You are on page 1of 512

-J,

This is a text on the subject of superconductivity, an area of intense research activity worldwide. The book is in three parts: the first deals with phenomenological aspects of superconductivity, the second with the microscopic theory of uniform superconductors, and the third with the microscopic theory of nonuniform superconductors. The first part of the book covers the London, Pippard, and Ginzburg-Landau theories, which are used to discuss a wide range of phenomena involving surface energies, vorticity, the intermediate and mixed states, boundaries and boundary conditions, the upper critical field in bulk, thin-film, and anisotropic superconductors, and surface superconductivity. The second part discusses the microscopic theory of Bardeen, Cooper, and Schrieffer. Finite temperature effects are treated using the Bcgoliubov-Valatin transformation. The theory is used to discuss quasiparticle tunneling and the Josephson effects from a microscopic point of view. The final part of the book treats nonuniform superconductors using the Bogoliubov-defiennes approach with which it is possible to extract many important results without invoking Green's function methods. This text will be of great interest to graduate students taking courses in superconductivity, superfluidity, many-body theory, and quantum liquids. It will also be of value to research workers in the field of superconductivity.

Superconductivity

Superconductivity
1. B. KETTERSON
Northwestern University

and S. N. SONG

CAMBRIDGE

...-----.---- ..--.-~ RUHR-UNlVERS:Ti'T ,-,. .: '.'


Fak. f. Phys':':~;
p.-:-,.;'
i'" ~.

UNIVERSITY PRESS

"',

In",v:r~~. ~;;'"

60 ~,·,5:s.

PUBLISHED

BY TIlE

PRESS

SYNDIC."TE

OFTHE

UNIVERSITY

OF

CAMBRIDGE

The Pitt Building, Trumpington


CA:VIBRIDGE UNIVERSITY PRESS

Street, Cambridge

CB21RP,

United Kingdom

The Edinburgh 10 Stamford

Building, Cambridge

CB2 2RU, LK 3166, Australia

http://www.cup.cam.ac.uk httpr/www.cup.org

40 West 20th Street, New York. NY 10011-4211, USA Road, Oakleigh, University Melbourne

<£) Cambridge

Press 1999 Subject to statutory exception

This hook is in copyright. and to the provisions no reproduction the written permission First published 1999

of relevant collective licensing agreements, of Cambridge University Press.

of any part may take place without

Printed in the United Kingdom

at the University [VN]

Press, Cambridge

Typeset in 9.5/13pt Times New Roman A catalogue recordfor Library a/Congress Ketterson, p.

this hook is available/rom

the British Library

Casaloquinq in Publication data

J. B. (John Boyd) - J. B. Ketterson (hardcover). and S. N. Song. - ISBN 0-521-56562-6 1. Song, S. N. (pbk.) ern. 2. Superconductors.

Superconductivity

ISBN 0--521-56295-3 1. Superconductivity. (Shengnian N.) QC611.95.K48 621.3'5-dc21 II. Title. 1998

97-3060

CTP

ISBN 0 521 562953 hardback ISBN 0 521 565626 paper hack

Contents

Preface Part I Phenomenological theories of superconductivity Introduction The London-London equation Pippard's equation Thermodynamics of a Type I superconductor The intermediate state Surface energy between a normal and a supcrconducting Quantized vorticity Type II superconductivity 8.1 Magnetic fields slightly greater than Hc1 8.2 The region Hc1 < < H < < Hc2 8.3 Microscopic magnetic probes of the mixed state The Ginzburg-Landau theory 9.1 Basic equations 9.2 Gauge invariance 9.3 Boundaries and boundary conditions The upper critical field of a Type II superconductor The anisotropic superconductor Thin superconducting slabs Surface superconductivity The Type II superconductor for H just below Hc2 The Josephson effects 15.1 The Josephson equations 15.2 Magnetic field effects: the two-junction SQUID 15.3 The extended Josephsonjunction 15.4 Effect of an applied rf field 15.5 The resistively shunted junction (RSJ) model 15.6 The rf biased SQUID The Josephson lattice in ID Vortex structure in layered superconductors 17.1 3D anisotropic London model

Xlii

1 2 3 4 5 6 7 8

metal

10

11

12
13

14 15

5 8 12 15 17 19 24 24 27 29 31 31 35 36 43 45 47 50 52 59 59 61 64 65 66
73

16 17

77

82 82
vii

VIII

~onrenrs

17.2 Lawrence-Doniach model 17.3 Vortex struct ure in a 2D film 18 Granular superconductors: the Josephson lattice in 2D and 3D 19 Wave propagation in Josephson junctions, superlattices, and arrays 19.1 Wave propagation in a junction 19.2 Wave propagation in a superlattice 19.3 The Josephson transmission line 20 Flux pinning and flux motion 20.1 Nonideal Type II superconductors 20.2 Microscopic description 20.3 The Lorentz force 20.4 Pinning centers and pinning forces 20.4.1 The core interaction 20.4.2 Surface magnetic interaction 20.4.3 Summation of the pinning forces 20.5 The equation of motion 20.6 The critical state 20.7 The elastic constants of a flux-line lattice 20.8 Collective flux pinning 20.9 Mechanisms of flux motion 20.10 Relaxation of the magnetization with time 20.11 Phase diagram of high To oxide superconductors 21 Time-dependent G~L theory 22 Fluctuation effects 22.1 The Ginzburg criterion 22.2 The diamagnetic susceptibility for T> t; 22.3 Paraconductivity for T> t; 23 G~L theory of an unconventional superfluid 23.1 The order parameter of an unconventional superfluid 23.1.1 Superfluid 3He: isotropic p-wave pairing 23.1.2 Isotropic d-wave pairing 23.2 Crystal-field and spin-orbit effects 23.3 TheG L theory of an unconventional superfluid 23.3.1 G L theory for 3He 23.3.2 G L theory for an isotropic d-paired superfluid 23.3.3 Unconventional G~L theory in metals 23.4 Inhomogeneities in the order parameter 23.5 Collective modes in an unconventional superfluid 23.5.1 Collective modes of 3 He B 23.5.2 E state collective modes 24 Landau Fermi liquid theory 24.1 Basic equations 24.2 Collisionless collective modes 24.2.1 The kinetic equation 24.2.2 Collisionless longitudinal zero sound

85 87 91 95 95 104 108 110 110 111 113 115 116 117 118 118 120 123 127 131 134 137 140 144 144 147 149 152 154 154 157 158 163 163 168 168 171 172 173 177 180 180 187 187 189

24.2.3 24.2.4

Collisionless transverse zero sound Collisionless spin waves

190 191

Part"

The microscopic

25 26 27 28 29 30 31

32 33 34

35

theory of a uniform superconductor The Cooper problem: pairing of two electrons above a filled Fermi sea The BCS theory of the superconducting ground state Elementary excitations: the Bogoliubov Valatin transformation Calculation of the thermodynamic properties using the Bogoliubov-Valatin method Quasiparticle tunneling Pair tunneling: the microscopic theory of the Josephson effects Simplified discussion of pairing mechanisms 31.1 The electron-phonon interaction 31.2 The spin fluctuation mechanism The effect of Coulomb repulsion on To The two-band superconductor Time-dependent perturbations 34.1 Ultrasonic attenuation 34.2 Nuclear spin relaxation Nonequilibrium superconductivity 35.1 Elastic and inelastic scattering processes 35.2 Quasiparticle and phonon populations in a noncquilibrium superconductor

195 199 208 212 216 222 230 230 234 240 243 245 248 248 251 251 254

Part '"

Nonuniform

superconductivity

Bogoliubov's self-consistent potential equations Self-consistency conditions and the free energy Linearized self-consistency condition and the correlation function 38.1 Treating the gap function as a perturbation 38.2 Relation to a correlation function 39 Behavior of the correlation function in the clean and dirty limits 39.1 A simple model for the clean limit 39.2 The dirty limit 39.3 The general case 40 The self-consistency condition 40.1 The dirty limit at zero magnetic field 40.2 The dirty limit at finite magnetic field 40.3 The clean limit at zero magnetic field 41 Effects involving electron spin 41.1 Spin generalized Bogoliubov equations 4l.2 The density matrix 4l.3 The linearized gap equation 41.4 Spin-dependent potentials

36 37 38

257 262 265 265 267 272 272 273 275 277 277 279 282 284 284 286 288 289

Contents

Paramagnetic impurities, electron paramagnetism, and spin orbit coupling 41.6 The Fulde-Ferrcll state 41.7 Gapless superconductivity 42 Boundary conditions 43 The proximity effect at zero field 43.1 Governing equations 43.2 The thin-film (Cooper-deGennes) limit 43.3 The general1D case 43.4 A microscopic theory of the ID Josephson superlattice 44 The proximity effect in a magnetic field 44.1 Governing equations in the presence of magnetic fields, spin susceptibility, paramagnetic impurities, and spin -orbit coupling 44.2 Representative numerical solutions 45 Derivation of the G-L theory 45.1 The first G-L equation 45.2 The gradient term in the clean limit 45.3 The gradient term in the dirty limit 45.4 The gradient term in the general case 45.5 The second G-L equation 46 Gauge invariance: diamagnetism in the low field limit 46.1 Gauge invariance 46.2 The magnetic field as a perturbation 46.3 The diamagnetic current 46.4 Diamagnetism of the superconducting Fermi gas 46.5 Magnetic field behavior near a vacuum-superconductor interface 46.6 Relation between normal-state conductivity and the superconducting diamagnetic response 46.7 Calculations of the diamagnetic response using Chambers' method 47 The q uasiclassical case 47.1 Quasiclassicallimit of the Schrodinger equation 47.2 Quasiclassical limit of the Bogoliubov equations 47.3 Andreev scattering 48 The isolated vortex line 48.1 Bogoliubov's equations for the isolated vortex line 48.2 The quasiclassical equations for a vortex line 48.3 A model calculation for the bound core states 49 Time-dependent Bogoliubov eq uations 49.1 Basic equations 49.2 The time-dependent, linearized, self-consistency condition 49.3 The linearized, time-dependent, G-L equation 50 The response of a superconductor to an electromagnetic field 41.5

290 293 295 298 304 304 306 307 314 316

316 321 326 326 327 328 329 330 332 332 332 335 338 341 343 347 350 350 353 355 360 360 363 365 369 369 370 372 374

Contents

xi

51 52

53

54

55

50.1 The vector potential as a time-dependent perturbation 50.2 Relation between the current density and the vector potential The Bogoliubov equations for an unconventional superfluid Mean potentials, density matrices, and distribution functions in a uniform superfluid 52.1 The mean potentials 52.2 Singlet-triplet separation 52.3 Density matrices and distribution functions Superfluid 3He 53.1 Some experimental properties of super fluid 3He 53.2 Structure of the gap in an t = 1, S = 1 superfluid 53.3 The Balian-Werthamcr (BW) phase 53.4 The Anderson-Brinkman-Morel (ABM) or A phase Collective modes in normal and supcrfluid Fermi systems 54.1 General formalism 54.2 Zero sound in a normal Fermi liquid 54.3 Collective modes in superfluid 3He B Green's functions 55.1 Green's functions for the Bogoliubov equations: the Nambu-Gorkov equations 55.2 Quasiclassical Green's functions in a normal system 55.3 Quasiclassical Green's functions in a superfluid system 55.3.1 Keldysh formalism 55.3.2 Quasiclassical equations of motion 55.3.3 Eilenberger equations 55.4 Temperature Green's functions 55.5 Dirty superconductors 55.5.1 The dirty-normal-metal Green's function 55.5.2 The dirty superconductor Green's function 55.6 G-L theory revisited 55.7 The Usadcl equations Identical particles and spin: the occupation number representation A.l The symmetry of many-particle wavefunctions A.2 Occupation number representation: Bose statistics A.3 Occupation number representation: Fermi statistics Some calculations involving the BCS wavefunction The gap as a perturbation through third order
transition temperature, thermodynamic criticaljield, and specific heat coefficientfor the elements

374 376 380 384 384 389 393 397 397 397 401 404 406 406 412 414 421 421 426 430 437 443 443 437 443 443 447 451 452

Appendix A

Appendix B Appendix C
Superconductinq

458 458 461 465 467 470


475

Debye temperature References Additional Index readinq List of mathematical

476
481 and physical symbols 483 493

Preface

A preface is supposed to alert potential readers to the contents and style of what lies within so they can decide whether to proceed further. Superconductivity is now a vast subject extending from the esoteric to the very practical; the people who study or work on it have different preparations, goals, and talents. No treatment can or should address all these dimensions. Part I is devoted to the phenomenological aspects of superconductivity, e.g., London's and Pippard's electrodynamics, the Ginzburg-Landau theory and the Landau Fermi liquid theory. These theories allow a discussion of the effects of magnetic fields, interfaces and boundaries. fluctuations, and collective response (which may all be thought of as different manifestations of inhomogeneities). Since there is currently much interest in unconventional (non-s-wave) superconductivity, we have included a discussion of the associated Ginzburg-Landau theory (which then has a multidimensional, complex order parameter). 3He is the only established example of an unconventional superftuid (triplet p-wave) and therefore our discussions of this substance are somewhat longer. Part II is devoted to the microscopic theory of uniform superconductors: the theory of Bardeen, Cooper, and Schrieffer (BCS) and the Bogoliubov-Valatin canonical transformation, where the latter so greatly simplifies the discussion of excited states and finite temperature effects. Although not strictly a uniform-superconductor phenomenon, the theory of tunneling and the accompanying Josephson effects are also discussed in Part II. Part III deals with the microscopic theory of nonuniform superconductors exclusively through the apparatus of the self-consistent Bogoliubov equations as developed extensively by deGennesand coworkers (and hence sometimes referred to as the Bogoliubov-decIennes theory). Inhomogeneities associated with a magnetic field, impurities, and boundaries are discussed; temporal 'inhomogeneities' (e.g., relaxation phenomena and collective modes) are also discussed via the time-dependent Bogoliubov equations. Bogoliubov theory is extended to include unconventional BCS superl1uids, such as 3Hc and (possibly) high temperature superconductors and heavy fermion superconductors (UPt3). Part III ends with a simplified discussion of Green's functions starting from the Bogoliubov theory. This formalism serves as an appropriate departure point for those wishing to go deeper into the theory of superconductivity. Real time Green's functions, thermal or Ma tsu bara Green's functions, as well as q uasiclassical Green's functions are all discussed. Our goal in this book is to focus primarily on the physics of superconductivity. The materials aspect is generally ignored; material properties enter the discussions only via idealized

xiv

Preface

parameters. For this reason the book may not contain enough material on high temperature superconductors to suit some readers' taste, although idealized layered systems are treated in some detail, However, it is our view that much of the high T; literature does, in fact, involve the systematics of the superconducting properties with regard to their chemical and physical makeup. On the other hand, the underlying physical pairing mechanism and the order parameter symmetry are still controversial at this time (1997). The level of the treatment varies between sections. Much of Part I is accessible to fourth year undergraduate students or first year graduate students in physics who have had exposure to mechanics, electricity and magnetism, and elementary quantum mechanics. Parts II and III involve some use of second quantization at a level usually arrived at by the end ofthe first year of graduate school. With some omissions the book could form the basis of a one semester graduate course on superconductivity; if all topics are discussed it would likely extend to a year. The microscopic origin of the attractive electron-electron interaction involving the exchange of phonons is discussed (in Part II) only in terms of the jellium model which is sufficient to bring out the basic physics of the electron-phonon interaction. A proper discussion, however, requires solving the coupled Green's function equations of motion for the electrons and phonons, which is beyond the scope of this book. A simplified discussion of the spin fluctuation pairing mechanism is included since pairing in superfluid 3He likely arises from this eJTect.Other pairing mechanisms, as have been proposed for high T; materials, are ignored entirely. The character of many of our discussions was strongly influenced by deGennes' 1966 classic, Superconduct ivity of AIetals and Alloys; any similari ty of our discussions and his is likely intentional. Both of the present authors are primarily experimentalists; their involvement in writing a largely theoretical book about superconductivity was primarily a self-education excrcise. We have tried to use the simplest mathematical methods we can while minimizing the (pedagogically useless) 'it-can-be-shown' approach. Our decision to use a uniform theoretical approach to the topics in Part III meant that in many cases we had to devise our own methods; in so doing we have strived to get things right but inevitably errors will creep in (herc and elsewhere), for which we apologize in advance. Incidentally we would appreciate hearing about any errors detected. This book is not meant to be a treatise on superconductivity (for this kind of treatment we recommend the well-known text edited by R. Parks); i.e., it is not our purpose to collect most of what is known and useful, but rather to present the nuts and bolts used to obtain some of the important results. The literature on superconductivity is enormous; it is therefore not possible to be aware of anything but a small fraction of it (this situation was 'worsened' by the discovery of high temperature superconductivity). For this reason we decided to minimize the number of original citations to those which are particularly important historically or pedagogically (a difficult and dangerous decision) or to those where the reader may seek further details. We hope the many workers not cited will forgive us. We have benefited from discussions with (or encouragement by) P. Auvil, K. Bennemann, S. Doniach, A. Patashinskii, P. Wolfle, and S. K. Yip. The text was typed (again and again) by Ms A. Jackson who deserves the most thanks. Northwestern University J. B. Ketterson S.N. Song

PART I

PHENOMENOLOGICAL THEORIES OF SUPERCONDUCTIVITY

Introduction

Many metals, alloys, and intermetallic erally low) temperature, enon which is call cd superconductivity. in 1911; it has been sensitively supercurrent terials exhibit phase transition may not vanish; using a sensitive no detectable probed decay technique

compounds

undergo

a phase resistance

transition

at some (genOnnes maof the

Tc' to a state having zero electrical


Superconductivity by observing of this current materials such as nuclear

(see Fig. 1.1), a phenomby Kamerlingh Favorable by a circulating

was discovered magnetic

the magnetic for periods

field produced resonance. limited conditions exceptions

by the patience

observer ( .~ years). However,

superconductivity

is better defined by the nature or measurement of these and

of the associated initially restrict

and it then includes here we postpone

where the resistance

the discussion

ourselves to perfect superconductors.


0.001

0.0015

I
!
:

, , , , ,

, , ,

.>V

j
1

~ ~

:::: 0.0010

'Hg

0.0005 < 10-6

I
I
4.30
T(K) 4.40

0.0000

-wo

\'
4.20

4.10

450

Figure 1.1 The original R vs T curve of Kamcrlingh Onnes showing the superconducting transition in mercury. (After Kamerlingh-Onnes (1911 ).) A second fundamental magnetic flux, discovered characteristic associated with a superconductor is the exclusion of

by W. Meissner

and R. Ochsenfeld

in 1933 (which we will often refer to

1. An alloy is a solid solution of two or more clements at least one of which is a metal. An intermetal1ic compound involves a metal and one or more other elements which form a chemical compound having nearly precise ratios of its constituents.

Part I

Phenomenological theories

Experiment

2:

Perfect conductor

Supercond uctor

Perfect conductor

Supercondncto

:'I

/; 1 t
l·rD.

iII

i II \,
,(" 1

I) r f./r\\.
1,1 I
. \''__/j

'\ \, ~! II
I.I \.I
:

..1,

fr~\ I\VJi \\ , (
!\
i
I

i/

, ill . I'

\I

i.:
I

\~J
r\

~ 1

Figure 1.2 Hypothetical experiments showing the difference between a superconductor and a perfect conductor. Experiment 1: sample cooled in zero magnetic field after which a field is applied; experiment 2: sample cooled in applied magnetic field.

simply as the Meissner effect in what follows). The degree of flux exclusion can depend on the material or measurement conditions; we will also assume this property is nearly perfectly displayed for our initial discussion. The combination of zero resistance and perfect diamagnetism results in a clear distinction between a superconductor (which as we will see is in a thermodynamic state) and a hypothetical 'perfect conductor' (which has the unique transport property of zero resistance); this difference is illustrated in Fig. 1.2 and involves the differing response each would have for different histories of cooling below the transition temperature and applying a magnetic field. If we start by cooling through T, and then apply a magnetic field, both the superconductor and the perfect conductor would exclude the field, For the perfect conductor, induced currents arising from Faraday's law would screen the flux, Flux exclusion in the superconductor could be assigned to the same mechanism or (more fundamentally) to the Meissner effect itself. If we reverse the order by first applying the field and then cooling through Te, the superconductor and the perfect conductor beha ve differently: the superconductor excludes the flux (the Meissner effect); the perfect conductor would remain fully permeated by the field. These experimental observations argue that the transition associated with superconductivity is indeed a phase transition since an equilibrium thermodynamic state is defined by its independent thermodynamic variables (in this case T and H), and is independent of its history (which as we see is not true for the perfect conductor). Superconductivity, and with it the Meissner effect, does not persist to arbitrarily high magnetic fields, For each temperature there is a well-defined critical field, He(T), at which superconductivity disappears." Fig. 1.3 shows a universal curve of the temperature dependence of He vs T.
2. This statement is restricted to so-called Type I materials having a shape for which there is no demagnetization effect, as will be discussed in later sections.

Introduction

Normal

Superconducting

Figure 1.3 The temperature dependence of the critical field H J T). , 8x 10-J
caVg
KI

I"
Tin

<;
si:
Q)

'"' en

U
0.

/
/

".~.:;' .,,'-::. .' .: o o


Temperature

.'

.'.' ...

Figure 1.4 Specific heat of tin as a function of temperature: open circles without an external magnetic field; filled circles with an external field H > He' Shown also are the individual contributions from the electrons and the lattice for H > He: chain line lattice contribution; dashed line - electron contribution. (After Keesorn and van Laer
(1938).)

A superconductor also exhibits a discontinuous increase in its specific heat at T; (there is no latent heat at zero field) below which it drops rapidly (approaching an exponential dependence at low temperatures), A discontinuity in the specific heat is a signature of a second order phase transition providing added evidence that superconductivity is associated with a distinct thermodynamic phase. Fig. 1.4 shows the heat capacity of a typical superconductor. There is a related low temperature phenomenon known as superfluidity which occurs in liquid helium. There are actually two such liquids involving the isotopes 3He and 4He having superfluid transitions near 2 x 10- 3 K and 2 K, respectively, depending on the pressure. For some purposes a superconductor may be regarded as a superfluid having an electric charge. 4He superfluidity involves a Bose condensation, a phenomenon which is related to superconductivity

Part I

Phenomenological theories

in a somewhat subtle or indirect way and will not be discussed in this book. The superfluidity of 3He is intimately related to superconductivity and will be addressed in later sections. Another related superfluid is the neutron liquid in the interior of neutron stars.

equation

The London-London

Wenow present two derivations properties of superconductors. (see also London

of equations

which are useful in describing used by F. London equation

many of the magnetic and H. London for electrons (1935) in a

The older approach,

(1950)), starts with the Drude-Lorentz

of motion to v/r:

metal, which is Newton's

law for the velocity, v, of an electron

with mass, m, and charge,

e, in an

electric field, E, with a phenomenological

viscous drag proportional

(2.1)

For a perfect conductor

---+

cc, Introducing

the current

density j

nev, where n is the conduc-

tion electron density, Eq. (2.1) can be written

dj ne1 -=-E dt In
which is referred

'

(2.2)

to as the first London

equation.

The time

derivative

of Maxwell's

fourth

equation (in cgs units) is DH V x --

at

=C

4rr oj
-

ot

+ - -. 2
c

r. o2E

Dt ·
the curl of (2.3) and using (2.2) we have
2

(2.3)

where

is the dielectric

constant.

Taking

Vx

V x --

CH)
at

= (4rrne _-

m('

r. ( + - -,)

c Dl-

V x E;

(2.4)

using V x E = - (ljc)DH/Dt

we have

Vx
where we introduced

V (a) x -. H at

a + ( -1 + -I: -.- ) -a H ;.~ Dt2 at


1
('2

o..

(2.5)

the London

depth, i.L, defined by


(2.6)

--,-. me:
Eq. (2.5) has been obtained experimentally observed for a perfect conductor model. exclude In order to conform with

the

Meissner

effect, we must

time-independent

field solutions

Part I

Phenomenological theories

arising from integrating (2.5) once with respect to time and we therefore write V
X

(V x H)

8 () + (1-:;-,----::;2 H + C~ -:;-. J' or

0;

(2.7)

this is referred to as the second London equation. In what follows we will refer to Eq, (2.7) simply as the London equation. An alternate derivation of (2. 7) is motivated by the idea that some of the moving electrons behave collectively as a superfiuid, a liquid possessing no viscosity. This concept is borrowed from the physics of liquid 4He; below 2.l9 K this system behaves as if it were composed of a mixture of two liquids: a superfluid, having no viscosity, and a normal liquid, having a finite viscosity. We assume that the free energy of a superfiuid consists of three parts
(2.8)

where F N is the free energy associated with the normal liquid, Ekin is the kinetic energy of the moving superfluid, and Emag is the magnetic field energy. We may write these latter two terms as
(2.9)

and
(2.10)

where p(r) is the mass density associated with the supcrfluid. Writing p = 11m and v using the fourth Maxwell equation V x H = (4rr/c)j, Eq, (2.l0) becomes

= (l/ne)j

and

(2.l1)

n IS now interpreted as the density of supcrconducting electrons, We will assume that the superconducting electrons adjust their motion so as to minimize the total free energy; this requires c5(Emag + Ekin) = 0 or

If

H(r)' bH(r)

me? + --2 4rrne

[V x H(r)] . [V x bH(r)]

d 3r = 0,

(2.12)

where bH(r) is a variation of the (initially unknown) function H(r). Integrating the second term by parts (and placing the resulting surface outside the superconductor) we obtain

[H(r)

+ A~V x

(V x H)]' c)H(r)d3r

o.

(2.13)

Since the variation c'iH(r) is arbitrary, the term in the square brackets must vanish; therefore V x (V x H)

+ ic~H

0,

(2.14)

which is eq uivalent to (2.7) (including the displacement term in Maxwell's equation yields the last term in (2.7), which is negligible for most applications). As a simple application of Eq. (2.l4) we now discuss the behavior of a superconductor in a

The London-London

equation

magnetic field near a plane boundary. Consider first the case of a field perpendicular to a superconductor surface lying in the x y plane with no current flowing in the z direction. From the second Maxwell equation, V' H = 0, we obtain DHz/uz = 0 or JJ = const. From the fourth Maxwell equation, V x H = (4n/c)j, the first term in (2.14) vanishes and hence H = 0 is the only solution. Thus a superconductor exhibiting the Meissner effect cannot have a field component perpendicular to its surface. As the second example consider a field lying parallel to the superconductor surface, e.g., H II X,which we may write as H = H(z)x (which satisfies V' H = 0). Using the vector identity V x (V x H) Eq. (2.14)becomes
__

V(V'H)

- V2H,

(2.15)

(j2H (jZ2

x __

Al

H =0
x

(2.16)

or (for a superconductor occupying the region z > 0)


(2.17)

A field parallel to the surface is therefore allowed; however, it decays exponentially, with a characteristic length, Au in the interior; ).L (T = 0) ranges from 500 to 10 000 A, depend ing on the material. Accompanying this parallel field is a surface current density, which, from Maxwell's fourth equation, is
j(z) = -

-.-HAO)c-Z!'.LY. 4nl'L

(2.18)

This current density shields or screens the magnetic field from the interior of the superconductor.

Gf----p-;-p-p-a-r-d-'-S-e-q-u-a-t-;-o-n---

At temperatures well below the superconducting transition temperature the heat capacity of a superconductor displays an exponential behavior C ~ e - MkB T (see Fig. 1.4).This suggests that the conduction electron spectrum develops an energy gap, L'\ (not to be confused with the gap in a semiconductor); recall that electrons in normal metals have a continuous (gapless) distribution of energy levels near the Fermi energy, fl. On dimensional grounds one can construct a quantity having the units of length from L'\ and the Fermi velocity, 1!F; we define the so-called coherence length by (3.1 ) This length bears no resemblance to the London depth, Au and hence represents a different length scale affecting the behavior of a superconductor; it can be interpreted as a characteristic length which measures the spatial response of the superconductor to some perturbation (e.g. the distance over which the superconducting state develops at a normal metal-superconductor boundary). Length scales of this kind were introduced independently by Ginzburg and Landau (1950) and by Pippard (1953).1 We first discuss Pippard's phenomenological theory (which scmiquantitatively captures the main features of the microscopic theory to be discussed later). We begin by writing London's equation in an alternative form. Substituting V x H = (4n/c)j (from the fourth Maxwell equation) in the London equation yields
. c VXJ=---H ne2 =--H. me

4n}~
(3.2)

We next write H = V x A. where A is the magnetic vector potential, and restrict the gauge to satisfy
V·A

(London gauge)

(3.3)

and the boundary condition


I. These length scales are not identical, however; the Pippard length is temperature-independent while the Ginzburg-Landau length depends on temperature. The Pippard coherence length is related to the ReS coherence length. 8

Pippard's equation

(3.4) where An is the component of A perpendicular to the superconductor may then be written surface. London's equation

i=

--A.

ne2 me

(3.5)

Note that the condition (3.4) builds in the reasonable boundary condition that the normal component ofthc supercurrent.j., vanishes at a boundary (this is a good boundary condition at a superconductor-insulator boundary but will require modification for metal-superconductor or superconductor-superconductor boundaries). To generalize (3.5), Pippard reasoned that the relation between j and A should be nonlocal, meaning that the current j(r) at a point r involves contributions from A(r') at neighboring points r' located in a volume with a radius of order ~o surrounding r. The mathematical form he selected was guided by the non local relation between the electric field, Etr'], and the current, j(r), which had been developed earlier by Chambers (1952). The expression employed by Pippard was J.(r ) =
-

f[A(r')'

R]R

R4

c -R'<odJ ,.

rr

'

(3.6)

where R = r - r'. The constant C is fixed by requiring (3.6) reduce to (3.5) in the quasiuniform limit where we may take A from under the integral sign; we then have C fCOs2(!dQ fe - R!~odR
=

ne , me

(3.7)

from which we obtain C = 3ne2 /4n~omc. Pippard's generalization of London's equation is then
J(r) = - ---

,.

3ne .f[A(r')'R]R 4n;ome R4

-R~

!,Od r.

J,

(3.8)

Since Eq. (3.8) involves two functions. A(r) and j(r), a complete description requires a second equation which is obtained by substituting H = V x A in the fourth Maxwell equation to obtain Vx(VxA)=~j 4n
e (3.9)

(resulting in an overall integrodifferential equation for A). Eq. (3.8) applies only to a bulk superconductor. An important question we would like to examine is the behavior of a magnetic field near a surface, which will require a modification (or reinterpretation) of (3.8). To model the effect of the surface the integration over points r' is restricted to the interior of the superconductor. If the surface is highly contorted, then it can happen that two points near the surface and separated by about a coherence length cannot be connected by a straight electron trajectory, without passing through the vacuum; one then has to account for this shadowing effect. We restrict ourselves here to plane boundaries which we take to be normal to the z direction. In the limit AL > > ~o Eq. (3.8) reduces to the London equation, as discussed above. (By expandingA(r') in a power series in R, we may obtain corrections to the London equation due to nonlocality.) In the opposite limit, AL < < ;0' A(r') varies rapidly. Let us assume that A(r) falls off

10

Part I

Phenomenological

theories

over a characteristic distance ;_ (which we will determine shortly through a self-consistency argument). When )~ < ~O, the value of the integral (3.8) will be reduced roughly by a factor (A./(o);
I.e.,

j(r)

;. ne2
=-~-

C;o mc

A(r). form

(3.10)

We may also write (3.10) in the 'London-like'

(3.11) This equation has solutions which decay in a characteristic consistency we set this length equal to A: length (I.~~0!;_)1/2; to achieve self(3.12) (A more rigorous derivation from the microscopic theory carried out in Sec. 46 yields 2 = 0.62i.t~0.) We conclude that in the Pippard limit the effective penetration depth A is larger than the London depth, )'L: ;'!;'L = (~oI)L)l/.l > 1. At the same time ), remains smaller than the coherence length: }'/(o = (I'LgO)2/3 < l. If our metal has impurities it is natural to assume the relation between the current and vector potential will be altered. To account for the effects of electron scattering, Pippard modified the coherence length factor in the exponent of(3.8) as 1!~0 --> (1/(0) + (lit) where t is the electron mean free path;' the coefficient in front of the integral was not altered. Eq. (3.8) then becomes
J(r)=---

3ne

47[~omc

..-

f[A(r')'R]R
R4

-R(.,'-+~)

'" Id'r.

1,

(3.l3)

In the limit ). > > (t, ~o) we may again take A(r) from under the integral sign; carrying out the integration we obtain
j(r) = -.. mcc;o

ne2

1 ·--A(r). 1 1

(3.14)

-+~o

For the case of a very dirty metal, where l < < ~o, j(r)
=--

ne? t me

~ C;o

A(r).

(3.15)

The effective penetration depth is then obtained from the expression (3.16)
If ;. < < (t, ~o) we continue to use Eq. (3.12).

The magnetic response of a superconductor depends on whether A ;;:; ~ or A .:S ~, as will be developed later." These regimes are designated in Table 3.l. A Type I superconductor displays a
2. Other expressions, such as ~ = ({ ';0)1/2, are sometimes used to estimate the effective coherence length. Such ambiguities are removed by the microscopic theory to be discussed in Part III.

Pippard's equation

11

Table 3.1 Regimes defininq Type J and


Type II superconductors?

Type I (or Pippard) Type II (or London) complete Meissner effect (flux exclusion) up to some critical field He' above which it becomes normal. The magnetic response of a Type II superconductor is more complex and will be developed later.

3. The precise criterion separating the regimes is A =

ufi.

.I--~-h-e-r-m-O-d-y-n-a-m-ic-s-o-f-a-Type I superconductor

In this section we consider the thermodynamics of a Type 1 superconductor (Gorter and Casimir 1934a,b); our discussion follows that of London (1950). We restrict the geometry of the superconductor to a form for which the external field, H, is not distorted by the presence of the superconductor (examples being an infinitely long cylinder with H parallel to the axis or a plane slab of infinite extent with H parallel to its surface). Far inside the superconductor (i.e., several London depths from the surface) the magnetic field essentially vanishes in the superconducting state and is equal to H in the normal state. In the thermodynamic identities that follow we identify this interior field as B, the flux density. The H field will be taken as the applied external field. The relation between Band H is shown in Fig. 4.l. We recall the thermodynamic identity for the response of a system in a magnetic field dIS = TdY

+ -H·dB,
4n

(4.1)

where Cf· is the energy density and Y is the entropy density. When T and B arc the independent variables we use the Helmholtz free energy density, Y = iff - TCf', and when T and H are the independent variables we usc the Gibbs free energy density, '§ = .'17 - (1/4n)B· H; taking the differential of these two quantities and using (4.1) yields 1 dY = - 9'dT+ -H·dB 4n and
d~1j= (4.2)

.'f'dT - -

4n

B· dH. properties:

(4.3)

A Type I superconductor B
=

displays the Mcissner-Ochsenfeld

0, H> He.

(4.4)

B=H,

(4.5)

He is called the thermodynamic critical field. Since H and Twill be our independent variables, we integrate (4.3) at constant T to obtain
<fJ(T, H)
=

,§(T,O)

Lf

B· dH;

12

Thermodynamics

of a Type I superconductor

13

/ / /

Figure 4.1 B vs H curve for a Type I superconductor.

therefore 1 ~fj(T,H) = 0'(T,O) - _(H2


Sn

- HD

(4.6)

or ~fj(T,H) Note that


qj

0'(T,O)

(4.7)

is continuous at the transition. We define a function


'fJo

==

'fJ(T,O)

+ -II~.
0

Sn

(4.8)

We may then write (H>HJ


(4.9)

and
'fJ(T,H)
= 'fJo - -H~

8n

(4.10) where qjN and qjs denote the normal and superconducting states, respectively. We may interpret as the Gibbs free energy of the normal metal at zero field (were it stable); hence the Gibbs free energy density of the superconducting state is lower than that of the normal state by (1/8n)H~(T); this quantity is referred to as the condensation energy. Since ~fj == .'F - (1/4n)H' B, §is(T,O) = qjo - (I/Sn)H~. The normal state Helmholtz free energy density is then
qjo

(4.ll) From Eq. (4.3),


.r;' = _ (~'fJ)

6T

;
H

(4.12)

14

Part I

Phenomenological theories

thus Y' --N-

aqjo

DT

(4.l3)

and
Js=
c:~ O,'lo --+-Hc 1
~,(f

aT

4n

('-H) aT
()

.
H

(4.14)

Note the entropy is discontinuous (when H =F 0):

across the transition and hence we have a first order transition

(4.l5 The heat of the transition is (4.16 this equation corresponds to the Clausius-Clapeyron heat (at constant H) is defined as equation of a (P, V,T) system. The specifn

(4.17 or

_ T[ ---H--+- 32Hc (OHc)2]


4n
c

cT2

aT

(4.l8

At 11 = 0, where the transition is second order, t;

/I,.«jHIT=T'=-4n

(311 aT

c)

T~T';

(4.15

this is sometimes called Rutgers' formula (Rutgers 1933).

.I---~-'h-e-in-t-e-r-m-e-d-ia-t-e-s-t-a-t-e--

If a superconducting body of arbitrary shape is placed in a magnetic field, the flux exclusion associated with the Meissner effect will in general distort the magnetic field. Exceptions are an infinite cylinder with the field parallel to the axis, or a sheet or half space with H 0 parallel to the plane of symmetry. For situations involving lower symmetry the local magnetic field can vary over the surface, being both higher and lower than the applied field, Ho. As a simple example consider the case of a spherical superconductor shown in Fig. 5.1. From magnetostatics the field willbe highest at the equator (on the cirele C in Fig. 5.1) where it is H = ~Ho- Hence flux enters the sample, not at the thermodynamic critical field, Hc' but at a value flo = ~Hc. For magnetic fields He> Ho > iHc the sample consists of alternating domains of normal metal and superconductor. A superconductor in such a regime is said to be in the intermediate state (Landau 1937).

Figure 5.1 The magnetic field distribution about a superconducting sphere of radius a. For an applied field Ho < ~Hc' there is a complete Meissner efTectand the field at the equator (at any point on circle C) is ~H 0; the field at the poles (Q, Q') is zero. For ~Hc < Ho < He' the sphere is in the intermediate state. 15

16

Part I

Phenomenological theories

Figure 5.2 Schematic of the magnetic field distribution in a slab in a perpendicular field. For all fields JJ 0 < ll ; the sample is in an intermediate state having a laminar structure of normal and superconducting domains.

We will limit our discussion to the simple case of a plane superconducting sheet with Ho parallel to the surface normal. From our earlier discussion we know that a superconductor cannot sustain a field component perpendicular to its surface. The field behavior is shown qualitatively in Fig. 5.2. It has the following features. (i) (ii) For a magnetic field 0 < Ho < He the sample consists of adjacent domains which are wholly superconducting (with no internal flux) or normal (with Hloeal = HcJ, In the interior of the superconductor and far from the surface, the domain walls are parallel to the applied field direction; the fraction of the cross section that is supercond ucting is fixed by Ho and H, such that the total flux through the sample is conserved normal cross section total cross section Hence the superconducting H0 He (5.1) (which vanishes, as it should, for

fraction is 1 - HolHe

Ho = Hcl. (iii) Near the surfaces the flux sheets 'flare out' (which reduces the field curvature, which would otherwise raise the local fields at the interface). Were the local field to remain fixed at He' to sustain the average flux, needle-like superconducting domains would have to encroach on the normal domains in the vicinity of the surface. Such needles do not occur since the field profile is controlled by a minimization of the surface energy (between supcrconducting and free surface or the normal regions, see next section), the superfluid kinetic energy, and the total magnetic field energy. A detailed treatment of the domain structure is mathematically complex and will not be dealt with here (see Landau, Lifshitz, and Pitaevskii (1984), deGcnncs (1966), London (1950)).

8f----swe have a phase boundary region

u-r-,.-a-c-e-e-n-e-r-g-Y-b-e-tw-e-e-na normal and a


superconducting metal

Consider a slab parallel to the x-y plane in a perpendicular perpendicular

magnetic

field parallel

to z. Assume occupying the

to the x axis with the superconductor model is: x H(r)]2 ] , H~(T)!8n,

x> O. The

total free energy, F, in the London F = sf


YJ

l I
o
dx

.'1's

H2(r),i2 + -- + ----'=:[V

Sn

Sn

(6.1) and the second energy

where s1 is the interface density, respectively.

area,:Fs

is the condensation field energy density

energy density,

and third terms are the magnetic we have (see Eqs. (4.6) and (4.7))

and the superfluid

electron-kinetic

At our phase boundary

in the intermediate

state, where H = He for x

< 0,
(6.2)

By definition

=F - -

eWI

,<, 0

4n

He' B(x)dx.

(6.3)

According to the discussion H(x) = HczeX /

at the beginning

of Sec. 4 we set B(x) = H(x); from Eq. (2.17) we have

AL•

Inserting

this form in (6.3) and using (6.1) we obtain

= -

Yc

"I' .7" S

d3"

+ ",yY, I
_4

(6.4) tension', given by

where y is the surface energy per unit area, i.e., the 'surface

(6.5) Note this surface energy is negative. This suggests the system can lower its energy by maximizing

17

18

Part I

Phenomenological

theories

the interfacial area (i.e., the system is unstable to the formation of multiple domains with associated interfaces), Type I superconductors (for geometries not possessing an intermediate state) display a single domain for H :::;; e; hence they must have a positive domain wall energy. Physically this H positive surface energy arises because superconductivity is destroyed over a region of order ~ perpendicular to the interface; i.e., we lose the condensation energy over a volume of order .#~, where ( is a coherence length. This is equivalent to a positive contribution to the surface energy of order H2
},=~_c.

Sn

(6.6)

In a Type I material ( > )'L and hence the positive contribution (6.6) outweighs the negative contribution (6.5) and the interface is stable. In Sec, 9.3 we will continue this discussion and derive an expression for the surface tension. For a Type II material the system does, in some sense, try to maximize the amount of internal interfacial area above some field (referred to as the lower critical field); however, it is subject to a constraint imposed by quantum mechanics, as we discuss in the next two sections.

8I---o-u-a-n-t-;z-e-d-v-o-r-t-ic-;-t-y---

The previous discussion of the surface energy of a normal metal/superconductor interface suggests that Type II materials, where ¢ ::S A are unstable to the formation of domain structures which in some way maximize the amount of interface area. Two possible domain geometries are: (i) an array of nested sheets (closed or open, depending on the geometry 1) and (ii) a twodimensional lattice of flux filaments. Calculations show the latter domain structure to be more stable. Since the filaments (are presumed to) admit flux into the interior of the superconductor, we envision them as having a normal core with a diameter of order ¢, outside of which supercurrents flowin a diameter of order }, which produce the internal field via Ampere's law. As a primitive model of a single flux filament we consider the extreme limit ¢ --> 0 for which the London approach should provide a good description. We recall Eq. (2.5) associated with our first derivation of the London equation

-a v
Dt

x (V x H)

+ 4nne -2-

me

H]

0;

(7.1)

we next integrate (7.1) over an area s~ intersecting the filament (for convenience we choose a plane perpendicular to its axis) and use Ampere's law (Maxwell's fourth equation) in the form
-;;-

ot

a flme

-2

ne

+H]

. d,xi

O.

(7.2)

In integrating (7.2) with respect to time we now allow the possibility of a nonzero constant of integration (since the flux filament phenomen violates the Meissner behavior); thus (7.3) Applying Stokes' law to the first term in (7.3) yields ne2 j'dt

mej1

+f

H·d.xi

4>.

(7.4)

Ifwe choose the contour to enclose a large area, we may expect the first term to be exponentially
1. For a superconducting slab we envision an array of interfaces parallel to the surface and for a cylindrical sample an array of coaxial cylinders. Other shapes would have more complex structures.

19

20

Part I

Phenomenological

theories

small (since the currents fall off exponentially with a characteristic length ),); we can then identif the constant of integration, ¢, as the total flux contained within the filament (most of which alsr falls inside a radius of order A). To gain further insight we substitute H = V x A into (7.4) and again apply Stokes' law tc obtain

f[:~j 1M
+A
C

= ¢;

(7.Sa

writing j

nev

ne(p/m), (7.Sa) becomes

l[p J + : A] . dt = : ¢.
C

(7.Sb

Wc identify the integrand as the canonical momentum associated with the motion of a charged particle in the Hamiltonian formulation of mechanics. F. London correctly concluded that superconductivity was a macroscopic quantum phenomenon, and guided by this insight he suggested that Eqs. (7.S) must conform with the Bohr-Sommerfeld quantization rule for the (quasiclassical) motion of an electron, i.e. (I e I/c)¢ = nh or ¢ = n(he/I e I), where n is an integer. However, this assumes the orbiting entities are single electrons; Ginzburg and Landau allowed for a marc general case where e -> e*; we then have (7.6a) where he

¢o=le*l·
From the Bardeen-Cooper-Schrieffer he (ECS) theory it is now known that e* = 2e; i.e.,

(7.6b)

¢o = 21 e 1 = 2.07
X

10-7 G crrr'

(7.6c)

which is referred to as the flux quantum. Hence flux enters a Type II superconductor as an array of quantized flux filaments; the lowest energy situation corresponds to singly quantizied (11 = 1) filaments eaeh carrying a flux quantum ¢o. In what immediately follows we adopt cylindrical coordinates (r, 0, z) and write the in-plane radius vector as r. Let us next examine Eqs. (7.S) for a contour of radius A > > r > > ¢; the amount of flux contained is then vanishingly small and the first term in the integrand of (7.Sb) dominates yielding the condition 2npr = h or p = h/r; BCS theory also dictates that the mass of the orbiting entity is m* = 2m; hence h~ v(r) =-{} 2mr (7.7)

where (j is an azimuthal unit vector. This velocity profile corresponds to the large r behavior of a vortex in a fluid, although with the vorticity quantized. One then refers to the filaments as quantized vortex lines or vortex lines for short.

Quantized vorticity

21

Vorticity involves

a nonvanishing region
-->

curl of the velocity;

i.e., V x v # 0; but the curl of v in is a it would be spread description out over a by adding

(7.7) vanishes for r # O. However,


located in an infinitesimal coherence length, ¢. In our ¢

the circulation near thc origin.

r ==

~v' dt # 0 and thus all the vorticity


an approximate

Physically

0 model we can obtain


equation

singular source term to the London

in the form (7.8)

where 6(2)(r)is a two-dimensional

() function

and as

z is a unit

vector (parallel

to the vortex axis). Eq.

(7.S)may be written in cylindrical

coordinates

).~ dH_) Hz - - -d (' r _r dr dr


The left-hand side of (7.9) is a special required singular behavior

21 = ¢o(Y(r).

(7.9)

case of Bessel's

equation/

and the solution

having

the

near r = 0 is (7.10)

where Ko is the zeroth

order modified

Bessel function and,

of imaginary

argument. tn(l/x);

From hence

Ampere's

law, j = (e/4n)V x H = - (c/4n)(dHz/dr)O

.io = (el 41rH¢o/2n).~r) = nehlanmr or v = hl2mr, in agreement with (7.7). For large x we use the form Ko(x) ~ (n/2x) 1 /2e -x or Hz ~ (¢o/2nJ.D(niJ2r)1/2e -riAr and the field drops off exponentially, as for small .x, Ko(x);:; The energy of a vortex line follows from Eqs. (2.8)-(2.11);

argued above.

E = 8~

where L is the length of the vortex line. We integrate E ).2 - = -'= H x (V x H)· dt

8n

f f

[H2(r)

+ ;_;:(V

x H)2]d2r,

(7.11) (7.11) by parts to obtain H' [H

+ -1

8n

+ ),tv

x (V x H)]d2r.

(7.12) side of (7.8). 0), which corc is is not H is in the

The quantity in square Were we to simply logarithmically finite everywhere, sufficiently accurate

brackets replace This

in the second term of (7.12) is equal to the left-hand integrate our we would model obtain H(r
-->

it by 6(2)(r) and indicates the energy. that

divergent.

simple

for the vortex problem

to evaluate thus eliminating

To avoid the divergence

we assume

b(2)(r) from the right-hand We must still account the line integral radial paths For small
I'

side of (7.8), which results

second term in (7.12) vanishing.

for the energy in the core of thc vortex, in (7.11) into three parts: a circle at a very due to the exponential fall-off of H(r)

however, which we do by separating at large r), two counter-traversing only nonvanishing contribution).

large radius (which makes a vanishingly

small contribution

from the outer circle to an inner circle of a very the inner circle (which makes the

small radius (which cancel each other) and finally a path around

the first term in (7.12) is (7.13)

2. Here we adopt the definitions of Abramowitz and Stegun (1970), pp. 374ff.

22

Part I

Phenomenological

theories

Figure 7.1 Structure of an isolated Abrikosov vortex line in a type II superconductor or (7.14 where we choose a radius r = ( for the (inner) line integral; this corresponds to the physicallj reasonable assumption that the field divergence is removed at the coherence length scale (as, more complete theory confirms). We expect that the l/r divergence of the superfluid velocitj ultimately destroys superconductivity in the vortex core; we may model this effect by assumim the density of supcrconducting electrons, n(r), approaches zero (sufficiently rapidly) as r --> O.Fig 7.1 shows the qualitative behavior of H(r) and Il/I(r) I near the center (core) of a vortex filamen where I l/I(r) I == [n(r)]t. In the next section, where we discuss the mixed state of a Type II superconductor, we wil require an expression for the interaction energy of two vortices. Returning to our London-lik: model, we generalize Eq. (7.8) to the case of two (parallel) vortices (7.15 Since this equation is linear the resulting magnetic field will be the sum of two terms having th same form as (7.10):

(7.H The total energy (which is quadratic in H) will involve three terms: two of these correspond to th 'self-energies' of the individual vortices (as given by (7.13)) and the third results from thei interaction, The interaction energy in the extreme London limit can be evaluated from the secor» term in (7.12) along with (7.10) (we assuoe the path of the first integral involves a single circle c very large radius encircling both vortices); the interaction energy is then given by

Ouantized vorticity

23

=2(~)2
4n)·L

Ko(lrl

~r21),
AL

(7.17)

where,due to our assumptions, we require 1r 1 - r21 ;<; ~. The sign of (7.17) is positive and hence the force (per unit length) between two vortices is repulsive. We may rewrite Eq. (7.17) in the form
E(l2) __ '" =~H(l2)

4n

'

(7.18)

where H(12) =0 ll~l)(r = r 2) = H~2)(r = r 1) with the latter given by (7.10); i.e., H(12) is the contributionto the field at one vortex resulting from the presence of another. Regarding ( - ¢oL/4n)z as a magneticmoment, p, we could write E(12) = - p" H. where H = H(12)Z.

v
8.1

Type II superconductivity

Magnetic fields slightly greater than

He1

Superconductors with ~ < )" termed Type II materials, behave differently in a magnetic field fron those with ~ > I" (Type I materials). Fig. S.1 shows the field dependence of the magnetization of a: ideal Type II superconductor (the sample geometry is assumed to be a plane slab or a cylinder tl avoid geometrically induced field inhomogeneities as noted earlier in our discussion of th intermediate state). For low fields the magnetization is - H/4n; i.e., the sample displays, Meissner-like behavior. However, at a field H = He1, called the lower criticalfield, flux abrupt! enters the sample (the susceptibility X = (dM/dH) --+ + x for an ideal sample at this field). Th magnetization increases (becomes less negative) continuously above Hel and reaches zero at ; field H = H c2 called the upper critical field. The regime Hel < H < Hc2 represents a nev thermodynamic superconducting state called the mixed state or Shubnikov phase (Shubnikov e al. 1937). We now show that the Meissner state of a Type II superconductor becomes unstable to tb entry of vortex filaments at a field which we identify with Hc1• For magnetic fields which are onl slightly above He! we may write the Gibbs free energy per unit volume as I#(H) = I#(H = 0)

+ nLEIL + - I E(I,JI
Li<i

..
-

B·H
--.

4n

(S.1

The second term represents the self-energy of the individual lines where nL is the number of line per em? with energy E/ L per em, the third term is the (repulsive) interaction energy between th: lines (and is summed over a unit of area), and the last term is the usual field term relating tlu Gibbs and Helmholtz free energies. In the presence of a uniform array of flux lines the magnetn induction is (S.2 Near He! (above which flux first enters), we will initially neglect the interaction term and, usiru (S.2), write (S.3 For H < 4nE/¢oL, 24 (S.3) is minimized by setting nL
=

O. However, for H > 4nE/4>oL Eq. (S.3

ype " euperconaucttvtty

-41TM

(a)

-41TM
Meissner Phase AbrikoSOlJ Mixed Phase
Normal

Phase

(b)

Figure 8.1 The reversible magnetization curves for a Type I (a) and a Type II (b) cylindrical superconductor. The magnetic field is applied parallel to the cylinder axis.

would suggest

that q} ex - nL; i.e., the system

can lower

its free energy

indefinitely

simply

by

creating more flux lines. It is clear that we can identify

He I as
(8Aa)

or
(8Ab) where we used (7.14) for the vortex self-energy in (SAb).1 The negative interaction (repulsion) divergence of CIJ for

H> Hel is eliminated


the minimum

if we include vortex-vortex to B of the quantity

effects; i.c., we must seek

with respect
!J.q}=-

4n

cI

1 4>0 -H+-z--K

2 2n}~

O}L

(d)] -

(8.5) of the vortex-vortex

The first two terms

are the same as in (S.3). The last term is the result

1. We can now make contact with the comment made at the end of Sec. 6. Taking the classical limit as h ---> 0, we see that the lower critical field would approach zero. At any finite field we would then have a divergent number of flux lines; i.e., the system would have a divergent internal interface area. Thus quantum mechanics imposes the constraint on the maximal amount of internal interfacial area.

26

Part I

Phenomenological

theories

0.10 0.05 0.00


!J.'§ -0.05

H/H cl = 1.1

(a)

1.2

-0.10 -0.15 -0.20 2

1.3 ,,_
3
4

dI", l

o
(b) -50 -100
I
\
\ \

H/Hcl
\

z:

8
->

!J.,§-150
-200 -250 -300 0.0

,,/\

-~
\
\ \ \

.>

9/

/'/-

-:

//

///

\_-

////10
0.4 diAL 0.6 0.8

0.2

Figure 8.2 Normalized Gibbs free energy density calculated from Eq. (8.5) as a function of the reduced vortex lattice spacing in the low applied field regime (a) and high applied field regime (b).

repulsion; the factor 1/2 assigns half of the interaction energy to each vortex (in Eq. (7.18)), z is the number of nearest neighbors in the vortex lattice (the exponential fall-off of Ko justifies including only nearest neighbor interactions at low fields where nL is small), and d (appearing in the argument of Ko) is the vortex-vortex spacing, which depends on the symmetry of the vortex lattice. Calculations show that a triangular (centered hexagonal) arrangement of lines has the lowest energy. For this lattice nL = 2/J3d2 = BNo or d2 = 24>o/J3B, and z = 6. On substituting these values of d and z in Eq. (8.5) we obtain !J.~tJ= !J.'§(B); this function is shown qualitatively in Fig. 8.2 and the minimum value of MtJ yields the magnetic induction, B, for a given external field, H. Carrying out the minimization process for each field H we can develop the function B = B(H) (or M = M(H) through B = H + 4n:M). (Further analysis of this model shows that X = (3M/aH)T does diverge as H --> He! from above.) The above analysis breaks down as nL increases (i.e., as B increases), and more powerful techniques are required. The simplest method involves the Ginzburg-Landau (G-L) theory, a discussion of which we begin in the next section. In discussing the thermodynamics of a Type II superconductor the concept of the ther-

Type" superconductivity

27

modynamic
He!

critical

field, He' is retained From the discussion

in terms of the condensation in Sec. 4 (where we showed

energy,

H~/Sn; however,
= 0) we

< He < Hc2'

(l/4n)Sg,B'dH

identified H;/Sn

as the condensation

energy. The fact that M has a different

field dependence in a

Type II superconductor

does not alter this identity.

Therefore
(S.6) compounds. This fact, coupled up to He2, makes with the Type II

f
Hc2

H'
c.

M(H)dH

= __

Sn

H~.

greatly exceeds He in certain observation

alloys and intermetallic

experimental

that a zero resistance in the manufacture

state often persists of magnets

materials of great importance

and related

technologies.?

8.2 In the region


He!

The region HC1 < < H < < HC2


<<
H

<<

He2

there

is a densely

packed ~

array

of vortices;

however,

the

spacing between vortices ( ~

nC 1/2), d, satisfies
London

the inequality equation

( < < A so as to guarantee


by the solution

the existence

of such a regime).

< < d < < A (here we assume that When d > > ~,H(r) is accurately given

of the inhomogeneous

(S.7)
where i runs over all vortices in the lattice. Assuming Fourier the vortices lie on a periodic

lattice:' H(r)

may be expanded in a two-dimensional H(r) = IHGe-iG,


c

series as (S.Sa)

where (S.Sb) here G denotes all vectors of the two-dimensional reciprocal lattice associated with the real space (S.Sa) into by eiG"', and

lattice of the vortex array and nL is the reciprocal integrating over d2r. and noting G'·ri is a multiple

of the area of the unit cell. Inserting of 2n we obtain

(8.7),using the fact that V x [V x H(r)] = - V2H(r) (since V' H(r) = 0), multiplying

,s¥' I (1 + A 2G2)H(A;G'
C

= ¢oNLz

2. In the presence of a transport current (as one has in a magnet) the nux lines arc subjected to a Lorentz force which would tend to make the flux lines move, which is a dissipative process. However, flux lines are usually pinned (immobilized) by various inhomogeneities present in the material (c.g., defects, grain boundaries, ctc.), 3. In an inhomogeneous superconductor, vortices may be attached to so-called pinning sites, resulting in deviations from periodicity. On the other hand, thermal agitation may cause the lattice to melt in a high temperature superconductor.

28

Part I

Phenomenological

theories

or (S.9 where sf is the total cross sectional area of the vortex lattice, IVL is the total numbers of lines ant n, == N Lied; henceforth we assume that H(r) and Hc are directed along the z axis and denote onltheir amplitudes, H(r) and Hc' The total energy of the vortex lattice follows from Eq. (7.11) as

Defining the energy density Iff'

= E/sf L

we obtain (S.10

Now the minimum nonzero value of I G I ~ 2n/d( ~ 2nn~!2) and with our assumption d < < ;. wr have ;.2G;;:'in > > 1; therefore we may approximate 1/(1 + }.2G2) by 1/).2G2. The sum in (S.10 depends on the particular lattice adopted by the vortices. However, following deGennes (1966) we will limit ourselves to a semiquantitative estimate and replace the sum by an integral; (S.11 From the above argument Gmill ~ 2n/d; on the other hand we expect Gmax to be cut off at the scale of an inverse coherence length and write Gmax = 2n/¢. We then obtain the total energy density 0 the la ttice as (S.12 here we used Eq. (S.4b) for He! and included a numerical factor f3 to offset partially the various approximations used to obtain (S.12). Matricon (see deGennes (1996)) finds f3 = 0.3S1 for a triangular lattice where d2 = 2¢o/J3B. We note in passing that by considering the changes in the free energy arising from small distortions of the lattice from its equilibrium form, we rna) calculate the magnetic contribution to the elastic constants of the lattice. We will return to this topic in Sec. 20.7. Ignoring the entropy contribution we may replace Iff by IF; using the definition '!J
= ff - -BH

4n

(8.13;

we may then calculate B = B(H) from the condition aG/3B = 0 which is appropriate for a constant-H environment. Carrying out the calculation (including the contribution from the implicit dependence of d on H) we obtain

Type" superconductivity

29

(S.14)

where

{J' == {Je-

I!2.

From the definition B


tn

+ 4nM

we have

He!

({J'~)c,

M=-

4;- ...

tn (

(l) .

(i\.15)

Sinced X. B-I!2 we see that M increases logarithmically (becomes less negative) with increasing magnetic field. This behavior agrees well with experimental data for materials with )_ > > ~.

8.3

Microscopic magnetic probes of the mixed state

Wewill now briefly discuss two experimental probes of the inhomogeneous magnetization in the mixedstate: neutron diffraction and nuclear magnetic resonance. Other probes which have been applied or proposed include: decorating the surface with magnetic atoms (the Bitter technique) and magnetic force microscopy. (For more coarse grained magnetic structures magneto-optic and scanning Hall probe microscopy are useful.) We begin with neutron diffraction. The neutron has a magnetic moment fln = (gn/2)(eh/ 2Mpc) where gn is the anomalous 9 factor of the neutron (gn/2 = - 1.91354) and eh/2Mpc is the nuclear magnet on with Mp the proton mass. The neutron interacts with a magnetic field via the usual spin Hamiltonian
f1

=-

fin·

H(r),

(S.16)
(J

where fin = fln!! where g is the vector Pauli matrix. The coherent scattering cross section, followsfrom the standard Born approximation expression

c-

wheref(q) is the scattering amplitude given by f(q) = -. 1 2nh

A1nf~·· r; H(r)e'Q"rd
3

(S.17)

here Mn is the neutron mass and q = k - k' with k and k' the wave vectors of the incident and scattered neutrons (I k I = I k' I for a diffraction experiment). Inserting Eq. (S.Sa), with H G given by (8.9),into (S.17) we have (S.lS)

30

Part I

Phenomenological theories

where V is the sample volume and the Kronecker delta function reflects the usual Ewald diffraction condition, q = G. Writing B = l1L<:Po, <:Po = 2nnc/2e and Jin ~ (gn/2)(elJ/2Mnc) we have
(g.19)

Note that as discussed above ;c2G2 > > 1; hence 6(q) = If(q) 12 o: (AG)-4. For this reason it is very difficult to observe more than the lowest order peak. Experiments were first performed on Nb by Cribier et al. (1964). Since high T; and heavy fermion materials have anomalously large London depths neutron diffraction has not been a useful probe of the vortex lattice in these materials. Wc now discuss the expected behavior of the line width, ~w, measured in an nmr experiment. For many purposes the line width may be inferred from the root mean square deviation of the magnetic field from its average value"; i.e., (S.20) and we take ~w = y~H, where y is the gyromagnetic ratio of the nuclei which are in resonance. Here the bar implies an average over the sample volume. For simplicity we assume the external (applied) magnetic field is strictly uniform and that the sample has the shape of a long cylinder parallel to the applied field so that demagnetization simply B
= I1r<:Po;

effects may be ignored. Now H(r) in (S.20) is

H2(r)

is calculated as follows using (S.Sa) and (S.9):

where "I is the area of the sample perpendicular


--2

to the magnetic field. Separating ofTthe G


2 2

term (which cancels the [H(r)] term in (S.20)), restricting to the limit A G replacing the sum by an integral we obtain

> > 1, and again

The integral is convergent at the upper limit so we extend Gmax to cc and we again take Gillin ~ 2n/d (valid for a square lattice) yielding (S.23) Note that our result is independent of magnetic field and within a factor of order unity is equal to H c L' Therefore nmr is readily observable in extreme Type II materials containing nuclei with sufficiently large moments. We emphasize that this line width is not to be interpreted as the reciprocal of a magnetic relaxation time; it is referred to as an inhomogeneous broadening. The relaxation of the magnetization will be discussed briefly in Subscc. 34.2.
4, For a discussion of moments in nmr see Abragam (1961),p. 106.

.f---~-'h-e-G-in-Z-b-U-r-g---L-a-n-d-a-u-theory

9.1

Basic equations

In 1937 Landau developed a model to describe second order phase transitions (those involving no latent heat). Central to this theory was the introduction of the concept of an order parameter. The order parameter is an appropriate quantity which vanishes in the high temperature phase (T> Te) but which acquires a nonzero value below the transition (T < Te)' The identification of the order parameter is often obvious from the nature of the second order transition. Thus for the ferromagnetic transition it is natural to identify the spontaneous magnetization, M, as the order parameter. The ferromagnet brings out a fundamental property of second order phase transitions: the development of a spontaneous magnetization is accompanied by a reduction in the symmetry of the system. Thus if the material is iron, where the high temperature (nonmagnetic) phase has a body centered cubic (bee) structure, on passing through the transition (Curie) temperature (Tc = 1043 K) the material chooses one of several symmetry-equivalent crystal axes along which the magnetization develops; the choosing of one among several directions (for iron the directions ofeasy magnetization are the cube axes and there are six such directions, ± x, ± y, ± z) lowers the symmetry of the crystal. The system is then said to have 'spontaneously broken symmetry'. It is a general property of second order phase transitions that the low temperature phase always has a lower symmetry. Since the order parameter evolves continuously from zero below Te, it is natural to expand the free energy as a power series in the order parameter. The free energy is a scalar but the order parameter may be a higher-dimensional object (e.g., a vector, tensor, complex numher). For our example of the ferromagnet, the order parameter is the magnetization, M, a vector. Thus in making the expansion of F(M) we can admit only symmetry-invariant combinations of the components, Mi; since the magnetization may develop along any ofthe (easy) crystal axes the free energy expansion must preserve the full symmetry of the high temperature phase. For a cubic ferromagnet the expansion of the free energy satisfying the above requirements has the form F(M, T)
=

F(O, T) 1

+ rx(M; + M; + M;) + ~f:il(M; + M; + M;)2


?

+ 2f:i2(MxMy + MyMz + MzMxl-·

(9.1)

Expression (9.1) is invariant under all the symmetry operations of a cube. However, to simplify 31

32

Part I

Phenomenological theories

F-F

F-F

0:

<0

~~o\~
(a) M (b)

LLM
···~~a2/2~

Figure 9.1 G L free energy functions for T> T, (ex> 0) and [or T < T, (et.< 0).

the discussion consider a (hypothetical) isotropic ferromagnetic liquid. The only rotationally invariant scalar that can be formed is then M· M and (9.1) simplifies to
(9.2)

Limiting the expansion to fourth order will lead to an adequate description of ferromagnetism. Thermodynamic stability requires /J > 0 (otherwise the system would seek a divergingly large magnetization). However, (f. may have either sign. If o. > 0 the minimum of(9.2) occurs at M = 0; if :x < 0 the minimum is for M #. O. These two situations are shown in Fig. 9.1. Hence we can model a second order phase transition simply by arranging for the sign of (f. to change at T; which is easily done by writing
(9.3)

it is sufficient to regard

fJ as a constant. Setting iJF/aM = 0 yields


(9.4)

This equation has the two solutions discussed above: ex> 0, ex < 0, In terms of Eq. (9.3) we have
1\.,f2 =

M=O

(T> Tel;
(T

(9.Sa)
(9.Sb)

<

Te).

(T> TJ;
(T< TJ.

(9.6a) (9.6b)

It is important to recognize that the ferromagnetic state is degenerate. For the case of the crystal it is degenerate with respect to the number of independent crystal axes (or directions) the magnetization might orient along (six for our case of bee iron with easy (100) axes). For our ferromagnetic liquid the ordered state is continuously degenerate: M can point in any direction. There is an important aspect of the second order phase transition which is neglected in the above model: fluctuations. All thermodynamic systems undergo fluctuations. As an example the density of a liquid may fluctuate by an amount o p. However, this is at the expense of an increase in the local pressure, ()P, governed by the bulk modules. This is a general feature: the 'restoring force' (pressure in our example) is linear in the 'displacement' (here the density). For a system

The Ginzburg-Landau

theory

33

whichundergoes a second order phase transition the order parameter becomes a thermodynamic variable below the transition. For our magnetic system the generalized restoring force accompanying a change in the magnetization 3M follows from expanding the free energy about its equilibrium value, M, However, as the transition is approached the linear restoring force goes to zero:just at the transition the restoring force is proportional to (()M)3. We see that fluctuations can have a profound effect on the system near its transition temperature. The Landau model therefore assumes that the temperature is far enough from the transition temperature that fluctuation effects may be ignored. This is referred to as the mean field model. We now take up the case of the supercond uctor. Our earlier discussion has brough t out the idea that superconductivity is some kind of macroscopic quantum state. Ginzburg and Landau built this idea into the Landau second order phase transition theory by assuming the existence of a macroscopic 'wave function', 1/1, which they took as the order parameter associated with superconductivity. Since wavcfunctions can be complex, only the form 1/11/1* may enter the free energy expansion; we therefore write (9.7) The minimization proceeds exactly as above, i.e.,

11/11

= 0,

(9.8a) {1
-

II/f 1= [

a(Tc

T)]1!2

(9.8b)

To describe situations where the superconducting state is inhomogenous we must generalize (9.7). F is then interpreted as a free energy density, .?(r), and we write F = Fo

l?(r)d3r

=Fo+

fd3r[xll/I(rW+~fill/l(r)14J

(9.9)

F is now the total free energy. Eq. (9.9) in its present form does not model the increase in energy

associated with a spatial distortion of the order parameter, i.e., effects associated with a coherence Iength.Z. To account for such effects Ginzburg and Landau added a 'gradient energy' term to (9.9) of the form (9.10) with m* as a parameter; the choice of the coefficient fj2/2m* makes (9.9) mimic the quantum mechanical kinetic energy (introduced earlier in Eq. (2.10)). Ginzburg and Landau assumed that if(9.10) was to be regarded as the kinetic energy contribution to the Hamiltonian density of the superconducting electrons, then (as in Hamiltonian mechanics) the interaction of the electrons with an electromagnetic field would be accomplished by the Hamiltonian prescription
e* p-->p--A c

34

Part I

Phenomenological theories

or, since p ---> (~) V in quantum ie*

mechanics,

V--->V--k the use of e* aIlows theory). Combining

he

'
entities to carry a different charge (e*

(9.11J

the superconducting the above we have

2e in BCS

FG
Finally

1i2 _

2m*

/ [ V - - A(r) ] ljI(r) /2 d3r. ie* he


of the magnetic field to the energy density

(9.12)

we must add the contribution

(2.9) Combining the above we have

F=F

+ fd
gn

3r

{x

IIjI(r) 12 +

!fi 1~J(r) 14 + 2rn* 2


112

[V _ ie* A(r)] ~J(r) /2

he

+ ~H2(r)l.
The minimization turn involves

5
out using the methods of the calculus ljI(r), 1jI*(r), and H(r)( to 1jI*(r) yields ie* V - - A(r)

(9.13)
of variations

of (9.13) must be carried involving functions

since F is a functional Minimizing

the free energy density ff(ljI(r),

I/J*(r), VIjI(r), V~J*(r), H(r)) which in V x A(r)).

the unknown

(9.13) with respect

()F =

f{ f
d 31'

2m*

1z2 [

lie

]2

ljI(r) +xljl(r)

+ fJll/J(r)

121j1(r) bljl*(r)

d2r'2rn*

lil [ V - --:-A(r) ] ljI(r)bljl*(r) ie* he


variable, yields the complex

(9.14) conjugate of

(variation

with respect

to 1jI, which is an independent

(9.14)). To minimize

iF we set the integrand

of the first part of (9.14) to zero; this yields the first

G-L equation
- -* 2m

112 [

ie*]2 V - --:- A(r)

he

ljI(r)

+ cxljl(r) + Ii IIjI(r) 121j1(r) = O.

(9.15) to

The surface term (which was generated establish certain boundary of(9.13) conditions with respect Variation v006b;011 provided we identify j as

by an integration to A (with H

by parts) can be used (with caution) later. law

and will be discussed

V x A(r)) yields Ampere's

V x H(r) = -j(r) c

4n

(9.16)

The Ginzburg-Landau theory

35

j(r)

e* = -*

~n

{[ 1jJ*(r)

e*] - ihV - - A(r) ljJ(r)

+ ljJ(r)

+ ihV -

e* - A(r) 1jJ*(r) , c (9.17a)

or equivalently, j(r)
= --

- ie*h 2m*

[~J*(r)VIjJ(r)

ljJ(r)VIjJ*(r)]- --IIjJ(r)
m*c

e*2

12 A(r).

(9.17b)

Eq. (9.17)is the second G-L equation; we note that (9.17) is the same as the expression for the current density in quantum mechanics. Note the current density satisfies the equationj(r) = e(I)F / bA(r));i.e., it is the variable conjugate to A(r).

9.2

Gauge invariance

Thesimplest solution of (9.15) is for the case of a uniform superconductor, IjJ #. ljJ(r), with A = 0, as given earlier in Eqs. (9.8). However, (9.15) possesses a continuum of other solutions having the same free energy, which we now show. As we can with any complex function, we write ~(r)= a(r)ei<l>(r) whereatr) and cJJ(r) rc the position-dependent amplitude and phase. respectively. a Let us examine a class of solutions which satisfy the complex equation
( ie* V - -A(r) he ) ljJ(r)

0,

(9.18)

whichis equivalent to the two real equations Va(r) and [ VcJJ - e* A]


he

=0

(9.19a)

= O.

(9.19b)

From (9.19a) we see that the only allowed solutions of(9.18) involve a constant amplitude, ,,; Eq. (9.19b),on the other hand, has infinitely many solutions involving a vector potential and a position-dependent phase (which does not affect the free energy) related by A = -VcJJ.
e* he

(9.20)

Any vector potential satisfying (9.20) results in a uniform free energy and (on substituting (9.18) into (9.17)) a vanishing current density (note that H = V x A = 0 for all A of the form (9.20)). The above exercise shows that the symmetry broken in superconductivity is qauqe symmetry, or equivalently, phase symmetry. Superconductors having different phase functions, (D(r),are in a real sense physically distinct/ this arbitrariness of the phase is the analogue for a supercon1. Strictly speaking we cannot determine the absolute phase of a superconductor, but in our later discussion of the Josephson effects we will show that phase differences can be measured.

36

Part I

Phenomenological

theories

ductor of the property that the magnetization ferromagnet.

may point in any direction in an isotropic (liquid

9.3

Boundaries and boundary conditions

We first examine a simple case involving an inhomogeneous order parameter generated by the presence of a boundary, in the absence of a magnetic field. Assume we have a superconducting half space occupying the region x > O. We further assume that the order parameter is driven tc zero at this interface. Experimentally this can be accomplished by coating the surface of the superconductor with a film of ferromagnetic material.' We then seek a solution to the onedimensional G-L equation
(9.21;

Noting o: is negative in the superconducting as


[2= __

state (rx =

- I xl), defining the G-L coherence length

- - 2rn* I o: I'
=

li2 -

(9.22)

and writing

(fJ/I o: l)tjJ2

I?

we may rewrite (9.21) as


(9.23)

- ~2r -f + f3 = O.
Multiplying by f' we may rewrite (9.23) as

(9.24)

hence the quantity in square brackets must be a constant. Far from the boundaryJ' [2 = 1 (equivalent to tjJ2 = I xl IfJ); then (9.24) becomes

0 and

(9.25)

which has the solution(

= tanh(x/J2~)

or
(9.26)

From (9.26) we see that ~ is a measure of distance over which the order parameter responds to a perturbation. Since x = a(T - Tel we have dT)=

"

(li2 --

2rn*aTc

)1,2 ( 1--T)-1/2
t;

(9.27)

2. Paramagnetic impurities (those bearing a spin in a host material) or interfaces with a ferromagnetic metal strongly depress superconductivity. A normal metal interface has a much smaller effectand an insulator or vacuum has a negligible effect for most purposes.

I ne l2mzburg-LanClau theory

37

We see that the G L coherence property of the coherence We now examine recognize the remainder n = IIjI IZ; this supports superconducting length

length

diverges

as 1/(1 order

T/TjiZ;

this divergence (although

is a general the exponent we

at all second In a limit equation,

phase transitions

differs in general from this 'mean field' value of 1/2 close to TJ Eq. (9.17b). as the London the identification where the first two terms wave function are negligible, associated as written caleulate in the form (3.5), provided the London penetration we identify with the

ofljl as a (condcnsate)

electrons.

We can immediately

depth as (9.2Sa)

or (9.28b) Comparing (1 Eqs. (9.27) and (9.2S) we see that


1i2.

in the G-L

theory

),1.

and

~ both

diverge

as

T/TJ-

Their ratio, called the G L parameter,

is therefore

a constant

which we write as (9.29)

Let us now return phase boundary given by

to the discussion

of the surface

tension

of a normal-superconductor is

begun in Sec. 6. From Eq. (6.3) the total Gibbs free energy, which is a constant,

oW

+
-'"0

dx [§(X) - __1__ He . B(X)J;

4n

(9.30) and the integral model, must now be the superconnominally, at of

here .91 is the interface ducting properties x = 0). Inserting

area, :1F is the G-L

free energy

density

extended into the region x > 0 since, due to the gradient do not turn on abruptly (9.13) into (9.30), employing

term in the G-L

at the interface equation

(which we still locate,

a gauge where A

A(x)z, and using the definition

i given by (6.4), we obtain the one-dimensional


}'=

+
_ <CD

dx

[rx2

17 -+-

1 dljl(x) 12

2f3

2m*

~dx

e*2 +--.zA1(x)lljJ(xW 2m*c


B(X)J
_._ •

+ rx IIjI(xW

1 BZ(x) H + - f311j1(x) 14 + ~. __ ~e
2 Sn

4n

(9.31) (9.15) allows us to

The vanishing of the cross term iA' V in both (9.31) and the first G-L equation choose IjI real; it then follows from (9.17) that

(9.32a) and

i, = i, = o.

(9.32b)

38

Part I

Phenomenological

theories

To compute (we must simultaneously solve the first and second G-L equations, (9.15) and (9.16), for ~J(x) and B(x)( = - (dA(x)/dx)), subject to boundary conditions which will be specified shortly, and insert the results in (9.31). To eliminate various parameters in the subsequent calculations we rewrite all equations in terms of the scaled variables: (9.33) In terms of these variables the first and second G-L equations become (where we now drop the bars over the variables) (9.34) and (9.35) The solutions of (9.34) and (9.35) for arbitrary boundary conditions are
~J
I(

must be obtained numerically. The appropriate

= 0, B

= A'

= 1 at

x = - ex)

(9.36a)

and

t/J

1, A'

0 at x =

ex;

(9.36b

(where x > 0 is nominally the superconducting side). 3 The behavior of B(x) and tjJ(x) in the small k regime where the field varies more rapidly than the order parameter is shown in Fig. 9.2. Tc obtain a first integral ofthcse equations we multiply (9.34) by tjJ', which leads immediately to 1d --(~J
1(2

dx

'2 )=-d

dx

[(1-A 2

2 -1 ) tjJ2 +-tjJ 4] 1
2

-tjJ

2-d

dx

(A2) -, 2

(9.37a

and multiply (9.34) by A', which yields ~(A'2) dx ,2 _ tjJ2~(A2) dx, 2 (9.37b

Combining (9.37a) and (9.37b) we have for our first integral (9.38 the value of the constant was fixed by the boundary conditions at either + 00 or - o:». Usin: (X2 !2~ = H:/8n and the scaled variables of Eq. (9.33) we may rewrite Eq. (9.31) as
(=--

) LHZc
Sn

3. From the structure of Eqs, (9.34) and (9.35) it follows from the boundary conditions (9.36a) and (9,36b) that tit' = 0 at x = ± 00. Our restriction to real t/I requires that the constant A (which is equivalent to a phase <1» vanishes at x = + Xl.

The Ginzburg-Landau

theory

39

r-----s--~
K

ll> I

Figure 9.2 Schematic diagram of the variation of Band 1// in a domain wall. The case /{ < < 1 refers to a Type I superconductor with a positive surface energy; the case /{ > > I refers to a Type II superconductor with negative surface energy.

Combining this expression with Eq, (9.38) we have


(= ~

) H2 4n

f [2
"%

_1jJ'2
K2

A'(A'

-1) dx;

(9.39)

givenA'(x)( = - B(x)) and tj/(x) we can compute (. Returning to Eq. (9.38) we verify that in the limit A' ofEq. (9.25);

0 we recover the dimensionless form

(9.40) Thesolution of this equation is


tj/(x) = tanh-~,------'-K(X -

xo)

)2

(9,41)

(whichis the analogue of Eq. (9.26)); here Xo is the nominal position of the boundary. The characteristic length scale of the order parameter variation associated with this normal/superconductor phase boundary is ~ (in un scaled units), as discussed earlier in connection with Eq. (9.26). his form docs not satisfy the boundary conditions (9.36), as expected, since a stable phase T boundary in an unbounded superconductor can exist only in the presence of a field, but rather satisfiesthe boundary conditions IjJ = + 1 or - 1 as x ---> + cc, However, assuming the presence ofan order parameter quenching mechanism at Xo = 0 (e.g., a thin ferromagnetic plane embedded in an otherwise homogeneous superconductor) we may evaluate the associated surface tensionby substituting (9,41) into (9.39); carrying out the integration (with A = 0) we obtain (9,42) From the definition of K given in (9.29) we see that :' ~ ~(H; /8n) as anticipated earlier in Eq. (6.6). When A is nonzero the characteristic length scale of a field variation may be estimated from Eq. (9.35). n a Type I material at a vacuum/superconductor I interface IjJ may be regarded as constant overdistances where the field varies. Since the length scale of Eq, (9.33) is unity this corresponds to the field varying over a distance AL (the London depth) in unsealed units. However, near a

40

Part I

Phenomenological theories

normal-metal-superconductor phase boundary the field variation occurs in a region where I/! is small. If we seek an approximate solution to (9.35) as an exponentially decaying form, A .~ e~X!,j where () is a characteristic length, then l/(F ;::::II/! 12. From an expansion of Eq, (9,41) about x = xo, I/! ~ K(X - xo). The average value of 1/!2 in a region of width b would be of order (K3f Combining these forms we see that the field decay would be governed by (K6)2 ;:::: l/b2 or () .~ K~ 1/2. This would result in a (negative) surface tension contribution of order - Kl/2y which, although less than the expression (9,42) in a Type I material, decreases slowly with K and hence limits the accuracy of this expression. With increasing K the surface tension continues to decrease and further analysis (see Lifshitz and Pitaevskii (1980), Sec. 46) shows that it passes through zero for K = 1//2. We end this section with a discussion of the boundary conditions at an interface between two dissimilar materials at least one of which is superconducting. A boundary condition which is appropriate for the case when no current flows parallel to the surface normal, ii, is (see Fig. 9.3(a))4 e* [ - ihV - ~A C

. iiI/!

1 = - -I/f,

(9,43)

as is easily verified by substituting in Eq. (9.17a); here b is a parameter having the units of length. The superconductor insulator case corresponds to the limit b - co, Lastly we consider contact between two superconductors in the presence of a current normal to the boundary. In the most general case the boundary, which we will locate at x = 0, may have properties different from both superconductors, an example being a thin layer of a third (normal) metal, which is referred to as an SNS (supcrconductor-normal-superconductor)junction. The most general (Cauchy) boundary conditions connecting the two sides of the junction, denoted 1 (x < 0) and 2 (x> 0), have the form (deGennes 1964)5 (9,44a) and (9,44b) If we choose a single gauge for regions 1 and 2 the vector potential will also be continuous, The coefficients Mij are not independent but arc constrained by the requirement of current conservation through the boundary: (9.45) as a result we may choose the Mij to be real and they must satisfy the condition MllM22 - M12M21

1.

(9,46)

The resulting expression for the current through the boundary.j., is


4. We emphasize that Eq. (9.43) is a macroscopic boundary condition. The behavior of the microscopic order parameter may differ substantially near the interface. S. A boundary condition of this type will be discussed from the microscopic point of view in Part III, Sec. 42.

The Ginzburg-Landau

theory

41

t \1If(x)\

x
(a)

hll}j
o
x
(b)

Figure 9.3 Schematic of the order parameter behavior: (a) near an NS boundary; (b) in an SNS junction.

-* -11/1
m JWl2

e*h

1(0)

111/12(0) I sin[(f)2(0)

- (f)1(0)].

(9.47)

The maximum value of this current is

Jm"x

e*/i
=~

11/11(0) 111/12(0) I·

(9.48) then we may

1Vl12

If jmax is much less than the bulk critical currents in the two superconductors

approximate the I/Ij(x) by the A = 0 forms analogous to (9.26) (9.49a) and (9.49b) here Xl and X2 are free parameters and ~1.2 are the G-L coherence lengths in the two media. From Eqs. (9.49) we have immediately (9.50a) and

(9.50b)

42

Part I

Phenomenological

theories

By substituting these quantities into the A = 0 limit of Eqs. (9.44) we may numerically solve for the parameters XI and Xz, assuming the coupling coefficients ]VIi} are known. The behavior of II/I;(x) I is shown schematically for an SNS junction in Fig. 9.3(b). We will encounter Eq. (9.47) in Sec, 15 when we discuss the Josephson effects at an SIS junction (where I designates a thin insulating barrier). If the coupling is weak, as in the case of an SIS junction, we have M11 = M 22 ~ 0 and M 21 = - I/M 12' In this case the values ofll/l I (0) I and II/f 2(0) I are only slightly shifted from their equilibrium bulk values, 11/1I ( - oc) I and 11/12( + oc) I·

Of--~-'h-e-u-p-p-e-r-c-rl-·t-ic-a-'-~-ie-'-d-o-fa Type II superconductor

Asthe magnetic field is raised above He1, and the density of flux lines nL increases, a point is eventually reached where the distance between flux lines becomes of the order of the vortex core diameter; i.e. 1/2 ~ ~. One would then expect a transition to the normal state, and the field at whichthis occurs, called the upper critical field, is designated He2• In the region just below H c2 the superconducting order parameter must be small and this is the regime where the G-L approach should provide a good description. We may calculate Hc2 by linearizing the first G-L equation, since the order parameter is vanishingly small just at Hc2' Eq. (9.15) then becomes

ne

_ -h

2 (

2m*

ie*)2 V_- A he

,I,

'I'

+ (X'/'

'I'

o·,

(10.1)

this equation is the same as the Schrodinger equation for a particle with energy, _ a, mass, m", and charge, e*, in a magnetic field, H, with an associated vector potential, A. We will assume a uniformfield Ho II i. The solution of(10.1) is easiest in the so-called Landau gauge where we write thevector poten tial as
A
=

HoYx;

(10.2)

theresulting G-L equation is then (10.3) We may separate the variables in (10.3) by writing IjJn.kz.kJX,y,Z) = e;kzz+;kyYun(x) (we do not normalize IjJ at this point as this property follows only from solving the nonlinear G-L equations, whichwe address in Sec. 14). Inserting this wavefunction in (10.3) we obtain: [
_ -_

112 d --2 2m* dx

1 + - m*w~(x 2

XO)2

unix)

= GnUn(X),

(lOA)

where Xo = (l1e/e*Ho)ky = (¢0/2nHo)ky, and f,n = - rx- (h2k;/2m*). Eq. (lOA) is the Schrodinger equation of a harmonic oscillator with frequency we = I e* I H o/m*e (the cyclotron frequency of the particle) with energies
(10.5)

43

44

Part I

Phenomenological

theories

having its origin at the point xo' The full (unnormalized) wavefunctions are
V' n.k;:,kv·

./,

(x ,~,V z)

2 e''(k )"y. +k),z e " (X-Xo~!)'2 "HH

n'

(x - xo)
--all

(10.6)

where afr == h/m*wc = hc/e*Ho = ¢o/2nHo· Only the smallest eigenvalue, n = 0, k; = 0 solution (corresponding to the highest field at which superconductivity can nucleate in the interior of a large sample) is meaningful since our linearized theory is valid only as a description of the onset of superconductivity. Hence (10.7a) or
H ----

c2 -

2m*ca he"
-T)

=--a(T

2m*c ne*

(10.7b)

or, as it is more commonly written,


H -~.
cZ -

2n~2' Eq. (1O.7c)may also be rewritten as


aHel = ~.From

(10.7c)

(1O.7b)we see that = B/4>o, we see that (10.7c) is in accord with the order-of-magnitude estimate of Hcz made at the beginning of this section. However, we are then struck with the fact that in our particle analogy, we have considered only a single quantum state (n = 0, the 'ground state'), and yet the material contains the largest possible number of flux lines. The resolution to this apparent contradiction is that our n = 0 quantum state is highly degenerate (as evidenced by the fact that the quantum number k; does not affect the energy). The relation between this high degeneracy and the high density of flux lines will be dealt with in Sec. 14, where we discuss the region of fields just below Hcz.
aH,

in terms of the length

lld goes to zero at

T; and increases linearly below that temperature. Recalling nL

.I--~-'h-e-a-n-i-s-o-t-ro-p-ic----superconductor

Anisotropy may be incorporated in the phenomenological G-L theory simply by introducing an 'effective-mass tensor' into the 'kinetic energy' term in Eq. (9.15) in the form [ - in

(v -

i:*

A)

J~ J[- (v [~* in 0 0

i;:

A)] ~ +

a~

+P I~

12~ = 0,
(11.1)

wherel/m* is an effective-mass tensor. For a system which may be regarded as uniaxial, the reciprocal effective-mass tensor may bewritten as

[~*lj

mx 0 0

0 mx 0

(11.2)

mz wherethe z axis is the symmetry direction. The inverse of Eq. (11.2) yields the mass tensor m*. We now obtain an expression for the upper critical field Hc2(T). Near the field-dependent transition temperature the order parameter is vanishingly small and we may again neglect the nonlinear term. Our equation is then formally identical with the Schrodinger equation of a particlewith charge e* and an anisotropic mass tensor m* in a uniform magnetic field Ho and the 'energy levels' have the harmonic oscillator form
a

(n

+ 1/2)nwc(O),

( 11.3)

wherewJ8) is the angle-dependent cyclotron frequency which is encountered in the effective-mass theory of cyclotron resonance in semiconductors. The latter can be worked out classically simply by solving Newton's equations of motion for our 'particle' moving under the influence of the Lorentz force:

m*'v = -v
c

e*

x Ho;

(11.4)

here v is the velocity. The solution of this equation involves elliptical orbits which are traversed at a frequency
(11.5)

45

46

Part I

Phenomenological

theories

here 0 measures the angle of the magnetic field from the z axis. The solution with the lowest free energy corresponds to n = 0 in Eq. (11.3) and the upper critical field is thus given by Hd8)
= .

2ea(Tc
2

T)
!

ne*[slll 8jrnxrnz

2 + cos 2 8/mJ 12'

(11.6)

If we define two coherence lengths,

and

and recall the 'flux quantum'

1>0 =

he/ I e*

I, then Eq. (11.6) may be written as


(l1.7a)

in particular, for fields parallel and perpendicular to the symmetry plane of the material we have, respectively, (11.8a) and
HCH

=~.

1>0

2n<,x

(11.8b)

The angular dependence of the upper critical field is thus characterized by the two parameters C and ((Lawrence and Doniach 1971). Note that (11.8a) and (11.8b) both lead to a linear behavior of the upper critical field in T; - T. In terms of the definitions (11.8a) and (11.8b) we may rewrite Eq. (l1.7a) as H~2(O)
2

HcH

cos () +
2
LJ

H~2(8).
2

Sill

() -

LJ

(l1.7b)

He211

which is the equation of an ellipse. Eqs. (9.17) can be generalized to the case of an anisotropic superconductor by replacing rn* -1 by Ijm* from Eq. (11.2); an anisotropic London penetration depth then results.

.f----~-h-i-n-s-u-p-e-r-c-o-n-d-u-c-t-in-g-slabs

Consider an isolated superconducting slab which has a thickness d. As d is reduced, the effects of boundary scattering and reduced dimensionality usually cause To to fall (and ultimately to change the character of the transition); however, we shall here regard To and d as independent parameters. We consider the case of a parallel magnetic field and work in the gauge A = - H oyz. We seek a solution ofEq. (10.1) of the form

which yields the differential equation (12.1) where OJe = I e* I Ho/m*c and Zo == enky/e* Ho. The lowest (most negative) eigenvalue corresponds to k; = O.In a bulk sample, Zo may have any value since it simply refers to a change in the position ofthe origin. For a slab, we must solve Eq. (12.1) subject to boundary conditions at the surface of the slab. If we require the surface term to vanish in Eq. (9.14) we are led to the boundary condition ( ie*) V - --:- A
he

t/J. ii = 0,

(12.2)

where fi is a unit vector normal to the surface. This is not a general boundary condition; it turns out to be correct only for a smooth nonmagnetic superconductor-insulator or the superconductor-vacuum interface. If we place the origin in the center of the slab, Eq. (12.2) requires that
ou(z) -

au

I
z=±dj2

=0

We shall not seek an exact solution to Eq. (12.1) subject to this boundary condition. However, for d ::S ~, we may regard the second term as a perturbation. The zero order solution satisfying the boundary condition is then u = const. With this wavefunction and our perturbation, we have (12.3) where we have set
Zo

0 (which minimizes ex). his leads immediately to T (12.4) 47

48

Part I

Phenomenological

theories

The behavior ofEq. (12.4) is radically different from that of Eq. (11.8aJ. The reduced dimensionality has the effect of replacing ~z by d/(12)l!2. Since only one factor of'; enters Eq. (12.4), the temperature dependence of He21i is altered from Hc2 'I «: T; - T to He21i C£ iT; - T)112. We can refer to these two behaviors as 3D and 20, respectively. Let us now consider the case where the field is at an arbitrary angle with respect to the thin slab. We take the vector potential in the form
A = y(xcos8

- zsinO)Ho.

(12.5)

Writing

ljJ(r) = u(x,z)ei\y,

the linearized G-L equation becomes (12.6)

.itu where

+ au = 0,

(12.7) with (12.8a) and


.it' = -112

2m* oz

e2

+ - m*(j)~(z
2

Zo)2 sin? 0

+ m*(!)~x(z

zo)sin Ocos 8;

(12.8b)

here we have defined

Continuing our analogy to charged particle motion in a magnetic field, we note that there is no constraint on the in-plane motion and thus the term in x2 would take on arbitrarily large values, were we to treat it as a perturbation on an x-independent trial wavefunction; we have therefore regarded this term as part of a zero order Hamiltonian, .ie' o- The eigenvalues of yt'o are simply harmonic oscillator levels with a frequency We cos 8. The requirement that eu/ez vanishes at z = ± d/2 again requires that (in lowest order) the wavefunction has no z dependence; i.e., u(x, z) = u(x). Since particle motion in the z direction is restricted, we can treat .ft' as a perturbation. The eigenvalue is again minimized by setting k; = 0 (the perturbation produced by the last term in Eq. (12.8b) then vanishes by symmetry). The resulting perturbation produced by x' is of the same form as Eq. (12.3). The total eigenvalue is then (12.9) Defining H e211
=

(12)1/2cPO/2n(d

and HcH
(J

cPo!2n(2 we may rewrite Eq. (12.9) as


(12.10)

" 1 - He2(O) I cos III + H~2(8) Sill 2 O', 2

».;

He211

note that the angular dependence of this expression, first suggested by Tinkham (1963), is radically different from that of Eq. (11.7b); in particular, it has a cusp at 8 = n/2. We next treat the problem where the boundary condition is altered such that u(z) I z +dl,
s=

I tun superconaucnng

slaos

= O.Tn practice, this condition is approximated by a superconducting film which is sandwiched between two films containing a large concentration of paramagnetic impurities (such as ferromagnetic films) where 'pair breaking' effectively drives the order parameter to zero at the boundary. We return to Eq. (12.1). For Ho = 0 (we = 0) the eigenfunction satisfying our boundary condition is

u with kdl2
=

= uocos(kz)

(12.11 )

± nl2

for the ground state. This leads to a suppression of the transition temperature h1n2 -rx=-2m*d2

governed by

or
(12.12)

here TeO is the bulk transition temperature and T, is the film transition temperature in the presence of the new boundary condition. Application of a parallel magnetic field induces an additional reduction in T; which we calculate from perturbation theory (W ong and Ketterson
1986):

<u(z) 11 m*w~z21 u(z) - brx --'----__:_:-=----_.::_---'----'-= <u(z) u(z)


1

1 2 m*wc 2,(1 d'

1) 12 - 2n2 (12.13)

= 0.0l63m*w;d2.
Combining Eqs. (12.12) and (12.13) gives

(12.14) Eq. (12.14) may be written in terms of T', as


2

Helll =

(m*C)2 2a(Tc - T) e* 0.0327m*d2·


1 1

(12.15)

Ifwe define a coherence length involving Tc' rather than Tco, by


~=

h2 )1/2 ( 2ma(Tc - T) ,

(12.16)

then we may write H


e211

---

5.53CPo 2n,;d'

(12.17)

which is to be compared with Eq. (12.4). The angular dependence follows from inserting (12.17) into (12.10).

~~s-u-r-t-a-c-e---------------superconductivity

A superconducting half space in a parallel field may exhibit the phenomena of surface superconductivity (Saint-James and deGennes 1963). We take the interface to be the plane z = O. The critical field is still determined by the eigenvalue of Eq. (12.1) (kx = 0 with the requirement that au/uz = 0 at z = 0). For values of Zo (i.e., ky) such that the 'orbit center' is deep in the bulk of the superconductor (zo -> - 00) the corresponding eigenvalue yields the bulk critical field; the boundary condition is adequately satisfied since the ground-state wavefunction of our harmonic oscillator is a Gaussian (centered on zo) having negligible amplitude at the boundary. For the opposite extreme, where we choose Zo = 0, the same Gaussian wavefunction also satisfies the boundary condition, since the derivative at the peak of the Gaussian (now centered on Zo = 0) vanishes. The question arises whether there is a smaller eigenvalue (than the bulk eigenvalue) for some value 0 > Zo > - 00. To examine this situation, we reflect the harmonic oscillator potential V = tm*w;(z - zo)2 through the plane z = 0 such that
(13.1)

Note that this is not a harmonic oscillator potential but rather has two minima (located at z = ± I Zo I) and the slope a V/cz is discontinuous at z = O. If we seek a nodeless solution for the motion in this potential (which is symmetric under a sign change of z), the boundary condition at z = 0 will automatically be satisfied, and it should correspond to a possible ground state. We can see immediately from Eq. (13.1) that an eigenvalue smaller than that of the bulk must exist, since the potential described by Eq. (l3.1) is always less than or equal to that obtained by extrapolating the potential from either half space into the other. The above-defined eigenvalue problem admits an exact solution in terms of Weber functions (Saint-James, Thomas, and Sarma 1969). However, an adequate solution is obtained using a trial wavefunction u ~ e-rz2 (deGennes 1966). Minimization of o: with respect to rand Zo yields
Ci

=( 1--

2)1!2IiWe
IT

I e* IIiHc3 -=060--2 . 2m*c'

(13.2)

where we have defined the upper critical field associated with this surface superconducting state as He3. The exact solution yields 0.59 rather than 0.60 for the numerical factor. The upper critical fields for bulk and surface superconductivity are then related by (13.3)

50

Surface superconductivity

51

Surface superconductivity persists to higher magnetic fields (in a Type II material) and thus may (in a conductivity measurement) mask the bulk transition. The state may be suppressed by the following: (1) rendering the surface rough (by sand blasting); (2) using sufficiently high measuring currents to destroy surface superconductivity; (3) depositing a normal-metal overlayer. We now briefly discuss the case in which we have a thick film, of thickness d (with the origin in the film center), rather than a superconducting half space. For films with d;:;; (, the amplitude ofthe G-L wavefunction at nucleation would be highest in the immediate vicinity of the surfaces at z = ± d/2 for He211 < H < He3• As d is reduced, the eigenvalues (satisfying the boundary conditions at z = ± d/2 associated with bulk and surface superconductivity) approach each other and, below some distance d = dc, only a state which has its wavefunction centered on z = 0 is stable (Saint-James et al. 1969). As we saw above, when the G-L equation for the magnetic field at an arbitrary angle is written in terms of coordinates for which thc boundary conditions are conveniently applied, it is nonseparable. Thus, for thick films where u ~ u(x, z) (rather than thin films where u ~ u(x)), we must resort to a numerical, variational or perturbational technique. This complicated, and to some extent not fully explored, problem has been discussed by Minenko (1983), Saint-James (1965),Yamafuji, Kawashima, and Irie (1966) and Saint-James et al. (1969). However, we note that, at some angle 8 = 8c' we expect the eigenvalues associated with the surface and bulk superconducting states to merge.

4I>~~-h-e-T-~-p-e-'-'----------superconductor belowHc2 for H just

The discussion of the upper critical field given in Sec. 10 involved the linearized G L theory and is strictly valid only for the transition line in the H-Tplane where 11 = HdT). For fields below this value we must include the nonlinear terms in Eq. (9.15). Tn this section we obtain an approximate, but analytic, solution for the G-L equations which is useful for a small range of fields just below He2. This problem was first examined by Abrikosov (1957). We recall the solution of the linearized G-L equation obtained in Sec. 10 for a 'nucleation wavefunction' at 11<2(for H II z) which follows from Eq. (10.6) with n = 0 and kz = 0: (14.1) where we dropped the subscript from k; and Xo == (hc/e*Hc2)k = 4Jok/2rrHc2 = (2k; here C is a 'normalization constant', which vanishes for H = He2. As noted in Sec. 10, this wavefunction is highly degenerate in that I< can have any value (so long as Xo lies within the superconductor). The wavefunction which is the solution to the full G-L equations for H just below 1102 is expected to resemble some particular solution ofthe linearized Eq. (10.3). At the same time, based on our earlier discussion, we also expect some sort of vortex lattice to be present. We therefore seek an approximate sol ution for t/Jas a superposition of terms of the form (14.1) chosen so as to be appropriately periodic in both the x and y directions (thus forming the lattice). Eq. (14.1) is periodic in the y direction as it stands with a period b = 2rr/l<. We may retain this periodicity in the y direction while simultaneously forming a function which is quasiperiodic in the x direction by substituting nk for k in (14.1) and summing over all n to yield
n=

t/J(x,y)

C
n~

+ o:

einkYe-(X-XY/2I",

(14.2)

where Xn = n~2k = 2rrn~2 /b and C is a (common) normalization constant. We define the period in the x direction as a == x, + 1 - Xn = 2rr~2 lb. and the area of the unit cell is constrained by ab = 2rr~2. Eq. (14.2) is not, however, strictly periodic in x: if we substitute x + a for x this is equivalent to replacing n by It + 1 in the summation, which chanqes the phase of the waiefunction by
eiky; i.c.,

t/J(x + a,y)

e-ikyt/J(x,y).

(14.3)

This behavior is inherent in our choice of the gauge Ay = 11x; under a coordinate displacement x -> x + a, Ay -> Ay + H a leading to a change in the canonical momentum nhk --+ nhk - (e* /c)Ha

52

Type /I superconductor

for H just below

HC2

53

whichfor H ~ llc2leads to nk

--->

(n

+ l)k. Eq. (14.3) shows that


(14.4)

i.e.,the proba bility density is strictly periodic. Eq. (14.2) is not however, the most general periodic function that can be formed by superimposing forms based on (14.1) since we may introduce a complex expansion coefficient associated with each term, provided we impose the periodicity condition Cn+v = en where v is an integer.Our wavefunction then has the form
n-

I/I(x,y)

= n= -

(14.5)

wherethe quasiperiod in the x direction is given by a = v(xn + 1 - xn) = 2rrv~2[h and the area of theunit cell is constrained by ab = 2rrv~2;v corresponds to the number of vortices per unit cell. In practice there are two lattices of interest: (i) the simple square lattice and (ii) the triangular(hexagonal) lattice. For the simple cubic case all the Cn are equal (v = 1) and I/I(x, y) has theform (14.2) with a = b = 2rr/k and x, = n~2 k = na; hence
IjJ {x, .v) =
I

cI

n::::

e2nin.v/(~e - (x

na)2/2

~2.

(14.6)

Forthe triangular case

en 2 = Cn and C1
f

If; L(X,)')

c[ I

iCo; Eq. (14.5) then has the form


(.y-x")lj2~1

einkYe

+i
11

I
=odd

e-inkYe

(X_X")1!2~2l·

(14.7)

n - even

Writingk = 2rr/b and a = 2~2k, noting b sums,Eq. (14.7) becomes

j3a

for a triangular lattice,

and combining the two

n = even

or (14.8) We fix the normalization constant, C (our only remaining internal free parameter), by requiringthat the free energy be stationary with respect to a variation of 1/1. If we temporarily ignorechanges in the vector potential this is equivalent to setting (aE/aC)A = 0, which yields (14.9) Denotingquantities of the form (ljV) written
-. 1 2m'"

Jf(r)d3r

by a bar, where Vis the volume, Eq. (14.9) may be

1(11 -:-V
1

e*)
C

If;

12 + ~ 1 1/1 1 + -/31 a
2

1/1 1 4 =

o.

(14.10)

I. Atriangular lattice is a centered rectangular lattice with h =

flu

and two vortices per unit cell.

54

Part I

Phenomenological

theories

We must also ineludc relative

the effect of a change

in the internal arising

field, the magnetic

induction

Hir]. W

write this as a sum of two contributions arising from circulating tions by supercurrents

from: (i) an explicit

shift in the external shift

fiel.
H(S)(I

to He2 (since we have an exact solution associated neglected

for the latter), and (ii) a (screening) in Sec. 10). We denote

with the vortices (this latter effect is vanishingl the sum of these contribu

small for H = Hc2 and was therefore

(14.11 with

Associated

with the internal

field

H(l)

is a vector potential

A( 1).

Expanding

(14.9) to first order i

A(l) yields (in a London

gauge) he" A(1)(f)· [(1jJ* V--, __ Ao 21m*c he cancel

PIIjJ14+-.
where the remaining terms

ie*)

1jJ-1jJ

(ie*)

V+-Ao he

1jJ* =0,(14.1:
equation a

l
s:

(since they satisfy the linear

Schrodinger-like we identify

He2)'
The last two terms may be written associated as (l/c)j(S)(f)' A(ll(f) where lattice. Introducing j(S)(f) as th current density with the vortex

screening integrating

Vx

H(s)

(4rr!c)jls)(r

by parts, and writing H<l)(f) = V x A(l)(f), Eq. (14.13) becomes (14.14

where

H(s)

is parallel

to

H(l)

by symmetry.
(H(s)

To the accuracy

we are carrying

the calculation, current,


1

H(s)

the diamagnetic To obtain calculation dimensional) momentum

response the field

field
H(s)

4rrM(s)).
the associated magnetization the contours of constant j(sl(f).
Th

we first calculate by noting coincide, that

of jIg) is facilitated streamlines operator

~/(f)

and the (two the canonica

n( = m*V
Ie'

of

j(s)

which we now demonstrate. operator)

We introduce

where V is the velocity

~ h e* II =-;-V--A.
which has the commutation monic oscillator II_ problem (destruction), relations we introduce

(14.15

[IIx, IIy]

= i(e*h/c)H.

In analogy

the raising

(creation),

n+

with the standard

har

= IIx

+ illy,

and lowerin

IIx - illy, operators."

The ground

state has the property

II_1jJ = 0 frorr

which we obtain

[( If we substitute
lowest order) as

h -;----Ax a e*
1

OX

ClOY

) -i

(Ii a

e* -;----A,
C = }

)l

1jJ=0.
(14.16), and identify

(14.16
the current (in

the wa vefunction

in the form IjJ

IIjJ 1e;<!> into

(14.17! 2. The linearized Schrodinger-like G-L Eq. (10.4) may be written Il


+

Il _ if;

O.

Type" superconductor tor Hjust below

HC2

55

weobtain (on equating the real and imaginary parts separately to zero)
j~)
=

e*/1 G 2m*

oy

I 1/1 I 2

(14.1Sa)

and (14.1Sb) Fromthe structure of (14.1Sa) and (14.1Sb) we see that V I!/II.lj(s), The magnetization field, H(s), followsfrom Maxwell's fourth equation, V x H(s)(r) = (4rr/c-jj<S\r); comparing this equation with Eqs.(14,IS) yields (14.19) wherethe constant of integration was fixed such that H(s) vanishes when II/II = o. We may interpret H(s) as a magnetization through the usual definition B = H + 4rrM(s) where M(S)(r) - (e*h/2m*c) I!/I12z; note it is negative as expected (since superconductors = netic). We now combine (14,12), (14.14), and (14.19) to obtain e*/1 !/I12 [ Ho - Hc2 - 4rr--I!/I e*h fJ ---- 14 + --I I!/I 2m*c are diamag-

2m*c

12

o.
I(

(14.20) and H c2

Writing!/I(r)= ~I of(r), where !/I6= I x] /fJ, and using the definitions (9.29) and (10.7) for wemay write (14.20) as

4( 1) f2(
1- 21(2

1 - -Ho) Hc2

O.

(14.21)

The quantities p and may be calculated provided we choose a symmetry (!/In or !/ID from(14.6) or (14.S)) and a lattice spacing. It turns out that near H c2 the ratio (14.22)

(introduced by Abrikosov (1957)) is independent of H 0' Introducing this quantity into (14.21) we have

(14.23a)

riA 1 -

2/,;2

1 )'

(14.23b)

56

Part I

Phenomenological

theories

--

From the above

I VI I 2

(I exl / fJ)f2

--

and

I if; 14

fJ A( I if; I 2)

,--',2

.
of a Type II superconductc average of the intern and Hc2 we have (14.2 as the position of
K

We now proceed field, H(r): B

to a consideration induction

of the thermodynamics which is defined

We begin with the magnetic

== Ho

+ II(s);

using (14.19) and the definitions

where obtain

we used the relation the thermodynamic

He2

J2KHc

in the second

step. Substituting

(14.23a) for f2

magnetic

induction

B = B(H 0) as (14.25

We may invert this relation

to obtain

Ho

Ha(B) as
(14.25

From (14.25a) we obtain

the magnetization,

(B - Ho)/4n,

as (14.25

note this is linear near H e2 as discussed We now calculate (9.13) for the full free energy we have

earlier. the variational expression (14.9) with E

the free energy. Comparing

Rewriting

the first integral

and substituting

the unaveraged

form of (14.24) this equation

bccom

(14.2 Substituting Eqs. (14.23) we obtain F = F(Ho) as

(14.2
Alternatively, using Eqs. (14.25), we may calculate F

F(B) for which we find

Fo

+ gn

V[

B - -1-+-r-'J -(2-K--;:.2---1)
A

(B-Hc2)2

. parallel

(14.2
to the axis or a thin film wi

For the case of a long cylinder

with an external

field,

Hex!'

Type 1/ superconductor

for H just below

Hc2

57

000

o o

o
(a) (b)

Figure 14.1 Schematic diagram of square and triangular vortex lattices. The dashed lines show the basic unit cells.

Figure 14.2 The spatial configuration of I if; I 2 near H 02 for a triangular vortex lattice. The numbers labeling the contours specify the square of the reduced order parameter. (After Kleiner, Roth, and Autler (1964).)

58

Part I

Phenomenological

theories

Hex' lying in the plane (where Hex' = Ho) wc would use (14,27): for the case where the external fiek is perpendicular to the film (where Hex' = B) we would use (14,28), The only remaining problem is to calculate the Abrikosov parameter f3 A' This calculation i somewhat tedious and we refer the reader to Saint-James et al. (1969) for the details. The result are

f3~
and

1.18

(square lattice)

(14.29a

rJ~ = 1.16

(triangular lattice).

(14.29b

The triangular lattice is thus slightly more stable. A schematic diagram of the square anc triangular vortex lattices is shown in Fig. 14.1. The spatial configuration of 1 if; 12 near Hc2 for, triangular vortex lattice is shown in Fig. 14.2. The above treatment has neglected the elTectsof an] in-plane anisotropy which, given that the structure is sensitive to f3 A' can (and in some cases does alter the observed symmetry.

.,.1------~ The Josephson effects


15.1

The Josephson equations

Letus recall our expression for the G-L current density ie*h j = - -(lp*Vif; 2m* - if;Vif;*) -

-I if; 12A.
m*e

e*2

(9.17b)

Wewrite if; in the form if; = I if; I ei<I> before, obtaining as e*fJ j=-1if;12 m* [ V<l>--A e* he

J;

(ISJ)

i.e.,the current in a superconductor involves the gradient of the (gauge-invariant) phase (recall that real G- L wavefunctions carry no current). The G-L boundary condition at an insulator-superconductor interface was given earlier as 11 e*) fi· ( i V - ~ A if; = 0, (12.2)

whichis equivalent to no current flow through the boundary. Ifwe have a 'junction' which weakly couples two superconductors (formed, for example, by a thin insulating layer between the two superconductors through which electrons can tunnel), we must modify this condition. Taking fi II we replace (12.2) by more general (yet linear) boundary conditions 1 x, oif;! ie* -..,- - -h A,if;1 = ox e and (1S.2b) theparameter i is the same in both equations since it is a property of the boundary and not the superconductor designations, 1 and 2. We insert (IS.2a) into Eq. (9.17b) yielding

IP2 ---;A

(lS.2a)

t. =
1. This form is identical

ie*h [ * oif;1 Oif;iJ e*2 2 -2 * if;! --;-. - if;! --;-. - -*-1 if;! I A,
m uX ox mc
= M22

to Eq. (9.44) with Mll

= 0,

- MI2

M';-/

A.

59

60

Part I

Phenomenological

theories

ie*h [ = _._ 2m*

VI* (if; __2. 1 i"

ie* + -A he

xli

ie* if; ) - if; (if; * - - A 1/1* ~ J.* he


x 1

)J

e*2 - --11/1 m*e

12A.
x

(15
In the absence reversal, of magnetic a toms, the supercond
--> -

ucting properties A. Under

are in variant

under til] both sides fro

which results

in if; --> 1/1*, j

j, and A

--> -

these operations

Eqs. (15.2) turn into their complex

conjugates

and hence J. must be reaL We then obtain

(15.3) (15.
Writing

if;;

ing material,

I if;; I ei<I\ and assuming I if; 1 I = I VI2 I , we have


j=irnsin<l>21

both sides arc prepared

from the same kind of superconduc

(15.:

where

(15.6,
and

(15.6t
Note j.; is the maximum In deriving present argument. independent. in the junction. Under current When density that may be carried field is present by the junction. flux density an is time in the junction, the phase Eq. (15.5) we have assumed no electric that no electric field and magnetic a field is present

We generalize

to the case in which

by using a gauge invariano

a gauge transformation A
-->

(recall H = V x A and E

=-

V V - (l/c)(c'A/at))

+ Vx

(15.7

and 1 aX V-->V-----'c Dt' where x(r, t) is an arbitrary contains single-valued function and Vis the potential.

(15.8
Since the G-L equation by

the form V - (ie* /l7e)A, we must change <l>


-->

the phase of the G-L wavefunction

<l>

+~

e*

he

x( t).

(15.9)

Comparing

Eqs. (15.8) and (15.9) we see that the form

---V=O
at

a<l>

e*

(15.10)
assumed to be independent of time and denoted as V21> then

is qauqe-incariant. If V is (initially) integration of (15.1 0) yields

(15.11)

The Josephson effects

61

or
J

= JrnSlll

.'

(0)

[ 11>21
w] =

e* Ii
.

V21t

J.

(15.12)

Introducing the frequency

ol1>21!ot we see that (15.11) leads to

(15.13)

Theoscillating currentj(t) given by (15.12) will be associated with an oscillating voltage which will be superimposed on the static voltage, V2 I' We next examine the Josephson effects in the presence of a magnetic field. We will restrict ourselves to the case of a relatively weak field where a quasielassical description is adequate; i.e., the dominant effect of a field, which is described by a vector potential A, is to make the phase position-dependent. From the discussion surrounding Eqs. (9.19) we know that for a 'pure' gauge field (one not involving a field H(r)) the only effect of the vector potential is to produce a position-dependent phase; this suggests that in the presence of a vector potential associated with a weak field, H(r), the effect may be approximately incorporated in the phase of the wa vefunction. Comparing Eqs. (15.7) and (15.9) we have the gauge-invariant form analogous to (15.10) as
e* VI1>--A=O. he (15.14)

Therefore, the gauge-invariant


_ (0)

phase difference is given by 2n +- f 1>0 1


2

11>21

-11>21

A dt.

(15.15)

where 1>0 == hel2e is the flux quantum. The fundamental equations governing the behavior of Josephson junctions are the current-phase relation (15.5), the voltage-phase relation (15.11), and the gauge-invariant phase relation (15.15). They are believed to be exact. In the subsequent sections these equations are applied to some simple junction structures and circuits. For example, electromagnetic radiation is emitted from a Josephson junction in the presence of a potential (sec Sec. 17). Eq. (15.13) now forms the basis for defining the standard volt in terms of a measured frequency and the fundamental constants, e and h (Taylor, Parker, and Langenberg 1969).

15.2

Magnetic field effects: the two-junction SQUID


circuit shown in Fig. 15.1 involving two Josephson junctions leads. The loop formed by this circuit is assumed to contain a

Consider the superconducting connected by superconducting

62

Part I

Phenomenological

theories

'll --

= 'musin~<l>ll

ab

--

@H

dc -- --'t = 'mtsin~<t>t
Figure 15.1 Schematic of a de SQUID consisting of two Josephson junctions connected in parallel by supcrconducting links. The path of integration C is shown by the dashed line.

magnetic flux, <D,arising from a position-dependent field, H(r), with some associated vector potential A(r). On entering the loop on the left in Fig. 15.1 (point 1), the current I splits into two components III and It, where the subscripts refer to the upper and lower paths, respectively. From Eq. (15.5) the total current arriving at point 2 is I=Iu+1t = ImusinL1l1>u Imt sin L1<Dt, + (15.16)

where I rnu and I mt correspond to the maximum currents associated with the upper and lower junctions, and L1l1>ll L111>{ the corresponding gauge-invariant phase shifts. Assuming a and are matched pair of junctions for simplicity (l mu = I mz" = 1m), we may rewrite (15.16) as (15.17) The gauge invariant phase shifts may be obtained by integrating VII> around the elosed path C shown in Fig. 15.1. Noting that <Dis a multivalued function that can change by 2nn upon completing the path we have

where n is an integer. The phase differences across the upper and lower Josephson junctions are given by Eq. (15.15) as II>b (D a

L1(D +.__
u

cPo

?nfb
a

A . dt

and

The second and fourth terms in Eq. (15.18) are phase differences in the superconducting leads themselves and are found by using the supercurrent equation (15.1) and the expression for the London penetration depth Eq. (2.6): <Dc <Db =

VII>·dt

=-

2n ¢o

c(

4n). + _ _!j ) 'dt


C

The Josephson effects

63

and <Pa-<Pd=

a V<D·dl=- fa ( A+
2n
d

cPo

4n :[2 __ .. l.j ) ·dt.


C

Substituting the above four equations in Eq. (15.18) gives


Ll<Du-L1<Dt=2nn+2n~ . A'dt+--2n 4nill

cPo

cPo

.j·dt.

(IS.19a)

• C'

The integration of A is around a complete closed path C and is equal to the total flux cP inside the area enclosed by the contour. The integration ofj follows a path C' which excludes the integration over the insulators. If the superconducting leads are thicker than the London penetration depth. the integration path can be taken deep inside the superconductors where the integral involving the supercurrent density is negligible. The phase difference is then simply related to the total flux by
(IS.19b)

Using this equation to eliminate L1cI\, rom Eq. (15.17), the total current is f
(15.20)

When the inductance L of the loop is taken into account, the total flux in Eq. (IS.19b) consists of the externally applied flux cPCXL and the flux generated by the screening circulating current I dr; i.e..
(15.21)

For the identical junction case,


(15.22)

In general, Eqs. (15.20), (15.21), and (15.22) must be solved self-consistently to describe the behavior of the two-josephson-junction loop. For simplicity, we assume the loop inductance is negligible and consider only the effect of the externally applied flux on the characteristics of the loop. The maximum supercurrent density which can be carried by the loop is found by maximizing Eq. (15.20) with respect to L1<D{; i.e.,
L1<D+ t

n~:XL

(n +

1/2)n.

Hence the maximum supercurrent density, Imax, is given by Imax


=

21m cos (ncPCXL) I

To I'

(15.23)

which is periodic in the external flux. Since cPo ~ 2.07 X 10-7 Gcm ', it is clear that the device pictured in Fig. 15.1 can be used to measure very small changes in magnetic field. It is sometimes

64

Part I

Phenomenological

theories

H @a

Figure 15.2 Schematic diagram of an SIS Josephson tunnel junction.

referred to as a two-junction SQUID where the latter is an acronym constructed from the words superconducting quantum interference device.

15.3

The extended Josephson junction

We next discuss the behavior of a single planar Josephsonjunction in a magnetic field. We refer to the junction cross section depicted in Fig. 1S.2. The magnetic field is directed into the page along y. The middle of the junction is taken as the origin and the vector potential in the three regions is
_ { _
-

Hxe-(o-a,2)'AL Hx Hxe(o+1I'2)/'L

(z > a/2) (a/2 (z

(lS.24a)
a/2) (1S.24b) (lS.24c)

Az

>z > a/2)

<-

where the exponential dependencies arise from the Meissner effect in the two superconductors. The phase difference encountered in crossing the junction from z = -::I:) to z = + CfJ at a given horizontal coordinate x is

(1S.25)

Assuming a rectangular junction junction is


1= jmLv

of dimensions L, and Lv, the total current, I, through the

1.,12

. sin L1<D(x)dx
-Lx/2

or

The Josephson effects

65

1.2 ,---,---,--r--,-,--.,.--,--,-----,,---, 1.0 0' 0.8 0.6 0.4 0.2

:=-E

:::E

Figure 15.3 J vs rP characteristics effect is negligible.


m

for a short Josephson junction when the self-field

1m

sin(rr¢l¢o). rr¢l¢o

sin <1>21 ,

(0)

where ¢ == H(a + 2),dLx, and 1m = jmLJ~y. The maximum value of the junction current, 1m"" occursfor sin <I>~O! = ± I (depending on the sign of sin (rr¢l¢o)); thus
I'llax

1 I sin(rr¢l¢o) I m rr¢l¢o '

(15.26)

We note we obtain a 'diffraction-like' (sin x/x) pattern (see Fig. 15.3) involving the variable ¢I¢a where¢ is the flux contained within the junction. The effective junction area entering the definition of ¢ involves an effective thickness (made up of the insulator thickness plus a contribution of one London depth in each superconductor) timesthe junction width, Lx'

15.4

Effect of an applied rf field

An interesting behavior results if an external rf voltage is applied to a de voltage-biased Josephson junction, as first observed by Shapiro (1963). We take the applied voltage to have the form

v=

Vo

VI cos tot.

(15.27)

From Eq. (15.12) the resulting current in the junction will be 1 = Imsin [;~

I(V

VI cos wt)dtj (15.28)

1m sin [<1>(0)

+ OlJt + ()sinwt],

where (!J(O) is an arbitrary phase, WJ is given by Eq. (15.13) and b == e* Vdhm is called the modulation (or deviation) index. Using expressions which are well known in the theory of frequency modulatiorr' we may rewrite (15.28) as
2. See Abramowitz and Stegun (1970), Eqs, (9.1.42) and (9.1.43), p. 361.

66

Part I

Phenomenological

theories

I(t)

i: {sin[<D(O)

+ wjtJ

[J 0(3) + 2

JI

J 211(6)cos(2nwt)

J
(15.29) Using sin a cos b =

+ cos[(D(O) + wjtJ2 n~o .! 2n+ 1((»)sin[(2n + l)wtJ },


where t[sin(a the

J n are

Bessel

functions

of first

kind

of integer

order.

+ b) + sin(a
I(t)

- b)J we may rewrite (15.29) as


=

t; (.! o((»)sin[w/ + <D(O)J


+

Jl

.!11(()){sin[(wj

nCU)l

+ <D(O)]+ ( -

1)"sin[(wj - nw)t

+ <D(O)J}).
(15.30)

If we vary
dependence istics. be 'transformed

Wj

(by sweeping

Va) such that the condition


and would appear junction and voltage hence

OJ

j=

associated

with this term in (15.30), having the amplitude

± nco is met, then the time In(()) and phase <D(O), ould w
characterof the than the resistance is a more analysis

to zero frequency'

as a 'spike' in the current-voltage much smaller current above. than

The impedance leads extending into than confirms) representation experiment behavior.

of a Josephson the cryostat the constant

is usually

a constant assumed (rather

source Further

accurate

source

shows (and

that this results

in a step-like

a spike-like)

current-voltage

15.5

The resistively shunted junction (RSJ) model


the behavior parameters. of circuits involving and small Josephson the distributed junctions, one may model the with so-called the capacity which is of various circuit dissipative processes
3

In analyzing
effects 'lumped'

junction

capacity

Fig. 15.4 shows such a model; here C and R represent where the latter is represented flowing through

and effective resistance across the junction given by"

of the junction,

by a cross. Vj is the voltage

while I is the total current

all three circuit elements

(15.31a)
=

Imsin<D

+ -<D + -<D,
2e 2eR equations

hC ..

Iz.

(15.31b)

where we have used the Josephson

3. The capacitance is largely determined by the geometry of the junction and may he regarded as constant. The resistance on the other hand may depend strongly on the junction voltage; for this reason an additional shunt resistance, which is much smaller than the junction resistance, is sometimes incorporated to bypass this effect. 4. We use MKS units in this subsection.

The Josephson effects

67

+---


Figure 15.4 Equivalent circuit of a resistively shunted Josephson junction.

I
I

<P I

m ..
Figure 15.5 The simple pendulum with an applied torque.

(15.32)
and

IJ

1m sin cD

(15.33)

in obtaining (15.31b). Let us examine the behavior of an RSJ to which a constant current is applied. If I is slowly increased the voltage across the junction will remain zero (VJ = 0), implying <!> sin -1(1 11m), = until I = 1m' Above this point a voltage develops and the junction becomes resistive with a time-dependent phase. To examine the beha vior in this regime it is useful to note that the phase, Cll(t), ofajunction described by Eq. (15.31b) is in one-to-one correspondence with the angle of rotation (which we will also call cI») of a damped pendulum, driven by a constant torque, in a constant gravitational field (see Fig. 15.5): the applied torque, .r = mwyr, gravitationally induced torque, mgl sin <!>, moment of inertia, .f = mr", and damping coefficient, k, are identified with I, lmsin<!>, !zeI2e, and hl2eR, respectively. The regime I < 1m corresponds to a situation in which the applied torque .r is less than a critical torque :Y m necessary to raise the pendulum to an angle Cll n/2 (where the opposing gravitational torque is maximal). for .'Y > :Y m the pendulum = rotates in a manner such that the average energy dissipated per rotation is equal to the average work per rotation (were there no damping the average angular velocity cb would increase without limit in this regime). We now go to a rotating reference frame by writing
<!>(t)
=

cot

+ <!>'(t),

(15.34)

where (15.35) is the average angular velocity and cD' is periodic (with an average value of zero); Eq. (15.31 b) then becomes

68

Part I

Phenomenological theories

he .. fz. -$' +-$'


2e 2eR In the limit where the departure may neglect the $' contribution harmonic oscillator. $'(t) I
m

+Imsin[wt

+<D'(t)J

O.

(15.36)

of cI) from a steady state rotation,

oit, is small, i.e. <D'(t) < < 1, we


for a driven

to the last term in (15.36). We then have the equation

Writing

<D~ sin(wt

+ fi) leads

to the pair of equations

- - w2<D~ cos fi - w$~ sin 2e 2eR

hC

e =0

(15.37)

and - - U)2$~ sin fi 2e From the second condition

hC

+-

Ii

2eR

w$~ cos (J = O.

(15.38)

we obtain

tan f =-_ wRC or sin (J where,

(15.39a)

1
(1
2 + orr) 1/2'
7

(15.39b)

==

RC. Substituting , <Do

(15.39) into (15.37) yields 1 (15.40)

Ii -:;;(1 + (J)2,2)1/Z'

2e ImR

Now
-s2e _ 2e <l> = - V =- I R.

(15.41) limit, OJ,

In the limit, (liT)/m/l.

(liT

<<

1, $~l (ImlI). In the opposite =


$;)

>>

1, (15.40) becomes capacitor

<D~

(1/

The requirement equation

<<

I is therefore to

best satisfied

in the limit of large I and large

WI.

The limit OJ, ~ 0 may be obtained differential (15.31b) then reduces


=

by eliminating

the shunting

in Fig. 15.4; the

I
or

. 1m Sill <D

+-

h.

2eR

<D

(15.42)

2iRlm

Iz

f<l>«)
<1>10)

d$ III", - sin <D'

which upon integration

yields
!X2 [( -

$(t) = 2tan-1

1)+ (nt) T
tan

-'X

(15.43)

The Josephson effects

69

~
4.0
'/

"

'/

3.0
/ / /

"/
/

"

'/

2.0
A
"/ "/
"/
/

"/ "/

1.0
/ /

1.0

2.0

3.0

4.0

5.0

T)

Figure 15.6 Time-averaged nonhysteretic I-V characteristic computed in the RS.J model in the zero capacitance limit; a = 111m, '1 = VIRlm. Points A and B give the limeaveraged voltages corresponding to o: = 1.2 and rJ. = 4, respectively. where we have defined (15.44) where IX is a parameter and Tis the period (which is the minimum time required for <D(t) to return T) then (15.45)

to its initial value <D(O)). he voltage across the junction T voltage is sensed with a device which records
V=-c))=---

from Eq. (15.5) is (h<D/2e). If we assume the of V(t) (over many periods
2

Iz..

h1

2e

2e T

the time average

T 0

d<D

h 2n -dt=--=Rlm(1X dt 2e T

-1)·.

1/2

Note that for I

> > 1m (IX > >

1) (15.45) approaches discussion

(15.41). Fig. 15.6 shows the J- V characterisis non hysteretic.


Wo

tics from Eq. (15.45); note that the J- V characteristic Wenow give a qualitative Eq. (15.31 b) in dimensionless where
__ 7 C Wo = (

of the general case when C time in units of

*'

O. We begin by rewriting a parameter

form. We measure

and introduce

fJ

he

I n.'

)1/2.

'
(referred to as the Josephson plasma as

(15.46) frequency).

• woisthe natural frequency

for small oscillations

Then Eq. (15.31 b) may be rewritten <i>


where the derivatives angular velocity O(t)

+ {3<D + sin
==

C[)

='1,
to the dimensionless is then <i> time variable.

(15.4 7)

are now with respect <1>; the angular

acceleration

n = (dO/dcD)ci> = (dO/d<D)O

We define

an (dj

dCll)(Q2/2) nd Eq. (15.47) becomes a

d<D

d (0 -

2 )

+ fin + sin

<D= IX. of (15.48) for various initial conditions

( 15.48) and parameter

The 'orbits' associated

with the solutions

70

Part I

Phenomenological theories

t"J7

------~----~--~~--~$

Figure 15.7 The

Q-1tJ plane trajectories (lower panel) associated with the solution of Eq. (15.49). There are two kinds of orbits: an open orbit for (1/2)Q2(<ll = 0) > 2 and a closed orbit for (1/2)Q2(<ll = 0) < 2. The upper panel shows the potential as a function of <ll.

values (z and are trajectories in the (HI> plane; we may think of this plane as the twodimensional 'phase space' associated with the motion. We now discuss the nature of these orbits. We begin by discussing the orbits when :x = {J = O. Our equation is then d<1>

d (0 2
2 )

+ sin cD

o.

{15.49)

This equation may be integrated analytically in terms of Euler elliptic functions, as discussed in mechanics textbooks. Wc will restrict ourselves to some qualitative statements. Since (15.49)is invariant under the transformation cD ~ - <1>, we expect extrema at <1> = 0 and (when they exist] ± tt. For the rotating case, <t> advances continuously in time with 0 having a maximum at <1> = 0 (where the kinetic energy of the pendulum is largest) and a minimum at cD = ± tt (where the potential energy is maximal); the phase space trajectory is shown in Fig. 15.7. Such 'open' orbits req uire that the kinetic energy always be larger than the potential energy; i.e., (15.50) When this condition is not satisfied the motion is oscillatory and the orbit in phase space is closed as shown in Fig. 15.7. We now turn on a small damping (Ii> 0), but still keeping the drive current (torque) zero (« = 0); we will then have 0(<1» > 0(<1> + 2n) > 0(<1> + 4n), etc., i.e., the rate of rotation slows down. Eventually a point is reached where 0 passes through zero before the system (pendulum) reaches the 'top' after which it reverses its motion. This point marks the transition between rotary

-IT

it

-it

it

-it

it

Figure 15.8 The qualitative phase-plane trajectories, corresponding to the three different situations (A, 8, C) in the fi ex plane in Fig. 15.10. (After Belykh, Pedersen, and Soerensen ( 1977).)

and oscillatory motion. After the transition the phase point proceeds along a spiral, asymptoticallyapproaching the origin. The above described sequence of events is shown as a phase space plot in Fig. 15.8. (Note that since dissipation is present the arrow of time is relevant.) We now consider the general case;« > 0,13 > O.When a > > 1 (l > > 1m), Q is practically constant and given by Q =a/f3. As a is lowered through unity from above the behavior of the systemdepends on the magnitude of 13. Since '-x is starting at a value greater than 1, the pendulum motionmust evolve from a rotary state. If we now decrease a to a value less than 1 (J < 1m) the pendulum can continue to rotate (i.e., we can maintain a steady state motion) provided the work performed per cycle by the external torque, 2na, can compensate for the energy dissipated per cycle; he latter is given by t {3

T
o

Q2(t)dt

{3

f21r
0

dr Q2(<D)-d<D d<D
Q(clJ)dclJ. (15.51 )

= 13

21r
0

Thebehavior ofQ(<1»is not known without actually integrating the equation of the orbit (15.48). However,we can estimate the critical torque, ac, by assuming Q(n) = 0 and Q(O)is the angular velocityresulting from 'free-fall' rotation under the influence of gra vity, which from Eq. (15.50) is Q(Cll = 0) = 2; we linearly interpolate between 0 and tt by writing Q(<D) Q(O)<D/n. We then have =
2nac
~

2jiQ(0)

1r<D

- d<D on

or

A more accurate expression can be obtained from the exact treatment of the free pendulum whichyields
(15.52)

whichis valid for ji ::s 0.2 (Stewart 1968, 1974). Wc see that as fl ~ O,«, ~ 0; i.e., the junction is hysteretic, switching from a constant-phase, zero-voltage state to a phase-precessing, finite-

72

Part I

Phenomenological theories

2.0

[3 = 10
/

r. /( d~o
It

/ / /

__

/
/

~/

0.1 1.0 2.0


T)

Figure 15.9 Analog computer results for the time average of:x( = I/Im) vs 11( = V/Rlm) characteristics for different values of the parameter fJ = l/woRC. (After Johnson (1968).)

1.0

I
I
I

I
I

.\

Bel

---':"-1

0.5

I
I

I I

I
I

I 0.5
1.0

Figure 15.10 Analog computer results for the hysteresis parameter !Xc as a function of {i. Tnthe region above the cxJfJ) curve only one stable solution exists. Tnthe region below the curve two stable solutions exist. (After Johnson (1968).)

voltage state as a is increased above 1, but returning to the original (constant-phase) state ata value a < 1. The J- V behavior of a hysteretic junction is shown in Fig. 15.9. The junction instability line in the a-Ii plane obtained by computer is shown in Fig. 15.10 (Johnson 1968). For a discussion of the case in which the junction resistance is voltage-dependent see Barone and Paterno (1982).

The Josephson effects

73

Figure 15.11 A single junction SQUID coupled to a tank circuit via a mutual inductance M.

15.6

The rf biased SQUID


a magnetic flux with a two-junction method circuits involves interferometer (a de SQUID), as the use of only a single junction SQUID, (Silver and which is a

Rather than measuring Zimmerman 1967).

discussed in Sec. 15.2, an alternative The device involves superconducting three

(see Fig. 15.11): (i) the primary shunted Josephson coupled

loop connected

to a resistively

junction;

(ii) a second loop an external oscillator flux, the the to

(not shown in the figure), which is inductively flux <Pex! = 11;[ex! Iex! as a result of a mutual
lex!'

to the first loop and introduces

inductance,

M ex!' coupling

the two loops and a current, by an external a periodically subtle, changing

applied to the second loop; and (iii) a third loop which is driven at a frequency the operation of order M. of the device, which is somewhat of the primary
=

(typically operating To explain action of


CPex!

107 Hz) which applies

¢rf' through a mutual inductance,


and

we must examine of an RSJ connected are"

CPrf

on the response
CPex! =

SQUID

loop. We begin by discussing

equation of motion

when

CPrf

O. The equivalent

circuit consists

an ind uctor, L (represented

by the loop in Fig. 15.11). The equa tions of motion

(15.53a) where (15.53b) Combining these equations


L)

with the forms in Eq. (15.31b) we have (where we multiply

through

by

- + LIm
2e
The phase variable, according to the relation Eq. (15.54) then becomes

h<D

sin <D

+ --

hLC..

2e

<D

+-

«i.:

2eR

<D

O. by the internally generated

(15.54) flux, flux,

(D, in Eq. (15.54) may

be replaced

cP,

cP

= (cI)/2n)cpo'

Incorporating

the contribution

of the external

CPex!,

5. We use MKS units in this subsection.

74

Part I

Phenomenological theories

(a)

(c)

q,

Figure 15.12 The static behavior of a single-junction superconducting loop in an applied external flux ¢ex,: (a) cj)mN>o = 0.25; (b) ¢m/¢O = 1.25; (c) a 'linearized' version of Fig. lS.12(b) showing the path traced out by (p and ¢ext for the case ¢ = ncj)o involving three branches.

(IS.S5)

where ¢m == LJ me w~ = 1/LC, and T = L/R. We examine the static solutions of this equation for the responses ¢ = ¢(¢ex,), which are shown in Fig. IS.12(a) and (b). Fig. IS.12(a) shows the behavior when ¢m/¢O = 0.2S, where a continuous, single-valued evolution of ¢ with ¢ext is obtained; a line with a slope of 1 is shown for reference. Fig. IS.12(b) on the other hand shows the behavior when ¢rn!c/JO 1.2S. We now have a multi-valued behavior. Between successive points = 4)c at which dcp/d¢ext diverges (indicated by the dashed vertical lines in Fig. IS.12(b)) there are alternating regions with positive and negative slopes; the former are stable while the latter are unstable and the system spontaneously oscillates. However, the frequency involved is typically in the microwave regime and if T is sufficiently short the oscillations damp quickly (relative to the period associated with ¢rr)' The system then, effectively, switches from one stable (positive slope) region to an adjacent one. (By Taylor expanding in the region c/J/¢o ~ n (an integer) one sees that the slope of the plateau at this point is (2n¢m/¢O + 1)-1.) The upward and downward directed i

, ne Josepnson eTTeCrS

/'0

arrows in Fig. IS.12(c) indicate how the system switches for increasing and decreasing ¢ex!' respectively. To use these characteristics to make measurements we need a means of locating a stable branchand the position alongit. Todo this we introduce the third loop, which is driven sinusoidally byan rfgenerator, and induces an additionalflux ¢rf(t) = ¢rf sin wrft into the primary loop. We will assumethat a capacitor is associated with this additional loop and that the pair resonate at the rf frequency.The quality factor, Q, of this resonant circuit along with the resonant frequency fixes its bandwidth, ~OJ = wrrlQ, which in turn fixes the response time Trf ~ (~w) -1. Suppose that at t = 0 the rf generator is connected to the resonant circuit. The amplitude, ¢rf, of the oscillatory flux will then increase with a characteristic time Trf. When the total externally induced flux, ¢ex, + ¢rf(t), exceedsone of the thresholds, ¢e' (associated with the primary SQUID flux, ¢, depicted in Fig. 15.l2(b)),the SQUID loop becomes resistive, since it is then in a finite voltage state (oscillating at somecharacteristic frequency much higher than Wrf' which we ignore). This results in a rapid dissipationleading to a partial quenching of the amplitude, ¢r6 however, the process then repeats itself.The peak amplitude of ¢rf, denoted rPrf (which can be measured with the appropriate electroniccircuitry), is then a measure of ¢ex! as we will now discuss. Referring to Fig. IS.l2(b), we identify two 'special' points: (i) the midpoint of a stable branch, where ¢ex, = n¢o (with n an integer), and (ii] the midpoint of an unstable branch, where fex! = (n + !)¢o' We first examine case (i). Assuming sufficient rf drive, as the oscillations build up intime, the thresholds, ¢u, connecting the 11, n + 1 and 11,11 - 1 branches are encountered and the SQUID switches symmetrically between the three branches. If the rf drive is further increased the thresholds connecting the 11 + 1,11 + 2 and the 11 - 1,11 - 2 branches also are encountered and theSQUID then oscillates symmetrically between five branches and so on. An example of a path traversed by the system for a circuit involving three stable regions is shown in Fig. IS.12(c). For case (ii), where ¢ex, = (n + ~)¢o, the path traversed by the SQUID is nominally centered at the midpoint of an unstable region. As the rf drive is increased from zero, a level is reachedwhere a complete circuit involves traversing the portion of the two stable branches lying directlyabove, n + 1, and below, 11, the unstable 11 + ~ section. The threshold amplitude corresponds to the extent of the unstable branch. Since this extent is smaller than that of the stable region,the threshold for first completing circuits in which the SQUID switches irreversibly is lowerfor case (ii) than for case (i). Fig. IS.13 shows the peak amplitude of the rf tank circuit, VT, as a function of the rf drive current, Irf• The slope is initially high but drops abruptly (nominally to zero for a peak reading detector) when the thresholds for SQUID transitions are encountered, involving 3, S, ... stable branches for case (i) and 2,4,6, ... for case (ii). From the symmetry it follows that cases (i) and (ii) represent the maximum and minimum values of VT (or rPrf) for a given drive level, I rf ' It is clear from the above discussion that if ¢ex, is continuously varied (at fixed Irf), the peak amplitude (fjrf will move back and forth (in a triangular fashion) between maxima, at = l1¢o, and minima, at ¢ext = (n +})¢o' By observing the change in the number of flux quanta, one can accurately measure a change in a current, Iext, associated with the loop generating ¢oxt. Fig. IS.14(a) shows the characteristic 'staircase' dependence on Trf for an rf SQUID operated at 27 MHz that was shown schematically in Fig. IS.13. The two curves correspond to the limiting cases for the de external flux. Fig. IS.14(b) shows the triangular dependence of the detected rf voltage Vr vs I ext characteristics. The different curves refer to different values of the rf drive current, T rf '

«;

76

Part I

Phenomenological theories

<il = (n+ 1/2)$0

;-/ /

Figure 15.13 Relation between the tank circuit voltage Vr and the rf current amplitude I"f for the cases of integral and half integral numbers of nux quanta. For intermediate values of flux, the voltage steps occur at intermediate values of VT•

>
'0
>

f-<

.,s
Cl

-g
o
2l

.... o

.,

(a)

rf drive, Irf

(b)

de field,

c/Jdc

Figure 15.14 Experimental responses of an rf SQUID in the dissipative mode: (a) detected tank circuit voltage Vr vs rf current amplitude J rf (the two curves are the two limiting cases for external dc flux (IJde = 11(/>0 and (/>dc = (n + })¢o); (b) VT vs applied magnetic flux curves at different values of I,.f. (After Zimmerman (1972).)

If the loop gencratingcjJext is coupled to another loop which, in turn, is coupled to a magnetic sample in a fixed field (via an all superconducting circuit), changes in the magnetization with temperature or some other parameter may be observed. (Note that current changes induced through magnetization changes arc persistent.) The self-inductance of the two loops, together with the mutual inductances coupling them to the SQUID and the sample, make up a 'flux transformer' the characteristics of which can be optimized for maximal sensitivity. For further discussion of the rf SQUID see Zimmerman (1972) or Barone and Paterno (1982), where many additional references may be found.

.f-----~-'h-e-J-o-S-e-p-h-s-o-n-,a-t-t-ic-e-in 1D

Certain intermetallic compounds have a layered structure in which the in-plane conductivity greatly exceeds the conductivity normal to the layers. This applies for most of the high temperature superconductors. Such structures may also be made artificially by depositing alternating layersof a superconductor and a low conductivity material, The simplest model is to assume infinitesimally thin superconducting layers which arc coupled via order parameter tunnelinq (J osephson coupling) through insulating layers of thickness s. Following Lawrence and Doniach (1971), we introduce a modified free-energy functional of the form
(16.1)

where

(16.2)

herewe have defined ,4= = (lIs) + 1), Azdz. The structure of this equation is similar to Eq. (9.13) with respect to the in-plane components while the intcrplane coupling is seen (on expanding the exponent) to be a generalization of the operator ie* ---A_ Dz he ~ to the case of finite differences. Our order parameter JjJ "(r) has a discrete dependence on the index n and a continuous dependence on r = xx + yy: the total vector potential A I + Azz is, however, defined at all points. Variation with respect to JjJ* yields
--

J~~

112 (
2m*

ie* V--A he

)2 !If

"

(16.3) 77

78

Part I

Phenomenological theories

Variation with respect to A_Land Az yields Eq, (9.16) with (16.4) and
.i11.,,+1.n
=

Eq. (16.5) for the tunneling current flowing between planes nand n + 1 is equivalent to Eq. (15.4) (written in a gauge-invariant form). Solutions of Eq. (16.3) were first examined in detail by Klemm, Luther, and Beasley (1975). As before, we confine our interest to the evaluation of Hc2, in which case we may neglect the last (nonlinear) term. The external magnetic field will be assumed to lie in the x z plane. In order to take advantage of the layer periodicity, we choose a gauge for which the vector potential has no z dependence: (16.6) We seek a solution of the form

noting that Ifz

=-

Hx sin 0, we find that

- --

1i2 d2 __ 2m* dx? 2

+ -m*w2(x - x
c

1 2

)2 cos?
0

e
=

1i - -;;2 [( cos mz s where

kzs

e* + --'H s x sin 8 )]}- 1 + ex u,,(x)


lie

0,

(16.7)

Xo = l1eky/e* H cos e is an orbit center. For the field perpendicular to the layers, we again have a harmonic oscillator problem and we recover our earlier result - c: = ~11(J)e(O). For H parallel to the planes Eq. (16.7) becomes

112 d - -- __2 - --/1 2 2m* dx ' m:s


2

[(

cos kzs

e* ,+ -, HS) x-I he

] + c: }

un = O.

(16.8)

Changing kz has the effect of shifting the origin on the x axis. We shall initially fix the origin such that the argument of the cosine vanishes when x = O.Ncar the zero-field transition temperature, where ex is small, we expect the critical magnetic field, H e211' also to be small and further that the lowest eigenfunction u, which by Floquet's theorem is periodic (since the potential is periodic), is concentrated ncar the minimum of the 'potential'
V(x) =
--2

m:s

li

2 [

1 - cos (e*HS_. x )] . -, he

(16.9)

Expanding Eq. (16.9) to second order, we obtain a harmonic oscillator potential which, on substitution into Eq. (16.8), yields

The Josephson lattice in 1D

79

_!i__ 2m*
where

ul! + [~m*w2 2

(~)X2 2

+ rxJu = 0'

(16.10)

Thelowest eigenvalue is - o: = ±hme(n/2), yielding an upper critical field H e211 = cjJo!2n((z where ~ == (h2/2m* I o: I )112 and (z == (h2!2m;I,:x I )112. For temperatures further from Te, where the eigenvalue - c: and the associated parallel critical field are both large, u becomes less concentrated at thepotential minimum and our expansion of Vex) breaks down. To examine this limit, we write Eq. (16.8)in its 'canonical' form (Abramowitz and Stegun 1970) by introducing new variables 2v and

== -. hc

e*Hs

+ tt., 2q ==

--0

2cjJ~ m* n2H-~m:
-

whichyields d2u
-2

dv

+ (a'
q

- 2qcos2v)u

O.

(16.11a)

For large H, both at and 1970,p. 722)

are small and arc related by the expansion (Abramowitz and Stegun

(16.11 b) Inserting our expressions for


a'

and

into Eq. (16.11b) we obtain (16.12)

Sincethis is a high field expansion, Eq, (16.l2) is valid only for temperatures such that (z ~ s/)2; however, Hc2 '! is real only for (z > s/j2. No (real) solution exists for (z < s/j2 and hence the critical field becomes formally infinite (in this model) at a temperature T* such that (z(T*) = s/ this divergence would be removed by the effect of paramagnetic limiting (see Part III, Eq. (41.45)). Deutcher and Entin-Wohlman (1978) generalized the above model to the case of a superlattice of thin slabs of thickness d separated by Josephson-coupled insulating layers of thickness s. We shall continue to take the z axis normal to the layers, and we place z = 0 at the interface with the lower surface of a metallic layer. The free energy is given by

)2. In a real system,

80

Part I

Phenomenological

theories

F=I
11

f {f
dxdy 2m;s"
1

IlD
11f)

dz

[h2 ~

1(

_In

ie*) V----:-A

II(

!11,,(X,y,Z)

12

+ Ct.10/ lx, + __ h_2_,

y, z)

1- + 21

(J

o/,,(X, y, Z)

14
-

o/ll+l(x,y,(n

I)D)exp(

ie*. All_

he

IS)
(16.13)

here ,4,,+ I == (I/s) J~nD\I.lDAJx, y, z)dz and D = d + s. We seek a solution to Eq, (16.13) in the limit d < < ICL and d < < ¢. The first of these assures us that we may neglect the diamagnetic screening currents and hence we shall take the magnetic only the case where H is parallel that the magnitude field as uniform to and given by the external field, H. The limit We consider however,

and work in the gauge A parameter is constant a solution

= - Hxz.

d < < ¢ implies

of the order

along

z [or a given x;

we do allow the phase to vary. We thus choose

of the form

The z component

of the gradient

energy in Ell. (16.13) is

C<D e* ( --+-Hxhe jjz

0/ 12
in this term for large x, we choose the phase in the form

To avoid an arbitrarily (D

large growth ---:-l1x(z

e*

he

- Zll);

211 is chosen to minimize the overall free energy. For an isolated slab it has the value Z/1 = nD + d/2, i.c., the origin must be taken to be in the center of the nth slab. For the ground state, we assert that, by symmetry, Zll will have the same value in the coupled superlattice. We include no variation in the phase along y since this results in an increase in the free energy. Varying Eq. (16.13) with respect to !II* yields the linearized equation

- --17 Integration symmetry) yields over

m;s2

cos (e*llDx) ~-~

he

-1u

l}
+ Ct.1I
dz
term

O.

(16.14)
vanishes

z (noting

that

the second

in the first parentheses

by

(16.15)
where rewrite we h.ave introduced Eq. (16.15) as the notation

we

e*H/m*c. Using Eqs. (12.3) and (12.4) we may

I'

The Josephson lattice in 1D

81

--

17

2m* d.x"

--0

2 11
-

--

112 [ cos (e*HD) x-I ----m;s2 he

+a

H2) u 1- H~

0,

(16.16) with Eq. (16.X)

where H II is the parallel except for the additional To obtain

critical field of the isolated term multiplying behavior, a. we again

slab. The equation

is identical

the low-field
X4

expand the cosine term; this time we shall


by perturbation Gaussian theory. We require wavefunction the
[/0

include the H4 term, the effect of which matrix element of with respect which is given by (Landau and Lifshitz

we calculate

to the unperturbed 1977, p. 136)

ground-state

(16.17) Our eigenvalue equation is then


=

':i

1. D (m"'.') -hwe- 2 s m~

112

+ m*(!)~d2
24

3 D2) 1 - - -----0 4 d:

+ ...

(16.18)

or ( 16.19) For H
->

0 (T

---7

Tel, we have the limiting behavior H _


e211 -

¢o . 2nr;~z(D!s)'
with He211 behavior,

(16.20)

this we refer to as 3D behavior To obtain the high-field (16.11).The constants

o: T; - T.
we must write Eq. (16.15) in the canonical form of Eq.

now become

and

4¢2 a' = (2n)2H2oD2r;2


Using Eq. (16.12), we have

H2
1 - 111i -

1'2 2:2 m*) . m;

(16.21)

(16.22)

isolated slab, where IIelll Lawrence-Doniach

cc; we obtain He211 = HII; i.e., we obtain the upper critical field of an o: tT; - T)112, which we refer to as 2D behavior. Thus as the temperature is lowered a 3D to 2D 'crossover' occurs. Note that the divergence encountered with the
---7

From Eq. (16.22), as H

superlattice

model

involving

infinitesimally

thin superconducting

layers is

avoided in the Deutcher-Entin-Wohlman

model where d # 0.

Vortex structure in layered superconductors

Dichalcogenides ' of transition metals such as NbSe2' superconductor/insulator supcrlattices, and high T; oxide superconductors all have a layered structure. Most of these systems arc Type II superconductors with relatively large I( values. One of the fascinating characteristics of the layered superconductors is their strongly anisotropic magnetic properties. Usually, the coherence length perpendicular to the layer plane ((J is much smaller than that parallel to the layer (SI)' The anisotropy can be characterized by an anisotropy ratio, y, defined as )' = ~II/¢ L' Other length scales of interest are the penetration length, A, and the scale of intrinsic inhomogeneities. Depending on the relative size of ~j_ and the layer spacing, s, we may identify three different regimes for the vortex structure in the layered, high-x superconductors: (1) If ¢j_(T) > > s, then the layered structure is largely irrelevant and the superconductor may be regarded as threedimensional: anisotropic, but uniform. Since the coherence length diverges as T, is approached, this regime will always occur sufficiently close to T" The vortex structure can be described using the London or G-L theories by introducing an anisotropic mass tensor as discussed in Sec. 11.(2) With decreasing temperature, we may have a regime where ¢J(T) < < s (especially in high T, oxides). If both regimes can be entered by sweeping the temperature a 3D-2D crossover will occur at some temperature. Below this temperature we have a quasi-2D regime; the layered structure is then relevant and the Lawrence-Doniach-Iike model discussed in Sec. 16 is adequate to describe the vortex structure. (3) For the case of extreme anisotropy, as in Bi- and Tl-based oxide superconductors, with }' <: 10, the interlayer Josephson coupling is very weak. The flux 'lines' are then better viewed as stacks of 2D vortices residing in the superconducting layers (so-called pancake vortices). In this section we will give a brief account of the vortex structure in each of these three regimes. We will focus on the weak field region and derive expressions for Hel, which contain information on the anisotropy in addition to that contained in Hc2'

17.1

3D anisotropic London model

As discussed in Sec. 11, the upper critical field, Hd, in layered Type II superconductors can be described by the G-L equations with a phenomenological anisotropic mass tensor. Due to the linearity of the equations near Helo a relatively simple solution can be found in this region (see
1. The chalcogenides are the group VI elements S, Te, and Se. 82

Vortex structure in layered superconductors

83

Sec. 11). However, close to the lower critical field, Hc1, it is difficult to solve the nonlinear G-L equations to obtain Hcl in the anisotropic case. Furthermore, the G-L expansion itself is problematic in this regime since we are far from T'; For this reason, we will use the less accurate London model, which provides a reasonable description, at least for a large G-L parameter K (see,for example, Kogan (1981)). For K( == AR) > > 1, the amplitude of the order parameter, VI I, variesconsiderably only for distances smaller than ~; one may then assume 11/1 1 to be constant everywhere except in a narrow core of radius ~. As in Sec. 11, we introduce a phenomenological effective mass tensor. In the reference frame aligned with the principal axes of a crystal with orthorhombic symmetry or higher, this mass tensor is diagonal with elements which we now write as Mj• For convenience we define a geometric mean mass M = (M 1M 2M 3)1/3 and normalize the mass tensor such that mj = MJ M. In terms of M we define a mean penetration depth, A, and mean coherence Iength.z, by using Eqs. (2.6) and (9.22), respectively. The penetration depths, )'j = Aml/2, describe the decay of the components of the screening supercurrent along the principal directions, i. The coherence lengths, ~j = ~/mj1/2, characterize the spatial variation of the order parameter along these directions.For a uniaxial crystal, m3 = 111j_, = m2 = mil (in high T, oxides, the anisotropy within the m1 Cu-O layers is very weak). The anisotropy ratio is defined as :! ~ ~ II/~j_= ;'1/ All = (M ljMII)1!2. The free energy, given in Eq. (2.8) for the isotropic case, may be generalized to the anisotropic case as
1

(17.1)

whereB is the local field. Straightforward London equation B

minimization of the energy (17.1) yields the anisotropic

+ A2V

x [m- (V x B)]

O.

(17.2)

Taking into account the boundary condition, the local magnetic field around a single flux line directedalong and carrying a flux quantum CPo is determined by

+ A 2V

x (m V x B) = CPozb(x)b(y).

(17.3)

In the isotropic case where mij = bij, Eq. (17.3) coincides with the usual London equation (2.14). Todeal with an arbitrarily directed flux line, we define two reference frames as shown in Fig. 17.1. Thevortex frame (x, y, z) is obtained from the crystal frame (xo, Yo, 1.0) by rotating by an angle e fromZoabout Yo axis. Restricting ourselves to the case of uniaxial symmetry, the components mij' transformed to the vortex frame, are
ml
COS2

e + m3 sin? () e

0
1111

m(8) =
(

0 (m1 - 1113)sin cos 0

(17.4)

The algebra is greatly simplified by working in the vortex frame, since both Band j are independent of z. Substituting (17.4) into Eq. (17.3) results in a set of equations for Bx, By, and B; Theseequations lead to a more complicated vortex structure in the anisotropic superconductor than the isotropic Abrikosov vortex. For example, transverse fields, Bx•y, are present even when

84

Part I

Phenomenological

theories

Figure 17.1 In the crystal lrame (xo, Yo, 7.0) the plane xoYo coincides with the basal crystal plane. The vortex frame (x, y, z) is obtained lrorn the crystal frame by rotating by an angle 0 from 1.0 about the Yo axis. The axis 1. is parallel to the vortex axis. the applied magnetic F or
mxx

field is along the z direction. field is applied the perpendicular field


H13•

For simplicity, to or parallel case,

we consider The

two special cases in which the mass tensor is diagonal;

to the layer plane.

perpendicular

() = O.

myy

m1,H1zz

=
B, -}.

Eq. (17.3) then becomes

(a Bz
2 ml-

AX

.. 2 ~

Bz) + ml- aoy .. ~2


2

= <Po()(x)6(y).

(17.5)

This equation

is equivalent

to Eq. (7.S) and the lower critical field is given by (S.4b) as

Hclj_

:::::: --,

<Po

4nJ~

. tn -;:-.
C;

(A )
=

(l7.6) to the layer plane,

When the applied also diagonal:


mxx

magnetic

1113, myy

zz

field is parallel
mi'

e = n12. The

mass tensor is

Eq. (17.3) now reads


(17.7a)

LC.;

(17.7b) By interchanging the Zo <-> x, this equation can be transformed back to the crystal frame with B along

Xo

axis:

(17.8)
Eq. (17.S) can be solved by a Fourier the isotropic case by defining (17.9) From Sec. 7 we know that the solution of Eq. (17.S) at large distances is given by (17.10) where Kn(fj) is a modified Bessel function of the second kind of order n. The free energy per unit transform method. Alternatively, we may map Eq, (17.8) to

Vortex structure in layered superconductors

85

length of the flux line, Ep can be calculated

from Eq. (17.1). To logarithmic

accuracy

it is given by

(17.11)

Note for this case thermodynamic

the normal

core

radius

~j_ serves

as the cut-off

scale.

From

the usual

relation

E1 = CPoHcl/4n,

the parallel

lower critical

field is

(17.12)

Comparing can be obtained

Eqs. (17.6) and (17.12), we see that important by measuring

information fields.

about

the anisotropy

H c 1 in both parallel and perpendicular

17.2

Lawrence-Doniach

model
which can be modeled to a value T* such that is lowered

In Sec. 16 we discussed (,{'P)

the upper critical fields for layered superconductors

asJosephson lattices in 10. We saw that whenthe temperature

anisotropic

a 30-20 crossover occurs. Above T*, the layered superconductor behaves as an 30 superconductor. At temperatures lower than 1'*, the barriers of the layered structure dominate, resulting in a 20 behavior of the temperature dependence of He211 (T). In this subsection we will show that below drastically modified. Wc will start with Eqs. (16.4) and (16.5) which were derived freeenergy functional London approximation. (16.5)we obtain (16.2). For high tc supercond We confine ourselves thefield on the value of 11/1 n I can be neglected, and we can regard to the parallel
=

s/J2,

1'* the vortex

structure

in layered

superconductors

is also

from the Lawrence-Doniach field, the influence of

uctors in a weak magnetic

11/1" I as a cons tan t; i.e., we use the


field case. We take the field along

the x direction and choose a gauge such that Az

O. Writing

1/1"

I VI" I ei<D", from Eqs. (16.4) and

4nA~. Aj_ = - --h


c and

+ -V<Pn

CPo 2n

(17.13)

(17.14) where )'L is the bulk London the nth superconducting isotropic with an intrinsic Fig. 17.2 to compute wehave penetration depth of the superconducting layers, <Pn is the phase in Josephson supercurrent layer is

layer and

i; ~ e*h

11/1 12 /m;s is the maximum


for simplicity,

density. Note that in this model it is assumed, the phase difference

that each superconducting


1-

bulk penetration depth AL• We use the rectangular across one unit cell: <P == <D"+

contour C shown in (Dn. From Eq. (17.13)

86

Part I

Phenomenological theories

Figure 17.2 Schematic of a Josephson-coupled superconductor insulator superlattice. The insulating layers of thickness d, alternate with superconducting layers (crosshatched) of thickness d.. The modulation wavelength s = d, + d.. The rectangular contour C is used to compute the magnetic flux in Eq. (17.15).

y + "y
y

dy'

(n
ns

1),

dz'B(y',z')=

~
(

"Aj_·dl

+ CPo [<D(y) 2n
The left-hand side in the above equation can be approximated
sLiyB(y, z) ~ --

- (D(y

A + tly)].

as sLiyB(y,z),

and we obtain

41TA£
c'

(jy n+

1-

i. n)Liy + jO

CPo
2n

[<D(y) - (D(y

+ Liy)].

Hence (17.I~ where we have following approximations: ojjoz ~ Uy,n+ 1 - .iy,n)!s and For small <D the Josephson current relation (17.14) may be as j, ~ jmcD and substituting this form into Eq. (17.15) we obtain used the
- cD(y)]/Liy,

c<DjDy ~ [<D(y

+ L1y)
-

approximated

4nA£ ojv
C

~ --

oz

4n).; ojz
C

~=
oy

B(y,z),

(17.1~

You might also like