Akad. nauk

USSR

12, 195.

shapiro,
shapiro,

(1963), Phys. Rev. Lett. 11, 80; see also [13], p. 223.

s.
s.,

shapoval,

and janus,

e. A.

a. r. (1964), [12], p. 321.

(1961), J.E.T.P.

14, 628 (1962).


sharvin, yu. v. (1 960), J.E.T.P.

USSR 41,

877; Soviet Phys. J.E.T.P.

USSR 38, 298

Soviet Phys. J.E.T.P.

11, 216.

shaw,
shaw,

r. w.,
r. w.,

and mapother, d. e. (1960), Phys. Rev. 118, 1474.


mapother, d. e., and hopkins, d. c. (1961), Phys. Rev.

121, 86.

sherrill, m. d., and Edwards, H. h. (1961), Phys. Rev. Lett. 6, 460.


SHIFFMAN, C. A. (1960), [6], p. 373.
shiffman, c. A. (1961), IBM Conf. on Superconductivity (unpublished).
shoenberg, d. (1940), Proc. Roy. Soc. A175, 49.
SHUBNIKOV, L. W., KOTKEVICH, W. I., SHEPELEV, J. D., and RIABININ,
J. N. (1937), J.E.T.P. USSR 7, 221.
silin, v. p. (1951), J.E.T.P. USSR 21, 1330.

Wash. Acad. Sci. 6, 597.


simmons, w. a., and douglass, jr., d.h.( 1962), Phys. Rev. Lett. 9, 153.
skalski, s., betbeder-matibet, o., and Weiss, p. r. (1964), Bull. Am.
silsbee,

f.

b. (1916), /.

Phys. Soc. 9, 30.


smith, p. h., shapiro,

s.,

miles,

J. l.,

and nicol,

J.

(1961), Phys. Rev.

Lett. 6, 686.

SPIEWAK, m. (1959), Phys. Rev. 113, 1479.


stout, j. w., and guttman, l. (1952), Phys. Rev. 88, 703.
strassler, s., and wyder, p. (1963), Phys. Rev. Lett. 10, 225.
stromberg, t. f., and swenson, c. a. (1962), Phys. Rev. Lett. 9, 370.
STRONGIN, M., PASKIN, A., SCHWEITZER, D. G., KAMMERER, O. F., and
craig, p. P. (1964), Phys. Rev. Lett. 12, 442.
suhl, h. (1962), Low Temperature Physics, C. De Witt, B. Dreyfus,
and P. G. De Gennes, eds. London, Gordon and Breach.
suhl, h., and Matthias, b. t. (1959), Phys. Rev. 114, 977.
suhl, h., Matthias, b. t., and corenzwit, e. (1959), /. Phys. Chem.
;

Solids 11, 347.

swartz, P. s. (1962), Phys. Rev. Lett. 9, 448.


swenson, c. a. (1960), Solid State Physics, Vol. 1 1, p. 41, F. Seitz and
D. Tumbull, eds.; New York, Academic Press.
swenson, c. a. ( 1 962), IBMJournal 6, 82 see also hinrichs, c. h., and
swenson, c. a. (1961), Phys. Rev. 123, 1 106; and schirber, j. e., and
swenson, c. a. (1961), Phys. Rev. 123, 1115.
;

181

swihart, j. c. (1959), Phys. Rev. 116, 45.


swthart, J. c. (1962), IBM Journal 6, 14.
swihart, j. c. (1963), Phys. Rev. 131, 73.
taylor, b. n., and burstein, e. (1963), Phys, Rev.
templeton, i. m. (1955a),/. Sci. Inst. 32, 172.
templeton, i. m. (1955b),/. Sci. Inst. 32, 314.

tewordt,
tewordt,

l.

(1962), Phys. Rev. 128, 12.

l.

(1963), Phys. Rev. 132, 595.

thouless, d.

j.

Lett. 10, 14.

(1960), Phys. Rev. 117, 1256.

tinkham, M. h. (1958), Phys. Rev. 110, 26.


ttnkham, m. h. (1962), IBM Journal 6, 49.
tinkham, m. h. (1963), Phys. Rev. 129, 2413.
tinkham, M. h. (1964), [13], p. 268.
tinkham, m. h., and ferrell, r. a. (1959), Phys. Rev. Lett. 2, 331.
tolmachev, v. v. (1958): see bogoliubov, n. n., tolmachev, v. v.,
and shirkov, d. v., A new Method inthe Theory ofSuperconductivity,

USSR Press, Moscow; Translation: ConNew York, 959) see also tolmachev, v. v.,
USSR 140, 563 (1961); Soviet Phys. Doklady 6,

Section 6.3 (Acad. Sci.


sultants Bureau, Inc.,

Dokl. Akad. nauk

800 (1962).
tomash, w. J., and Joseph, a. s. (1963), Phys. Rev. Lett. 12, 148.
townsend, p., and sutton, j. (1962), Phys. Rev. 128, 591.
toxen, a. m. (1962), Phys. Rev. 127, 382.
toxen, a. m., chang, G. K., and JONES, r. e. (1962), Phys. Rev. 126,
919.

tsuneto,
tsuneto,

Phys. Rev. 118, 1029.


Progr. Theor. Phys. 28, 857.
vroomen, a. r. de (1955), Conf. Phys. Basses Temp., Paris, p.580.
vroomen, a. r. de, and baarle, c. van (1 957), Physica 23, 785.
waldram, j. (1961), thesis, Cambridge University, private comt. (1960),

t. (1962),

munication.

watson, J. h. p., and graham, g. m. (1963), Can. J. Phys. 41, 1738.


werthamer, n. r. (1963a), Phys. Rev. 132, 663.
werthamer, n. r. (1963b), Phys. Rev. 132, 2440.
whitehead, c. s. (1956), Proc. Roy. Soc. A238, 175.
wipf, s. (1961), thesis, University of London; see also coles, b. r.,

IBM Journal 6, 6$ (1962).


wipf, s., and coles, b. r. (1959), Cambridge Superconductivity Conference (unpublished); see also coles, b. r., IBM Journal 6, 68
(1962).

yaqub, m. (1960), Cryogenics

1, 101, 166.

82

Superconductivity

yntema, g. b. (1955), Phys. Rev. 98, 1 197.


young, d. r. (\959), Progr. Cryogenics, Vol. I, p. 1, K. Mendelssohn,
ed.; London, Heywood & Co.
zavaritskii, n. v. (1951), Dokl. Akad. nauk USSR 78, 665.
zavaritskh, n. v. (1952), Dokl. Akad. nauk USSR 85, 749.
zavaritskii, n. v. (1958a), J.E.T.P.
J.E.T.P. 6, 837.
zavaritskii, n. v. (1958b), J.E.T.P.
J.E.T.P. 7, 773.

USSR

33, 1805; Soviet Phys.

USSR

34, 1116; Soviet Phys.

zavaritskii, n. v. (1959), J.E.T.P.


J.E.T.P. 10, 1069.

USSR

37,

1506; Soviet Phvs.

J.E.T.P.

USSR

39, 1193; Soviet Phys.

J.E.T.P. 12, 831 (1961).


zavaritskii, n. v. (1960b), J.E.T.P.
J.E.T.P. 12, 1093 (1961).

USSR

39, 1571; Soviet Phys.

USSR

41, 657; Soviet Phys.

zavaritskh, n.

v. (1960a),

zavaritskii, n. v. (1961), J.E.T.P.


J.E.T.P. 14, 470 (1962).

Index
Adiabatic Magnetization, 15
Anisotropy of energy gap

Bardeen-Cooper-Schrieffer (BCS)
theory - cont.
high frequency conductivity, 103,
114
Knight shift, 139^10
nuclear relaxation rate, 104-6, 137
penetration depth, 37, 52, 130
range of coherence, 48, 115
similarity principle, 122, 130
specific heat, 97, 130, 134-6
thermal conductivity, 91, 92, 136
thermal properties, 127-36
weak coupling limit, 131

BCS theory, 131


decrease with impurity, 97, 100,
and

106, 142
deduction from
infrared absorption, 100
nuclear spin relaxation, 106

96
thermal conductivity, 92
ultrasonic attenuation (table),
138

specific heat,

Anomalous skin effect, 44,


Anomaly in lattice specific
Atomic mass,

effect

101

103

heat, 9

Bulk modulus, 16

Tc 6

on

{See also isotope effect)

Atomic volume,

effect

on

Tc

Coefficient of thermal expansion, 17

Coherence see range of coherence


Coherence effects, 136-8
Cold work, 143
:

Bardeen-Cooper-Schrieffer (BCS)
theory, 12, 117-40, 150
(See also energy gap, interaction parameter
V,
quasi

Condensation energy,

particles)

and G-L theory,

52, 139

effect,

Condensation of electrons

mentum

103

films,

mo-

pairs, 12, 32, 52, 118-21,


125, 140, 144, 151
Critical current, 4, 80, 155, 161
Critical magnetic field, 4

effects,

current and field in thin

80

in

130
critical temperature, 129, 143
electromagnetic properties, 103,
critical field,

in

BCS theory, 130


G-L theory, 51

of lead and mercury, 132


of small specimens, 46, 58, 75-8,
80, 109, 155
of
superconducting
elements

114, 139-40

electron -phonon

in

space, 12, 14, 19, 21, 31,

Cooper

136-8
collective excitations, 100
critical

85, 91,

41

basic hypothesis, 120

coherence

19,

117, 126

Pippard non-local relations,


44

anomalous skin

Collective excitations, 100


Colloidal particles, 29, 35, 46
Compressibility, 17

interaction,

118-20, 125
ground state energy, 120-6

(table), 5

of thin
183

films, 46, 76-7, 80, 109

Index

184

magnetic field - cont.


precise measurements, 83-5
pressure effects, 1617
relation of thin film value to A6

Critical

and

lo,

relation

to

thermal

properties,

135
temperature dependence, 4, 18,
84-6, 130, 135
very high values, 27, 70-1, 109,
4, 85,

155
Critical field for supercooling, 66,

74,76
for surface superconductivity, 74
Critical temperature, 3

{See also isotope effect)

dependence on
atomic mass, 6
atomic volume, 6
discontinuity of specific heat,

specific heat, 143

thermal conductivity, 87-8, 90


variation of surface energy, 59, 70

9,

15, 17, 132, 135

of magnetic impurities,
145-50
effect of non-magnetic impurities,
141-4
in BCS theory, 129-43
Matthias' rules,

6, 141,

superconducting

Effective charge, 49, 52


Elastic properties, 9

Electron irradiation, 143


Electron-electron interaction,
82-3, 95, 117-25

Energy

148
elements

(table), 5

161-3
Cryotron, 157-63
Cylindrical specimens, 13, 16, 24
{See also thin wires)
cell,

10, 87, 124-5,

131, 134, 143

Demagnetization coefficient, 23
Density of electron states, 104-8,
123, 127, 129, 134, 143
Diffusion, 152
Dilute alloys
critical temperature, 141-4
magnetization curve, 144-5

11,91,96-7

thermal conductivity, 92, 95


tunnelling, 79, 106-9
ultrasonic attenuation, 137-8
dependence on
field, 79-80, 92, 109, 139
phonon spectrum, 132
position, 150-2
quasi-particle energy, 131-3
size, 79, 139
temperature, 90, 103, 108, 129,
139
in BCS theory, 120, 126-34, 139
in thin films, 79-80
elements
of superconducting
(table), 99
order parameter, 52, 79,
139
Meissner effect and perfect
conductivity, 110, 114-15,
120
penetration depth, 1 1
range of coherence, 44

Thomson
14,

150
Ferromagnetism, 145-8
Flux creep, 73
Flux quantization, 32-3, 72, 120
Free energy, 13-4, 19-20, 47-8,
57, 75, 93, 128, 148-9

heat, 95

18,38,78,128

185

G-L

parameter k - cont.
value for negative surface
energy, 59, 66, 67, 70

critical

deduction from
penetration depth, 51-3
supercooling, 51-3, 66
in thin films, 54, 78

normal
Gapless superconductivity, 149

Gauge

Gap

Entropy,

114-15,

relation to

{See also anisotropy of energy


gap)
correlation with TJ9, 134
deduction from
infrared absorption, 98-100
infrared transmission, 113-14
microwave absorption, 100-4
nuclear spin relaxation, 104-6

G-L

rule,

12,

relation to

Debye temperature,

sum

Ferrell-Glover

cont.

specific heat,

effect

Crowe

Dilute alloys

77

13-19,21,86, 132, 135


similarity of reduced field curves,

of

Index

in

in

and

conductivity

specific heat constant,

invariance

69

surface energy, 58

BCS theory, 139


G-L theory, 49

Geometry, influence of, 23-6


Ginzburg-Landau (G-L) theory,

12,

48-54, 150
basic equations, 50, 139
critical field of small specimens,
75-7
extension to lower temperatures,

49,80

temperature dependence, 70
Gorter-Casimir thermodynamic
treatment, 11, 13-19
Gorter-Casimir two-fluid model,
11, 19-21
{See also two-fluid model;
two-fluid order parameter)
application to G-L theory, 49
relation to penetration depth, 36

Gyromagnetic

ratio,

22

free energy, 48-9, 57, 75

limitations, 49, 52-3, 66, 80

non-local modifications, 48, 50


range of coherence, 58
relation to

BCS

theory, 48, 52, 139


London equations, 50, 54

small specimens, 76-80

to

energy gap,

relation to penetration depth, 49,

53^, 78
-,

see

mean

free

path

Infrared absorption, 98-100


Infrared transmission, 45, 99, 109114, 115
Interaction parameter V, 121

BCS

cut-off, 121, 124-5, 131

of
non- magnetic impurities, 1 43-4
magnetic impurities, 149

effect

influence

on

isotope

effect,

124-6
quasi-particle

lifetime

124-5, 131
variation
with
energy, 131
Intermediate state,

effects,

quasi-particle
14,

23-6,

29,

59-61,94

52, 79, 139

G-L parameter

effects

effects

anisotropy, 131

superconductors of second kind.


67-72
supercooling, 65-7
surface energy, 57-9
G-L order parameter, 48
and free energy, 48-50, 57
effect of magnetic field, 53, 78,
139
gradual spatial variation, 49-50,
57-8, 73
proportionality

Impurity

5 1 -3, 58, 66,

and range of coherence, 58

68-70

Isotope effect, 12, 81-3, 95, 117,


124-6
absence in transition metals, 12,
82, 125

Index

186
Isotope effect - cont.

Index

Magnetic impurities, 145-50

of quasi-particle
124-6
in the BCS theory, 124
table of values, 82
effect

life

time,

16, 26, 75

effect, 109,

ideal

Knight shift, 77, 139^K)


Kramers-Kronig relations, 112, 114
Latent heat, 1 5, 78
Lattice parameters, 10
Laves compounds, 147-8

1 ,

depth, 29, 37-8, 43


microscopic implications, 11, 31,
43
non-linear extension, 38
prediction of penetration depth,

29,36

Low

frequency behaviour
diamagnetic description, 22-6
influence of geometry, 23-6
relation to high frequency response and energy gap, 1 14-15
small specimens, 75-80

field

distribution,

field

13-14,

small specimens, 75

superconductors of second kind,


68-9, 156

free path effects on


anisotropy of energy gap, 97, 100
106, 142
critical temperature, 141-4
G-L parameter k, 69
infrared absorption, 100
nuclear relaxation rate, 106
penetration depth, 35, 38, 41-3,

45-6,58, 112
range of coherence, 42-3, 45-7,
152
surface energy, 58-9, 70, 77
Mechanical effects, 16-17, 143

Mendelssohn 'sponge', 78, 155


Microwave absorption, 100-4
Mixed state, 71-74

Neutron bombardment, 143


Nuclear spin relaxation, 104-6
7-8,

field dependence of
energy gap, 79-80
entropy, 37
free energy, 13-14, 49
G-L order parameter, 53, 78
penetration depth, 35, 38-9, 53, 76

Magnetic

superconductors,

24-6,68

Mean

basic equations, 29, 42, 44,46, 1 1


incorrect values of penetration

penetration:

see

Nuclcation of superconducting
phase, 61-3

susceptibility,

34-5, 75, 77, 83

53

Quenching, 143

mean

field, 35,

38-9, 53, 76

free path, 35, 38, 41-3,

range of coherence, 43, 46, 112


79
temperature, 35-8, 51-3, 130
in BCS theory, 37, 52, 130
in Pippard theory, 45-6, 112
incorrectness of London values,
29,37-8,43, 112
methods of measurement, 35-6

superconducting
the
phase, 14, 19
Order parameter, see G-L order
parameter; two-fluid order para-

Order

in

meter

23,

Penetration depth, 29, 58


defining equations, 28, 34, 36

42, 45-7, 152


in

in

energy gap, 115


entropy, 38
frequency variation of conductivity, 111-2
G-L order parameter, 51, 78
surface energy, 55-7
susceptibility, 34-5
thin film critical field, 77
values in superconducting elements (tables), 38, 65
Perfect conductivity of superconductors, 4, 29-30, 114-15
Perfect conductor, 4, 6-8, 27-8
Persistent current, 3, 6, 26, 158-60
Phase propagation, 63-5
Phonon spectrum, 132
Pippard non-local theory, 12, 41-6
(See also range of coherence)
basic equations, 42, 44, 45
field

penetration

46
through thin

films,

45

penetration depth, 42-3, 45-6


reduction to local form (London
limit) 12,

45-6

of thin

58

energy gap, 44
field dependence of penetration depth,

mean

40

free path effect

on

Tc

and

BCS

Relation between magnetic and


thermal properties, 13-21, 86,
96, 132, 135
Rutgers' relation, 15, 17

Semiconductors, superconducting,
6
Silsbee's rule, 5

Similarity, 85-6, 96, 116, 121-2, 130

Size effect

on

76-7, 155
supercooling field, 67
critical temperature, 143
energy gap, 79-80
critical field, 46,
critical

magnetic

susceptibility,

34

penetration depth, 38, 46, 79


range of coherence, 46
Skin depth, 35-6, 43
critical field, 46,

films,

77

75-8

G-L

theory, 46, 54, 67, 75-80


in Pippard theory, 46
in

Pressure effects, 12, 16-17

142
penetration depth, 40, 44-6
sharpness of transition 40-1
surface energy, 57
uncertainty principle 40, 45
values for Al, In, Sn (table), 65

Small specimens

relation to energy gap


susceptibility

BCS theory, 44
G-L theory, 49,

relation to

relation to

theory, 44-5
14,

Range of coherence, 1
and superimposed metals, 150-1
dependence on mean free path,

size, 38, 46,

critical field in thin films,

penetration depth

Magnetic

flux, 32-3, 72, 120


Quasi-particles, 124-5, 132

45-6,58,112

Magnetostriction, 16
Matthias' rules, 6, 141, 148

Lifetime effects, 124-6, 131


Localized magnetic moment, 145-6
London theory, 1 28-32, 36, 41-2

Magnetic
22-4
Magnetic

magnetic

superconductors,

78

133-4

Quantized

cont.

dependence on
field direction, 39,

187

frequency, 39

dilute alloys, 144

filamentary

Josephson

Penetration

Magnetization
area under magnetization curve,

depth -

low frequency behaviour, 75-80

Small specimens -

cont.

penetration depth, 35-8, 43,


45-6, 79
range of coherence, 45-6
Sommerfeld specific heat constant, 9
in dilute alloys, 143

independence of isotopicmass,85
relation to

G-L parameter

Specific heat of the electrons

comparison

of

17,

and

19-21,

dependence on temperature, 9,
11,18,20-1, 86,91,96-7,134

Tc

9, 15, 17, 78,

132, 135

BCS

theory, 130, 135

relation to

21,86, 96
energy gap, 11, 90, 96-7, 132
thermal conductivity, 91, 95
Rutgers' relation, 15, 17
Specific heat of the lattice, 9-10, 18
Spherical specimens
critical field of small spheres, 76
magnetization, 8, 26
penetration
depth
of small
spheres, 29, 35, 46
supercooling in small spheres, 67
critical field, 15, 17,

Spin, effect on Tc , 147


Strain, 26-7, 62, 77, 155

Superconducting alloys and compounds, 5-6


dilute alloys, 58-9, 87-90, 141-5
ferromagnetism, 145-8
high critical fields, 154-6
Laves compounds, 147-8
magnetic impurities, 145-50
Matthias' rules,

intermediate state, 59
phase nucleation and propa-

gation, 61-6
range of coherence, 56-8

154

reversing switches, 154

Superconducting elements

(table),

6,

141

non-magnetic impurities, 141-5


rare earth and transition metal
solutes, 145-9
thermal conductivity, 88-9

Superconducting filaments, 78, 155


Superconducting ring, 3, 8, 26, 30,
32
Superconducting transition
contrast with perfect conductor,
6-8
discontinuity of
specific

heat, 9,

15,

17,

78,

134
entropy, 14
free energy, 13-14,48
in dilute alloys, 144
in thin films, 78-81
length

and

volume

changes,

16-17
reversibility, 7, 13, 17,

144

speed, 159

Superconductors of second kind,


67-73, 155-6
52-3,

values for AI, In, Sn (table), 65


Surface impedance, 35-6, 52-3, 95,
101-4, 111, 139

critical fields

61-3,

65-7,

74

and temperatures,

surface energies, 65

Temperature dependence of
4,

18,

84-6,

130,

135
energy gap, 90, 103, 108, 130
G-L parameter k, 70
penetration depth,

35-8,

51-2,

130
specific heat, 9, 11, 18, 20-1, 86,

91,96-7, 130, 134

Superheating, 61
Superimposed metals, 150-2

Surface currents, 22
Surface energy, 55-74, 150, 156
dependence on temperature, 64-5
effect of strain, 62, 77, 156
in G-L theory, 57-8
in inhomogeneous specimens, 77,

156

energy gap, 91-3, 95


gap anisotropy, 92
specific heat, 91

Thermal expansion coefficient, 17


Thermodynamics of superconductors

BCS theory, 128


G-L theory, 48-9
Gorter-Casimir treatment, 11,
13-19
relation between magnetic and
thermal properties, 13-21, 86,
96, 132, 135

surface energy, 64-5, 70


surface impedance, 101-3, 139
thermal conductivity, 87-94, 136
two-fluid order parameter, 19-21,

36

Thermal conductivity, 87-94


in

BCS

of thin

films

(See also superimposed metals)


critical current, 80
critical field, 47,

energy gap values, 99


energy gap anisotropy, 138
isotope effect exponents, 82
penetration depth values, 38, 65
ranges of coherence, 65
specific heat discontinuities, 17,
135
superconducting elements, 5

theory, 92, 136


films, 79,

93

cont.

relation to

Thin
Table of

critical field,

order, 72, 78-81

Supercooling,

Pippard theory, 56-7


free path effect, 58-9, 67,
77

relation to

86, 134, 135

discontinuity at

devices,

189

Thermal conductivity -

cont.

negative values, 58-9, 62, 67, 69,


70, 157

heat switches, 94, 1 53


leads, 153
magnets, 78, 154-7

rectifiers,

magnetic

calorimetric data,

in

mean

60-4
radiation detectors, 154

69

*,

Surface energy

Superconducting devices
cavities, 154
computer elements, 157-64
d.c. amplifiers, 153-4
galvanometers, 153

memory

19-21,135

critical field,

in

Index

Index

188

75-80

critical thickness for second

order

transition, 78

cryotrons, 157-63

energy gap, 79, 93


infrared transmission, 46, 109-15
in G-L theory, 54, 75-80
in perpendicular field,

magnetic
75-80

behaviour,

73
33,

73-4,

penetration depth, 29, 35-6, 46,


79
relation of critical field to X
b

and

&77
second order transition, 78
supercooling, 67
susceptibility, 34-5, 76, 77
thermal conductivity, 79, 93
total field penetration, 36, 45
variation of G-L order parameter, 78
Thin wires, 29, 35, 67, 76
Thomson heat, 95
Threshold magnetic field: see critical magnetic field
Time-reversed
wave
functions,
144

Index

190

Transition metals
absence of isotope effect, 12, 82,

125-6
effect

on

Trapped

Tc

145-50

flux, 3,

26-7, 32, 120, 144,

161
Tunnelling, 106-9, 132-4

Two-fluid model
(See also Gorter-Casimir twofluid

and

model)

BCS

theory, 127
extension of G-L theory, 49

Two-fluid order parameter, 19-21,


36
gradual spatial variation, 40,
56-7
relation to

penetration depth, 36
surface energy, 56-7
thermal conductivity, 89
rigidity in London theory, 39

Ultrasonic attenuation, 105, 136-8

Uncertainty principle, 31, 40, 45

relation to

nuclear spin relaxation, 104


penetration depth, 36
thermal conductivity, 87

Valence elections,
143, 145

Vortex

lines,

72-4

effect

on

Tc

6,

s>

Monographs on Physical Subjects

continued

IONIZATION AND BREAKDOWN IN GASBS


F. Llewellyn Jones

LOW TEMPERATURE PHYSICS

L. C. JaCKSOl

magnetic amplifiers George M. EtringC


magnetic materials F. Brailsford
MASERS AND LASERS G. J. F. TrOUp
THE MEASUREMENT OF RADIO ISOTOPBS

LIVERPOOL

Taylor

MBCHANICAL AND ELECTRICAL VIBRATION


J.

turn this

book

to th

2 the last date

si

R. Barker

mercury arcs F. J. Teago


THE METHOD OF DIMENSIONS Alfred W. P
MICROWAVE LBNSBS J. Brown
microwavb spectroscopy M. W. P. Stran
MOLECULAR BEAMS K. F. Smith
NUCLEAR RADIATION DBTBCTORS J. Sharp
the nuclear reactor Alan Salmon
optical masers O. S. Heavens
ORDER-DISORDER PHENOMENA E. W. ElCQ
PHOTONS AND ELECTRONS K. H. Spring
PHYSICAL CONSTANTS W. H. J. Childs
physical formulae T. S. E. Thomas
THE PHYSICAL PRINCIPLES OF WIRBLBSS
J.

A. Ratcliffe

PRINCIPLES OF APPLIED GEOPHYSICS


D. Parasnis
relativity physics W. H. McCrea

sbismology K. E. Bullen
SBMI-CONDUCTORS D.A.Wright
SHOCK TUBES J. K. Wright
THB SPECIAL THBORY OF RELATIVITY

M<

Dingle

superconductivity Ernest A. Lyman


THE THBORY OF GAMBS AND LINEAR

gramming

S. Vajda

THERMIONIC VACUUM TUBBS


Sir

W. H. Aldoi

Edward Appleton

thermodynamics Alfred W. Porter


wave filtbrs L. C. Jackson
wave guides H. R. L. Lamont
WAVB MBCHANICS

H. T.

Flint

x-ray crystallography R. W. James


x-ray optics A. J. C. Wilson

Printed in Great Britain

Monographs on Physical Subjects


ALTERNATING CURRENT MEASUREMENTS David Owen
APPLICATIONS OF INTERFEROMETRY W. Ewait Williams
APPLICATIONS OF THERMOELECTRICITY H. J. Goldsmid
atmospheric electricity B. F. J. Schonland
ATMOSPHERIC TURBULENCE O. G. Sutton
atomic spectra R. C. Johnson
cartesian tensors George Temple
classical mechanics J. W. Leech
THB CONDUCTION OF ELBCTRICITY THROUGH GASES K. G.
Emeteus

the cosmic radiation

J.

E.

Hooper and M. Scharff

dielectric aerials D. G. Kiely


dipole moments R. J. W. le Fevre
the earth's magnetism Sydney Chapman
ELASTICITY, FRACTURE AND

FLOW

J.

,4

H.

J.

C^
^3
^_{

^t

Josephs

high energy nuclear physics W. Owen Lock


HI-GH FREQUENCY TRANSMISSION LINES Willis Jackson
INTEGRAL TRANSFORMS IN MATHEMATICAL PHYSICS C.

J.

Tranter

AN INTRODUCTION TO ELECTRONIC ANALOGUE COMPUTERS


M. G. Hartley
AN INTRODUCTION TO ELECTRON OPTICS L. Jacob
AN INTRODUCTION TO FOURIER ANALYSIS R. D. Stuart
AN INTRODUCTION TO THE LAPLACE TRANSFORMATION
C. Jaegar

AN INTRODUCTION TO SERVOMECHANISMS A. Porter


AN INTRODUCTION TO VECTOR ANALYSIS B. Hague
AN INTRODUCTION TO TENSOR CALCULUS AND RBLATIVITY
Derek F. Lawden
AN INTRODUCTION TO PHASE-INTEGRAL METHODS
J.

G. O. Jones

heaviside's electric circuit theory

J.

^3
^J

C. Jaegar

THB electric arc J. M. Somerville


elements OF pulse circuits F. J. M. Farley
BLBMENTS OF TENSOR CALCULUS H. Lichnerowicz
fluid dynamics G. H. A. Cole
frequbncy modulation L. B. Arguimbau and R. D. Stuart
FRICTION AND lubrication F. P. Bowden and D. Tabor
FUNDAMENTAL OF DISCHARGE TUBE CIRCUITS V. J. Francis
gbnbral circuit theory Gordon Newstead
THE GENERAL PRINCIPLES OF QUANTUM THEORY G. Temple
glass

CQ
^J

Heading

[continued on back flap]

Das könnte Ihnen auch gefallen