• No results found

Josephson junction thermodynamics and the superconductivity phase transition in a SQUID device - 25608y

N/A
N/A
Protected

Academic year: 2021

Share "Josephson junction thermodynamics and the superconductivity phase transition in a SQUID device - 25608y"

Copied!
5
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

UvA-DARE is a service provided by the library of the University of Amsterdam (https://dare.uva.nl)

UvA-DARE (Digital Academic Repository)

Josephson junction thermodynamics and the superconductivity phase transition

in a SQUID device

Maassen van den Brink, A.; Dekker, H.

DOI

10.1103/PhysRevB.55.R8697

Publication date

1997

Published in

Physical Review. B, Condensed Matter

Link to publication

Citation for published version (APA):

Maassen van den Brink, A., & Dekker, H. (1997). Josephson junction thermodynamics and

the superconductivity phase transition in a SQUID device. Physical Review. B, Condensed

Matter, 55, R8697-R8700. https://doi.org/10.1103/PhysRevB.55.R8697

General rights

It is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), other than for strictly personal, individual use, unless the work is under an open content license (like Creative Commons).

Disclaimer/Complaints regulations

If you believe that digital publication of certain material infringes any of your rights or (privacy) interests, please let the Library know, stating your reasons. In case of a legitimate complaint, the Library will make the material inaccessible and/or remove it from the website. Please Ask the Library: https://uba.uva.nl/en/contact, or a letter to: Library of the University of Amsterdam, Secretariat, Singel 425, 1012 WP Amsterdam, The Netherlands. You will be contacted as soon as possible.

(2)

Josephson-junction thermodynamics and the superconducting phase transition

in a SQUID device

Alec Maassen van den Brink*

Institute for Theoretical Physics, University of Amsterdam, Valckenierstraat 65, 1018 XE Amsterdam, The Netherlands and Department of Physics, The Chinese University of Hong Kong, Shatin, New Territories, Hong Kong

H. Dekker

Institute for Theoretical Physics, University of Amsterdam, Valckenierstraat 65, 1018 XE Amsterdam, The Netherlands and TNO Physics and Electronics Laboratory, P.O. Box 96864, 2509 JG Den Haag, The Netherlands

~Received 18 October 1996!

In a model of two ideal BCS superconductors coupled by a tunneling Hamiltonian the nonvanishing of the Josephson internal energy~and entropy! for T→Tc2is shown to be a consequence of superconducting corre-lations, which persist in the thermodynamic limit even in the mean-field approximation. The ensuing rapid increase of the Josephson free energy as the temperature of a tunneling junction drops below the supercon-ducting bulk transition temperature Tcmakes this transition of first order whenever the phase difference across

the junction is fixed to a nonzero value. Taking this into account results in an availability potential governing the nonequilibrium thermodynamics of the junction which, in contrast with previously published results, has no unphysical features like latent heat released upon entering~or a superconducting phase dependent value in! the normal state. The analysis inter alia predicts a lowering of the critical temperature~to Tc8) for the junction,

which has meanwhile been observed in high-quality superconducting quantum interference devices. @S0163-1829~97!51714-0#

I. INTRODUCTION

The Josephson energy GJ(f,T) has the thermodynamic

significance of a ~Gibbs! free energy, for changes

dGJ(f,T) are equal to electrical work V(t)IJ(f,T) dt done

at constant temperature. Given an expression for GJ(f,T),

one should thus be able to calculate the associated

en-tropy SJ(f,T)52]TGJ(f,T)uf and internal energy

UJ5GJ1TSJ. These quantities play an important role in the

nonequilibrium thermodynamic theory of Josephson

devices.1–4

The Josephson entropy does not vanish as the temperature

of the system approaches the bulk critical temperature Tcof

the superconducting phase transition. Indeed, for one of the

simplest formulas for GJ @Eq. ~13!# this entropy tends to a

constant value, as the gapD(T)}ATc2T if T→Tc2. SJ

van-ishes for T.Tc, and therefore it has a finite jump dSJ at

T5Tc. The case is even more dramatic for a junction

be-tween unequal superconductors a and b @we take

(Tc)a,(Tc)b#, where the phase-dependent part of GJ(f,T)

is proportional to DaDb when Da!T, and hence SJ even

diverges if T→(Tc)a2.

While an entropy jump in itself is familiar from first-order transitions, presently S is greater in the low- than in the

high-temperature phase, for ]D(T)/]T,0. Hence, latent

heat would be released upon entering the normal state, im-plying an unphysical instability.

The scenario can be studied in a superconducting

quan-tum interference device ~SQUID!1–3 by incorporating the

junction in a superconducting ring. Indeed, the phase differ-ence f52pF/F0 ~with F05h/2e) is externally controlled

in a ring with negligible self-inductance so that the flux F

equalsFext. We explain the mentioned instability, and show

that for fÞ0 the Josephson coupling lowers the actual Tc,

while the transition becomes first instead of second order. In Sec. II we investigate the nonvanishing of the

Joseph-son internal energy for T→Tc2 in a model of two BCS

su-perconductors coupled by a tunneling Hamiltonian. In Sec. III we explain the effect of the Josephson coupling on the superconducting transition, and examine the consequences for a theory of nonisothermal flux dynamics. The predicted

decrease in Tcis estimated in Sec. IV. A detailed account is

available in Ref. 5.

II. MICROSCOPIC THEORY

The system is modeled by the Hamiltonian

H5H01HT5Ha1Hb1HT, ~1! with (c5a,b; Vc.0) Hc5Hc,kin1Hc,int 5

(

pa jp cpacpa2Vc

(

pp8

8

cpc2p↓c2p8cp8↑, ~2! HT5

(

pqa Tpq apa † bqa1H.c. ~3!

We work in the grand canonical ensemble and the term

2mNcis included in Hc, i.e.,jp5p2/2m2m. The prime on

the sum implies restriction to p(

8

) withuj

p(8)u,vD, the

De-bye frequency. Neglecting the cutoff amounts to taking the

55

(3)

limit Dc,T!vD!m ~throughout \5kB51) and is allowed

in all steps except when using the BCS gap equation to arrive

at the last line of Eq.~18!.

To introduce the Green’s functions Gc of the uncoupled

superconductors we define the Nambu field as6

Cc~p,t!5„cp~t!,2c¯2p↓~t!…T, ~4! so that Gc~p,t!52

^

Tt$Cc~p,t!C¯c~2p!%

&

0 ~5! 5

S

FGc~p,t! Fc~p,t! c 1~p,t! 2G c~2p,2t!

D

~6! yields Gc~p,v!5 1 v21j p 21D c 2

S

2iv2jp Dceifc Dce2ifc 2iv1jp

D

, ~7!

Dcbeing the magnitude of the gap.

In Nambu notation the tunneling Hamiltonian reads

HT5

(

pq C ¯ a~2p!TpqCb~q!1H.c., ~8! where Tpq5diag~Tpq,2T2p2q* !. ~9!

Evaluating the change in grand canonical potential due to tunneling in lowest-order perturbation theory one finds

dG[G~$Tpq%!2G~$Tpq50%! 5212T

E

0 b dt1

E

0 b dt2

^

Tt$HT~t1!HT~t2!%

&

0 5T

E

0 b dt1

E

0 b dt2 3

(

pq Tr$Ga~p,t22t1!TpqGb~q,t12t2!Tpq% 5T

(

pqv Tr$Ga~p,v!TpqGb~q,v!Tpq † %, ~10!

where we used Eq. ~8! for HTand applied Wick’s theorem.

To evaluate Eq. ~10!, we assume that the tunneling

ampli-tudes obey time-reversal symmetry Tpq5Tpqs3 and are

en-ergy independent~on the scale of Tc). Decomposing the sum

over p as (p5Na(0)*2`1`djp*(dpˆ/4p) @where the

approxi-mation Na(jp)'Na(0) is correct to first order in (Tc)a/m#,

the integrals over jp,q can be performed with Ga,b as in

Eq. ~7!. This yields

dG522Tg

(

v Q P

21, ~11!

with P[

A

v21Da2

A

v21Db2, Q[v21DaDbcosf

(f[fa2fb), and the dimensionless conductance

g5p2

^

uT

pqu2

&

Na(0)Nb(0)5p/(4e2RN), the relation

be-tween the tunneling amplitudes and the normal-state

resis-tance RNbeing standard.7

The sum in Eq. ~11! diverges fordG, but converges for

its difference between the superconducting and normal states

dGS2N[dGS2dGN, which is the quantity of interest. A

general property is its nonnegativity: dGS2N

}(v(12Q/P)>0 because

P22Q2>~v21Da2!~v21Db2!2~v21DaDb!2

5v2~D

a2Db!2>0. ~12!

For equal gapsDa5Db5D one finds ~see Ref. 7!

GJ[dGS2N52TgD2~12cosf!

(

v ~v

21D2!21

5gDtanh~D/2T!~12cosf!, ~13!

consistent with Anderson’s theorem,8 implying that

GJ(f50)50 ~to lowest order in Tc/m), which will be

cru-cial in Sec. III.

For unequal gapsDa,b!T one obtains

GJ5 g 4T~Da 21D b 222D aDbcosf!1O@~D/T! 4#, ~14!

while for two superconductors with very different Tcnear the

lower of these temperatures, i.e., for Da!T!Db, the result

reads5(gEis Euler’s constant!

g21G J5 2 pDb2 pT2 3Db 1p1

H

Da 2 Db 22Dacosf

JH

ln

S

4Db pT

D

1gE

J

1T$O@~T/Db!3#1O@TDa/Db 2 #1O@~Da/T!3#%. ~15!

Since Da,b}A(Tc)a,b2T for T→(Tc)a,b2 , with

U5]b(bG) the above formulas for GJ predict a constant @Eq. ~13!# or even divergent @Eqs. ~14! and ~15!# internal

energy ~and entropy! difference dUS2N upon approaching

the ~lowest! Tcfrom below. This property is a consequence

of deviations from BCS mean-field theory ~even though

these superconducting correlations are of order Na,b21), as is

demonstrated by an explicit calculation ofdU in which such

correlations are neglected. Consider

dU5d

^

Ha

&

1d

^

Hb

&

1

^

HT

&

, ~16!

where d

^

&

5

^

&

g2

^

&

0. In first order

^

HT

&

52dG, with

dG as in Eq. ~10!. For

^

Ha

&

one has

^

Ha

&

5

(

pa jp

^

apaapa

&

2Va

(

pp8

8 ^

apa2p↓a2p8ap8

&

. ~17!

Expanding the second term in terms of two-point functions,

the Hartree-Fock contribution to d

^

Ha

&

MF vanishes in the

thermodynamic limit. One is left with

(4)

^

Ha

&

MF5

(

pa jp

^

apaapa

&

2Va

(

pp8

8 ^

apa2p↓

&^

a2p8ap8

&

⇒ d

^

Ha

&

MF5

(

pa jpd

^

apaapa

&

2Va

(

pp8

8

2 Re@Fa~p

8

,t50!d

^

a2p↓ap

&

# 5

(

pa jpd

^

apaapa

&

2Da

(

p 2 Re@eifad

^

a 2p↓a p↑ †

&

# 5

(

p Tr$~Daeifas3s12jps3!d

^

Ca~p!C¯a~2p!

&

%, ~18!

where we usedd

^

ap(†)a2p↓(†)

&

5O@Na21# so that only the linear

term contributes tod

^

Ha

&

. To evaluate the rhs of Eq.~18! in

lowest order transform to frequencies, and use

Daeifas3s12jps35Ga(p,v)212iv to arrive at d

^

Ha

&

MF5T

(

pqv Tr$@iv2Ga~p,v!21# 3Ga~p,v!TpqGb~q,v!TpqG a~p,v!% 52T

(

pqv v Tr$]vGa~p,v!…TpqGb~q,v!Tpq% 21 2

^

HT

&

, ~19!

d

^

Hb

&

MF following by symmetry. Substitution into Eq.~16!

shows that

^

HT

&

cancels indUMF, viz.,

dUMF52T

(

pqv v ]v Tr$Ga~p,v!TpqGb~q,v!Tpq% 5]buDa,bb

F

Tpq

(

v Tr$Ga~p,v!TpqGb~q,v!Tpq † %

G

5]b~bdG!Da,b. ~20!

As our derivation is valid in the normal~N! and the

super-conducting ~S! phase, this relation also holds between

dUMF S2N

anddGS2N. By Eqs.~13!–~15! and ~20!,dUMF

S2N

van-ishes as T→Tc2, which is what we set out to show.

III. THERMODYNAMICS

The free energy of the system~with respect to the N state!

also contains the phase ~or SQUID flux! independent bulk

condensation energy, which is quadratic in Tc2T near Tc. It

usually completely dominates the Josephson term GJ@>0 by

virtue of Eq.~12!#. Accounting for the bulk energy is crucial

for resolving the thermodynamic instability of the junction

below Tc~see Sec. I!.

For a fixed fÞ0, GS2N varies with temperature as in

Fig. 1. Just below Tcthe total free energy in the S phase is

higher than in the N phase, so that the system remains in the

state D50 ~always a solution of the gap equation! until

GS2N50 again, at Tc

8

@see Eq. ~24!#. At this Tc

8

,Tc, D

jumps to a finite value. This first-order transition is indeed

accompanied by latent heat~see Sec. I!, as follows from the

kink in G(T). The significance of Anderson’s theorem ~see

Sec. II! is that the transition remains of second order if no

flux is applied externally.

Now consider the nonequilibrium thermodynamics of the

system~see Refs. 1–3!. Equation ~19! of Ref. 3 and the one

below it yield A52EJcos

S

2p F F0

DH

123T ˜2T r22T˜3 Tc3

J

1 ~F2Fext!2 2L 1a 1 3T 321 2 T 2T r Tc3 ~21!

for the availability A5G1(T2Tr)S of a SQUID in contact

with a bath at temperature Tr. Here EJ5Ic(T50)/2e, a

arises in a Debye-type bulk free energy Gbulk

521

6a(T/Tc)3, and T˜[min(T,Tc).

Following Eq. ~21!, an availability barrier persists in the

N state nearF512F0. However, this barrier is entirely due to

the Josephson coupling and related to the unphysical nega-tive latent heat discussed in Sec. I. This artifact is remedied by accounting for the bulk condensation energy, which we

now model a` la Refs. 1–3. Since Eq. ~21! implies Sbulk

51 2aT 2/T c 3, we take5S bulk S 51 2(a/Tc)(T/Tc)21h (h.0). For

the S2 phase one obtains

FIG. 1. The Josephson, bulk and total free energy of a junction withfÞ0 ~arbitrary units!. The dashed line indicates the free en-ergy of the superconducting phase, the thick line the equilibrium free energy G(T).

(5)

AS5EJ

F

12cos

S

2p F F0

DGF

123T 2T r22T3 Tc3

G

1 ~F2Fext!2 2L 1a

H

2~311 h! 2161

S

1222~311 h!

DS

TT c

D

31h 212

S

TT c

D

21hT r Tc

J

, ~22!

while AN follows from Eq. ~22! with EJ5h50. The free

energy GS,N results from AS,N by setting Tr5T. Our final

result for A(T) then follows from9

A~T!5$u~T2Tc!1u~Tc2T!u„GS~T!2GN~T!…%AN~T! 1u~Tc2T!u„GN~T!2GS~T!…AS~T!. ~23!

A calculation of the junction’s actual critical temperature Tc

8

~see above in this section! is now easy in the limit

EJ!a, in which the effect of the Josephson coupling is small

and hence «[(Tc2Tc

8

)/Tc!1. Expansion of GS5GN in

EJ/a!1 yields «512EJ ah

F

12cos

S

2p F F0

DG

. ~24!

IV. FINAL REMARKS

The thermodynamic analysis of Sec. III can also be mo-tivated in the context of a dynamical theory of nonisothermal behavior, which has potential relevance for a wide class of

physical systems.4The relevant ‘‘reaction coordinate’’~here

the flux! and the temperature then become fluctuating

quan-tities, the values of which spread over the entire accessible state space. Hence, there is no a priori possibility to restrict

the variables to a bounded region such as T,Tc ~or

T.Tc), so that covering the phase transition becomes

cru-cial.

Of course, the phenomenology of Sec. III is no substitute

for solving Gor’kov equations for a finite-size SQUID. Also,

the model~1! obscures the fact that tunneling is an interface

phenomenon. However, this should allow a

Ginzburg-Landau description near Tc, e.g., to determine the effective

interaction volume V ~see below!.

To estimate Tc

8

, take F512F0 and

EJ5hD(0)/8e2RN'10219/RNJ ~for Al! in Eq. ~24!. From

Eq. ~22! we find GbulkN2S51

4ah(12T/Tc)2, which yields

ah'6VBc(0)2/m0 upon comparison with Gbulk

N2S'

(V/2m0)$1.73Bc(0)(12T/Tc)%2@Ref. 6, Eq. ~36.12!#, where

m054p1027 H m21 and Bc(0)'1022 T @for Al ~Ref.

10!#, so that «'5310221/VRN. For V50.1 mm3 and RN550 V, one arrives at «'1023. With Tc'1 K this means that the effect is in the mK range, and hence observ-able in state-of-the-art devices of sufficiently small

dimen-sions and of a high material quality such that Tcis sharply

defined.

In conclusion, this work highlights hitherto neglected con-sequences of the phase-independent part of the Josephson

coupling @the first term in Eq. ~13!#, which near Tcis not an

arbitrary constant. In the meantime, its predictions

concern-ing the oscillations of Tcversus the applied magnetic flux in

a SQUID device have indeed been confirmed

experimentally.11

ACKNOWLEDGMENTS

We thank R. de Bruyn Ouboter and A.N. Omel’yanchuk for their interest and discussions.

*Electronic address: alec@phy.cuhk.edu.hk. Current address in

Hong Kong.

1R. de Bruyn Ouboter, Physica B 154, 42~1988!.

2R. de Bruyn Ouboter and E. de Wolff, Physica B 159, 234~1989!. 3E. de Wolff and R. de Bruyn Ouboter, Physica B 168, 67~1991!. 4H. Dekker, Physica A 173, 381~1991!; 173, 411 ~1991!; Phys.

Rev. A 43, 4224~1991!.

5

A. Maassen van den Brink and H. Dekker, Physica A 237, 471 ~1997!; A. Maassen van den Brink, Ph.D. thesis, University of Amsterdam, 1996.

6A.A. Abrikosov, L.P. Gor’kov, and I.E. Dzyaloshinski, Methods

of Quantum Field Theory in Statistical Physics ~Dover, New York, 1975!.

7I.O. Kulik and I.K. Yanson, The Josephson Effect in

Supercon-ductive Tunneling Structures ~Israel Program for Scientific Translations, Jerusalem, 1972!.

8G. Rickayzen, Theory of Superconductivity ~Wiley, New York,

1965!.

9The relative stability of phases can also be determined in terms of

the availability~instead of G), but then A has to be considered as a function of the enthalpy and the analysis becomes more involved; see further Ref. 5.

10Handbook of Chemistry and Physics, edited by D.R. Lide~CRC

Press, Boca Raton, 1992!.

11B.J. Vleeming, M.S.P. Andriesse, A. Maassen van den Brink, H.

Dekker, and R. de Bruyn Ouboter~unpublished!.

Referenties

GERELATEERDE DOCUMENTEN

We discuss thc analogy betwccn thc axial mode spcctium of an optical resonator with one or two phase-conjugating mirrors, and the quasipaiticle excitation spectium ot an NS or

Mean (± SE) weight change of four groups of ten worms (Eisenia fetida) each exposed to different concentrations of zinc (mg/kg) in saline and non saline OECD

It consists of a chaotic cavity in a time-reversal- symmetry breaking magnetic field B, which is coupled to two superconductors and to one or two normal metals by ballistic

We consider the case when the normal region has a chaotic classical dynamics on time scales greater than the ergodic time T crg.. In this section, we assume that r erg « Ä/Δ, so

This signifies that on average there is no change in the phase difference over the junction along the different directions in the plane of the junction, presumably due to a

The question of whether mesoscopic fluctuation can re- verse the supercorrent through a disorderd Josephson junc- tion was posed ten years ago by Spivak and Kivelson, 1 m their

Het antwoord op deze laatste vraag hangt samen met de waarde, die we aan het meetkundeonderwijs toekennen. Het zal iedereen wel duidelijk zijn, dat er eerst een antwoord op de

Some time ago it was realized that Andreev reflection has a similar effect on the chaotic motion in an electron billiard coupled to a supercon- ductor [5].. An electron trajectory