• Sonuç bulunamadı

Weak-localization effect in superconductors from radiation damage

N/A
N/A
Protected

Academic year: 2021

Share "Weak-localization effect in superconductors from radiation damage"

Copied!
9
0
0

Yükleniyor.... (view fulltext now)

Tam metin

(1)

Weak-localization effect in superconductors from radiation damage

Mi-Ae Park

Department of Physics, University of Puerto Rico at Humacao, Humacao, Puerto Rico 00791

Yong-Jihn Kim

Department of Physics, Bilkent University, 06533 Bilkent, Ankara, Turkey 共Received 16 September 1999; revised manuscript received 3 November 1999兲

Large reductions of the superconducting transition temperature Tcand the accompanying loss of the thermal electrical resistivity共electron-phonon interaction兲 due to radiation damage have been observed for several A15 compounds, Chevrel phase and ternary superconductors, and NbSe2in the high-fluence regime. We examine

these behaviors based on a recent theory of the weak localization effect in superconductors. We find a good fitting to the experimental data. In particular, the weak localization correction to the phonon-mediated inter-action is derived from the density correlation function. It is shown that weak localization has a strong influence on both the phonon-mediated interaction and the electron-phonon interaction, which leads to the universal correlation of Tc and the resistance ratio.

I. INTRODUCTION

Much attention has been paid to experimental and theo-retical investigations of the radiation effects on superconductors.1–3 For practical applications of the super-conductors in the magnet coils of a fusion reactor, the radia-tion response of the materials is important because they are subjected to irradiation. On the theoretical side, the disorder effects in superconductors caused by irradiation are interest-ing. The radiation effects in elemental type II superconduct-ors showed that the superconducting transition temperature Tc does not change significantly for relatively high-fluence

irradiations.3 The slight reduction was attributed to the re-duction of the gap anisotropy. Subsequent annealing leads to a partial recovery of the Tc changes, implying the

impor-tance of the microscopic details of the disorder structures. We note that most elemental type II superconductors are ra-diation tolerant. For instance, the He-4 dose which resulted in ⌬␳0⬃100␮⍀ cm in Nb3Ge causes ⌬␳0⬃2 ␮⍀ cm in Nb.4 Here ␳0 denotes the residual resistivity. On the other hand, A15 compounds show universal large reductions of Tc

and critical currents Ic for high-fluence irradiations.4–14The

residual resistivity␳0 also increases over⬃100␮⍀ cm, in-dicating that A15 compounds are radiation susceptible. Note that the layered compound NbSe2, ternary superconductors LuRh4B4 and ErRh4B4,15 and Chevrel phase superconductors,3 such as PbMo6S8, PbMo6S7, and SnMo5S8, also show the large Tc reduction in the

high-fluence regime.

In this paper we propose an explanation of the micro-scopic mechanism underlying the universal large reductions of Tc and Ic in A15 compounds and other materials. In

par-ticular, we stress the experimental observation of the corre-lation between the electrical-resistance ratio and Tc.

4,8,9,11 Testardi and co-workers4,8,9 considered Nb-Ge, V3Si, and V3Ge for a variety of samples produced with differing chemical composition, preparation conditions, and with varying amounts of 4He-induced defects. They found a close relation between the resistance ratio and Tcfor those samples

irrespective of the manner how the disorder was achieved. Furthermore, they found that decreasing Tc is accompanied by a decrease of the thermal electrical resistivity 共electron-phonon interaction兲.4 It was also reported that tunneling ex-periments in Nb3Ge and Nb-Sn clearly show a decrease of the electron-phonon coupling constant ␭ accompanying the decrease of Tcwith disorder.

16–18

However, previous theoretical studies focused not on changes of the electron-phonon interaction but on the smear-ing of the electronic density of states near the Fermi level N(EF), the microscopic details of the disorder, the gap

an-isotropy, and the enhancement of the Coulomb repulsion.19–25It is obvious that a consistent explanation of the existing experimental data was not possible in those prior theories. Recently, Kim and Overhauser26 pointed out that Anderson’s theorem27 is valid only to the first power in the impurity concentration and the phonon-mediated interaction decreases exponentially by Anderson localization, in agree-ment with the above experiagree-mental findings. As expected, it was shown that the same weak localization correction terms occur in both the conductivity and phonon-mediated interaction.28,29Based on the reduced phonon-mediated inter-action, we propose an explanation of the universal reductions of Tc and Ic and the universal correlation of Tc and

resis-tance ratio in Sec. III.30

Several comments are in order.共1兲 It is obvious that both impurity doping and irradiation共or implantation兲 can be used to study the disorder effects in superconductors and metals. In particular, compensation of the Tc reduction caused by

magnetic impurities has been observed as a consequence of both radiation damage and ordinary impurity doping.31–34 This compensation phenomenon has been predicted by Kim and Overhauser.35,36Recently, it has also been observed that impurity doping and/or ion-beam-induced damage in high-Tc superconductors cause a metal-insulator transition and thereby suppress Tc.37–40These reductions may also be

un-derstood by the weak localization effect on superconductors. The only difference is the strong renormalization of the im-purity potential due to the strong electron-electron

interac-PRB 61

(2)

tion in high-Tc superconductors. 共2兲 Although Anderson’s

theorem suggests no change of the electron-phonon interac-tion due to the very strong disorder, there is overwhelming experimental evidence for a decrease of the electron-phonon interaction in the strongly disordered samples and in the high-fluence regime. For instance, tunneling,16–18 specific heat,41 x-ray photoemission spectroscopy 共XPS兲,42 loss of thermal electrical resistivity,4 and the correlation of Tc and

the resistance ratio4,8,9,11reveal the decrease of the electron-phonon interaction when the electrons are weakly localized. Weak localization leads to a decrease of the amplitude of the electron wave function. As a result, the phonon-mediated matrix elements are also decreasing.28,29共3兲 Tunneling data do not show any enhancement of the Coulomb repulsion.16–18In addition, the loss of the thermal electrical resistivity with decreasing Tc and the universal correlation

between Tcand resistance ratio cannot be explained in terms

of the increase of the Coulomb interaction. 共4兲 Irradiation also leads to strong Tcreductions in Chevrel phase materials,

such as PbMo6S8,PbMo6S7, and SnMo5S8,3,43,44 and NbSe2,

43

at fluences above⬃1018 neutron/cm2. These mate-rials are more radiation sensitive than A15 compounds. It is clear that the origin of the strong Tc reduction is not related to the microscopic details of the disorder, but related to the universal nature of the electronic state in the irradiated samples.

In Sec. II, we briefly review the experimental results of the radiation damage effects on A15 compounds. The univer-sal large reduction of Tc, the accompanying decrease of

thermal electrical resistivity, and the correlation of Tc and

the resistance ratio will be emphasized. In Sec. III, the weak localization correction on the phonon-mediated interaction is derived. The resulting Tc decrease will be compared with

experiments in Sec. IV.

II. RADIATION EFFECTS IN A15 COMPOUND SUPERCONDUCTORS: UNIVERSAL TcREDUCTION

AND RESISTANCE RATIO

Several A15 compounds have been investigated, includ-ing Nb3Sn,5–7,11–13,20, Nb3Al,6,7,10, Nb3Ge,4,6–8,11,13,

Nb3Ga6,7, Nb3Pt,24, V3Si,4,7,11, and V3Ge.4,11. Both high-energy neutron5,6,10 and other energetic charged particles, such as protons,14 ␣ particles,4,8,11,13 oxygens,12 and electrons,20 were used to irradiate a variety of A15 com-pounds. Table I summarizes the experimental results of the irradiation effects on A15 compound superconductors, Chev-rel phases, and NbSe2. Note that the large Tcreductions are

found in not only A15 compound superconductors but also Chevrel phase superconductors, ternary superconductors, and NbSe2, implying the universality of the phenomenon.

The response of the superconducting properties of A15 compounds to irradiation can be classified into behavior at low fluences and at higher fluences. In the low-fluence re-gime, little or no change in Tc occurs, while universal large

reductions of Tcare observed for higher fluences. We focus

on the universal Tc reduction in this paper. The boundary between the two regimes depends on the irradiating particles, since heavy ions give rise to more severe radiation damage. For instance, the low-fluence regime corresponds to neutron fluence ⬍⬃1018 neutron/cm2 and to 4He fluence ⬍⬃1015 4He/cm2. For much more higher fluences the satu-ration of Tcis often found. It is noteworthy that the saturated Tc state is accompanied by a saturated value of the residual

resistivity ␳0.4 Accordingly, classification based on the re-sidual resistivity共not the fluence兲 may be more appropriate. In terms of the residual resistivity, the low-fluence regime corresponds to␳0⬍⬃10␮⍀ cm irrespective of the irradiat-ing particles. The saturations of Tc and the residual

resistiv-ity are easily understood in this classification scheme. From Table I, it is clear that the universal Tcreduction is not crucially dependent on any specific irradiating particle, any specific material, and any specific defect. Many physi-cists noticed that the universal Tc reduction is governed by

the total residual resistivity ␳0 due to the radiation damage and the inherent damage present in the sample.45,4,8,9,11 Fur-thermore, the close relation between the Tc decrease and resistance ratio was established in Nb-Ge, V3Si, and V3Ge.4,8,9,11 The relation was also noticed in Nb-O solid solutions.46 Testardi and co-workers4,8,9,11 reported that the correlation of Tcand the resistance ratio is independent of all TABLE I. Irradiation effects on A15 compounds, Chevrel phases, and NbSe2.

Sample Irradiating particle Tc0 ⌬Tc Maximum fluence Reference

Nb3Ge ␣ particle ⬃20 K ⬃8 K 1017␣/cm2 8 Neutron ⬃20 K ⬃16 K 5⫻1019neutron/cm2 3 Nb3Sn ␣ particle ⬃18 K ⬃15 K 7⫻1017␣/cm2 11 Neutron ⬃18 K ⬃7 K 2⫻1019neutron/cm2 5 Electron ⬃17.8 K ⬃3.8 K 4⫻1020electron/cm2 20 Nb3Al Neutron ⬃18 K ⬃14 K 5⫻1019neutron/cm2 6 Nb3Pt Neutron ⬃10.6 K ⬃8.4 K 3⫻1019neutron/cm2 24 V3Si ␣ particle 16.8 K ⬃14.5 K 7⫻1017␣/cm2 11 Neutron ⬃16.5 K ⬃13.5 K 2.5⫻1019neutron/cm2 3 V3Ge ␣ particle 6.5 K ⬃5.5 K 5⫻1017␣/cm2 11

PbMo6S8 Neutron 12.8 K ⬃8.6 K 1⫻1019neutron/cm2 3

PbMo6S7 Neutron 61% 1.5⫻1019neutron/cm2 3

SnMo5S8 Neutron 51% 1.5⫻1019neutron/cm2 3

(3)

sputtering conditions, film thickness, composition, and radia-tion damage. This result implys that defects produced during irradiation are similar in their effect on Tcto those produced

during the film growth process. Consequently, the correlation of Tc and the resistance ratio is also universal. Until the

resistance ratio is about 5, Tcdoes not change much. When it

is smaller than 2, Tcdrops quickly. Finally superconductivity

disappears if the resistance ratio is around 1. Testardi et al.4 also found that decreasing Tc is accompanied by the loss of

the thermal electrical resistivity 共electron-phonon interac-tion兲, which indicates the significant role of defects in both superconducting and normal-state behavior. This finding, consistent with the correlation of Tcand the resistance ratio, predicts the complete destruction of superconductivity for a resistance ratio less than 1 because of the complete loss of the electron-phonon interaction.

Other evidence for the decrease of a electron-phonon in-teraction in the high fluence regime is the following: 共1兲 By channeling measurements in single-crystal V3Si, Testardi et al.47 also found that radiation damage leads to large in-crease of the resistivity and a reduction of the electron-phonon interaction. 共2兲 Viswanathan and Caton48 reported the correlation of Tc and the residual resistivity in

neutron-irradiated V3Si. 共3兲 Tsuei, Molnar, and Coey41 did a com-parative study of the superconducting and normal-state prop-erties of the amorphous and crystalline phases of Nb3Ge. They found that the drastic reduction of Tc is due to the

changes in the strength of the electron-phonon interaction. Pollak, Tsuei, and Johnson42 did an XPS study of the crys-talline and amorphous phases of Nb3Ge and found that the crystalline phase has a higher Tcbecause of the enhancement

of the electron-phonon coupling.

The response of the critical current Icalso depends on the

fluence. For the low-fluence regime, Ic decreases first with

fluence, and then increases with increasing fluence, implying the importance of the flux-pinning mechanism.49,50 For higher fluences, the universal reduction of Tc drives down

Ic.5,51,52 The Ic drop is field independent.53

III. WEAK LOCALIZATION CORRECTION TO THE PHONON-MEDIATED INTERACTION

In the presence of an impurity potential U0, the Hamil-tonian is given by H

dr

␣ ⌿␣ †共r兲

p 2 2m⫹U0共r兲

⌿␣共r兲 ⫺V

共r兲⌿ 共r兲⌿ 共r兲⌿共r兲dr, 共1兲

where⌿†(r) and⌿(r) are creation and annihilation opera-tors for electrons. In terms of the exact scattered states

n(r), we expand the field operator⌿␣(r) as

⌿␣共r兲⫽

nn共r兲cn

, 共2兲

where cn␣ is a destruction operator of the electron. Upon

substituting Eq. 共2兲 into Eq. 共1兲, we find

H

nn cncn

nn,n⫽”nVnncnncn¯ ↓cn, 共3兲 where Vnn⫽V

n*⬘共r兲¯n*⬘共r兲¯n共r兲n共r兲dr. 共4兲

Here ⑀n is the normal-state eigenenergy and n¯ denotes the

time-reversed partner of the scattered state n. Equation 共4兲 was first obtained by Ma and Lee.54

A. Andersons’ theorem

By a unitary transformation between the scattered states and the plane-wave states,

n␣⫽

kជ

kជ

k兩n

, 共5兲

Equation共4兲 can be rewritten as26 Vnn⫽V

kជ,kជqជ

⫺k

兩n

典具

k兩n

*

k⫺q兩n

⬘典具

⫺k

⫺q兩n

⬘典

*.

共6兲 Anderson27assumed that the transformed BCS part in Eq.共6兲 plays a much more important role and each individual matrix element of the remaining interaction is so small as to be safely disregarded. Then the normalization condition of the scattered states leads to

Vnn⬵VnnBCS

⫽V

kជ,kជ

k兩n

兩2兩

n

兩k

⬘典

兩2⫽V, 共7兲 which is the essence of Anderson’s theorem.

However, the remaining term

Vnnnon-BCS ⫽V

kជ⫽”⫺kជ,kជqជ

⫺k

兩n

典具

k兩n

*

k⫺q兩n

⬘典

⫺k

⫺q兩n

⬘典

* 共8兲

cannot always be ignored. As Anderson suggested, the above term is indeed negligible in low-fluence regime, where re-sidual resistivity is smaller than 10␮⍀ cm. In fact, the low-fluence regime corresponds to the dirty limit where 1/kFl

⬍0.1. kF and l denote the Fermi wave vector and mean free

path, respectively, whereas for higher fluences, the remain-ing term contributes significantly. In this regime, the electron wave functions are weakly localized. Note that weak local-ization yields the well-known weak locallocal-ization correction to the conductivity.

Now we calculate Vnn⬘including both BCS and non-BCS

terms. In order to do this, we use Eq.共4兲 关not Eq. 共6兲兴 which is more physically transparent. In the dirty limit, the exact eigenstates␺n(r) can be approximated by the incoherent

su-perpositions of plane-wave states suggested by Thouless,55 which leads to the Boltzmann conductivity. The wave func-tion, with energy ប2kn2/2m, is written as

(4)

n共r兲⫽

kជ

akជneikជ•r. 共9兲

The amplitudes akជn are assumed to be independent normally distributed random variables with variance

akជn*akជn⬘ ⬵␦nn⬘␦kជkជ⬘ ␲ ⍀kn 2l 1 共k⫺kn兲2⫹1/4l2 , 共10兲

for large knl. Here⍀ denotes the volume of the system.

Inserting Eq.共9兲 into Eq. 共4兲 we obtain Vnn⫽V

兩␺n共r兲兩2兩␺n共r兲兩2dr ⫽V

kជ,qជ akជnaqជn*eikជ•re⫺iqជ•r

kជ,qជakជnaqជn⬘ *eikជ•re⫺iqជ•rdr ⬵V

kជ 兩akជ n 兩2

kជ兩akជn⬘ 兩2dr⫽V. 共11兲

共Here we assume ⍀⫽1 for the phonon-mediated matrix el-ement, in accordance with the usual notation.27,54兲 We have made use of Eq. 共10兲 which eliminated the non-BCS cross terms since n and n

are different54 and are not dummy in-dices. As Vkk does not contribute to superconductivity in

homogeneous systems,56 Vnn does not contribute to

super-conductivity in disordered systems.54Here Vkkis the

共forbid-den兲 diagonal term in the BCS reduced interaction. As a result, Anderson’s theorem is proved for large knl under this

assumption.

B. Weak localization correction

For the high-fluence regime, we may use the weakly lo-calized wave functions suggested by Kaveh and Mott.57,58 For three dimensions, the weakly localized wave functions consist of power-law and extended wave functions,

kជ共r兲⫽Aeikជ•r⫹B eikr r2 , 共12兲 where A2⫽1⫺4␲B2

1 l⫺ 1 L

, B 2 3 8␲ 1 kF2l. 共13兲 L denotes inelastic diffusion length. We then write an eigen-state ␺n as ␺n共r兲⫽

kជ akជnkជ共r兲⫽

kជ akជn

Aeikជ•r⫹Be ikr r2

. 共14兲 Comparing to Thouless’ wave function, Eq. 共9兲, Kaveh and Mott’s wave function includes the power-law component which originated from the diffusive motion of the electrons. While Thouless’ wave function corresponds to the Green’s function in the self-consistent Born approximation,59,60 Kaveh and Mott’s wave function corresponds to the Green’s function which includes both the Born scattering and the coherent backscattering due to impurities.61,58 Since the

power-law wave function 1/r2does not contribute to the cur-rent, the conductivity is reduced as

3d BA4⫽␴B

1⫺ 3 共kFl兲2

1⫺ l L

. 共15兲 A similar situation occurs in the phonon-mediated interac-tion. The power-law component does not contribute to the phonon-mediated matrix element either. The reason is the following: since the power-law component peaks at some point, its contribution to the bound state of Cooper pairs far from the point is almost negligible. This is analogous to the insensitivity of the localized共bound兲 state with the change of the boundary conditions.62 Accordingly, substitution of Eq. 共14兲 into Eq. 共4兲 leads to the weak localization correction to the phonon-mediated interaction,

Vnn⫽V

kជ 兩akជ n 兩2

kជ兩akជn⬘ 兩2 kជ共r兲兩2兩␺kជ共r兲兩2dr ⫽V

兩␺kជ共r兲兩2兩␺kជ共r兲兩2dr 共16兲 ⬵VA4⫽V

1 3 共kFl兲2

1⫺ l L

. 共17兲 We have made use of the fact that Eq.共16兲 does not depend on kor k

.

We can also derive the weak localization correction term in Eq.共17兲 without using Eqs. 共12兲–共14兲 based on the diffu-sive density correlation for the eigenstates. In order to do this, it is important to note that the matrix element Vnn

denotes the correlation function between two eigenstates␺n and␺n⬘, as is clear from the expression

63 Vnn⫽V

兩␺n共r兲兩2兩␺n共r兲兩2drV

qជ

n兩eiqជ•r兩␺n

兩2. 共18兲 We have evaluated the sum over qជ. For quantum diffusion of electrons, it was shown63–66that

n兩eiqជ•r兩␺n

AV 2 1 2␲បN0共EFDqជ2 共Dqជ22⫹共 n⫺⑀n⬘兲2/ប2 , 共19兲 where AV means the average over all states, and N0(EF) and

D are the density of states and the diffusion constant, respec-tively. This quantity is proportional to the spectral function for the density correlation function, A(q,␻), which is de-fined as66 A共qជ,␻兲⫽

⫺⬁ ⬁ dtdrdr

ei␻teiqជ•(r⫺r⬘)

关␳共r

,t兲,␳共r,0兲兴

, 共20兲 where␳ is the density operator. In Eq.共19兲 we may assume

n⫽⑀n⬘, since we are interested in states very near the Fermi

(5)

Note also that in the presence of impurities, the correla-tion funccorrela-tion has a free-particle form for t⬍␶ 共scattering time兲 and a diffusive form for t⬎␶.25As a result, for t⬎␶共or r⬎l), one finds63 R

t⬎␶兩␺n共r兲兩 2 n共r兲兩2dr⫽ 1 ⍀

qជ

n兩eiqជ•r兩␺n

AV 2 ⫽1

␲/L⬍qជ⬍␲/l 1 2␲បN0共EF兲Dqជ2 共21兲 ⫽ 3 2共kFl兲2

1⫺ l L

. 共22兲

In Eq. 共21兲, the lower limit is ⬃␲/L, the upper limit being ⬃␲/l, corresponding to the diffusive motion of the electron in real space. Note that N0(EF)⫽mkF/2␲2ប2 and D

⫽(1/3)vFl, whereas the contribution from the

free-particle-like density correlation is

Vnn⫽V

t⬍␶兩␺n共r兲兩 2 n共r兲兩2dr⬵VA4 ⫽V

1⫺ 3 共kFl兲2

1⫺ l L

, 共23兲 with A2⫽1⫺R.63 Since the phonon-mediated interaction is retarded for tret⬃1/␻D, only the free-particle-like density

correlation contributes to the phonon-mediated matrix ele-ment. This leads to the same weak localization correction to both the conductivity and the phonon-mediated matrix ele-ment. Here␻D is the Debye frequency.

The BCS Tc equation is, now,

Tc⫽1.13␻De⫺1/␭e f f, 共24兲 where ␭e f f⫽N0V

1⫺ 3 共kFl兲 2

1⫺ l L

. 共25兲 The initial change in Tc relative to Tc0 共for pure metal兲 is

then ⌬Tc Tc0 ⬵ 1 ␭ 3 共kFl兲2

1⫺ l L

⬀␳0 2 , 共26兲

where the BCS ␭ is N0V. This result is in good agreement with experiments.13,67–69

C. Strong-coupling theory

In the strong-coupling theory,70,71 the electron-phonon coupling constant is defined by71

␭⫽2

␣ 2兲F共d␻ 共27兲 ⫽N0

I 2

M

␻2

. 共28兲 Here F() is the phonon density of states and M is the ionic mass.

I2

and

␻2

are the average over the Fermi surface of the square of the electronic matrix element and the pho-non frequency.71 For a homogeneous system with the Ein-stein model, it is written as

␭0⫽N0 I0 2

MD2 , 共29兲 where I0is the electronic matrix element for the plane-wave states and ␻D denotes the Einstein phonon frequency. In

BCS theory, Eq.共4兲 leads to the BCS coupling constant

BCS⫽N0V, BCS theory. 共30兲

Comparing Eqs.共29兲 and 共30兲, we get V

I02

MD2. 共31兲 In general, using the equivalent electron-electron potential in the electron-phonon problem,72,73

V共x⫺x

兲→ I0 2

MD

2 D共x⫺x

兲, 共32兲 with x⫽(r,t), the Fro¨hlich interaction at finite temperatures for an Einstein model may be obtained by

Vkជkជ⬘共␻,␻

兲⫽ I02 MD 2

冕 冕

drdr

kជ*⬘共r兲⫺kជ* ⬘共r

兲D共r⫺r

,␻ ⫺␻

兲␾⫺kជ共r

兲␾kជ共r兲I0 2 MD2

兩␾kជ共r兲兩 2 kជ共r兲兩2drD 2 ␻D 2⫹共

2 ⫽Vkជkជ⬘ ␻D 2 ␻D 2 ⫹共␻⫺␻

兲2, 共33兲 where73 D共r⫺r

,␻⫺␻

兲⫽ 1 ⍀

qជD 2 共␻⫺␻

兲2 D 2 e iqជ•(r⫺r⬘) ⫽ ␻D 2 共␻⫺␻

兲2 D 2 ␦共r⫺r

兲. 共34兲 Here ␻ means the Matsubara frequency and␾kជ denotes the

plane-wave state. Note that the spatial part of the phonon Green’s function D(r⫺r

,␻⫺␻

) becomes the Dirac delta function, since the phonon frequency does not depend on the momentum. Consequently, Eq. 共33兲 leads to the coupling constant

(6)

␭0⫽N0

Vkជkជ⬘共0,0兲

⫽N0

I02

MD2 共35兲 and the strong-coupling gap equation74

⌬共kជ,␻兲⫽T

␻⬘

kជVkជkជ⬘共␻,␻

⌬共k

,␻

兲 ␻

2⫹E kជ⬘ 2 共␻

⫽T

␻⬘ ␻D 2 共␻⫺␻

兲2 D 2

kជVkជkជ⌬共k

,␻

兲 ␻

2⫹E kជ⬘ 2 共␻

兲, 共36兲 where Ekជ⬘共␻

兲⫽

k⬘ 2 ⫹⌬k⬘ 2 共␻

兲. 共37兲

In the presence of impurities, weak localization leads to a correction to ␣2 or

I2

,关disregarding the changes of F(␻) and N0兴. From Eq. 共32兲, one finds

Vnn⬘共␻,␻

兲⫽ I02 MD2

冕 冕

drdr

n⬘ *共r兲 n ¯ *共r

兲D共r⫺r

,␻ ⫺␻

兲␺共r

兲␺n共r兲I0 2 MD2

兩␺n共r兲兩 2 n共r兲兩2drD 2 ␻D 2 ⫹共␻⫺␻

兲2 ⫽Vnn⬘ ␻D 2 ␻D 2⫹共

2 共38兲 and ␭⫽N0

Vnn共0,0兲

⫽N0 I0 2 MD2

兩␺n共r兲兩 2 n共r兲兩2dr

, 共39兲 which agrees with BCS theory:

e f f⫽N0V

兩␺n共r兲兩2兩␺n共r兲兩2dr

. 共40兲

Therefore, both the weak- and strong-coupling gap equations give basically the same result of the weak localization effect in superconductors.

D. Resistance ratio

According to Matthiessen’s rule, the resistivity ␳(T) caused by static and thermal disorder is additive, i.e.,

共T兲⫽␳0⫹␳ph共T兲, 共41兲

where ␳ph is mostly due to electron-phonon scattering. At

high temperatures, the phonon limited electrical resistivity is given by75 ␳ph共T兲⫽ 4␲mkBT ne2ប

tr 2Fd␻, 共42兲

where ␣tr includes an average of a geometrical factor 1 ⫺cos␪kជkជ⬘. Assuming ␣tr 22, we obtainph共T兲⬵ 2␲mkBT ne2e f f⬵ 2␲mkBT ne2 N0 I02 MD2

1⫺ 3 共kFl兲2

. 共43兲 Note that decreasing Tc is accompanied by the loss of the

thermal resistivity ␳ph(T), in good agreement with experiment.4The ternary superconductor LuRh4B4 共Ref. 15兲 also shows the same behavior. The room temperature resis-tance ratio is then written as

␳共300 K兲 ␳0 ⫽ ␳0⫹␳ph共300 K兲 ␳0 ⬵1⫹ 2␲␶⫻300 K ប ␭e f f. 共44兲 When ␭e f f goes to zero, the system is not superconducting and resistance ratio becomes 1, which is in good agreement with experiments.4,8,9,11,46 More details will be published elsewhere.

IV. COMPARISON WITH EXPERIMENTS

Wiesmann et al.13 examined the Tc change of

vapor-deposited Nb3Ge and Nb3Sn as a function of␣ particle flu-ence. The 2.5-MeV␣ particles irradiated the samples, which were held at 30 K. The samples were then cooled, and both Tc and the residual resistivity ␳0 were measured. Figure 1 shows the dependence of Tc on ␳0 in Nb3Ge and Nb3Sn. Thin lines are our theoretical results obtained from Eqs.共24兲 and 共25兲. We find good agreement between theory and ex-periment. The Debye temperature and Tc0 共for the pure

sample兲 are ␻D⫽302 K,Tc0⫽23 K and ␻D⫽290 K,Tc0

⫽18 K for Nb3Ge and Nb3Sn, respectively. In the absence of experimental data for the inelastic diffusion length, we used the same value of L

Di

l⫻387 Å/T for both

materials.76Here␶imeans the inelastic scattering time and T

denotes temperature/K. We assumed ␶i⬀T⫺2 corresponding

to the electron-electron interactions.58,76Since it is very dif-ficult to evaluate kFl accurately,77 we assumed that ␳

FIG. 1. Calculated Tc’s vs residual resistivity␳0for Nb3Ge and Nb3Sn. Experimental data are due to Wiesmann et al., Ref. 13.

(7)

⫽100␮⍀ cm corresponds to kFl⫽3.65 and 3.60 for Nb3Ge

and Nb3Sn with the same value of kF⫽0.3 Å⫺1. These

val-ues also give a good fitting to the dependence of Tc on the

residual resistivity in impurity-doped samples.78This is per-suasive evidence that the Tcbehavior is not crucially

depen-dent on any specific defect; rather its behavior is governed by the residual resistivity.

Testardi and co-workers9prepared about 130 Nb-Ge films and examined the dependence of Tcon resistivity, resistance

ratio, chemical composition, and sputtering conditions. The Nb/Ge ratios were in the range of ⬃2.4–5.5 and film thick-ness were about 2000–3500 Å. Only films which show a width of the superconducting transition less than ⬃2 –3 K were chosen to ensure the macroscopic homogeneity of the samples. They found a universal correlation of Tcand

resis-tance ratio irrespective of all sputtering conditions, composi-tion, and specific nature of the disorder.4,8,9,11,79Figure 2共a兲 presents a sampling of Tc-vs-resistance-ratio data for 130

Nb-Ge films by them. The correlation between Tc and the resistance ratio is obvious. Resistance ratios less than 1 were generally found in films which are not superconducting, which agrees with theoretical expression, Eq.共44兲. Our the-oretical curve, which was obtained from Eqs.共24兲, 共25兲, and 共44兲, is also shown in the same figure. We again find good agreement between the theoretical curve and experiment. Since A15 compounds show deviations from Matthiessen’s rule possibly due to saturation,4,80 we adjusted␳ph(T) to fit

experimental values4 at 300 K in the following manner:

ph(300 K)⬵90␮⍀ cm⫻␭e f f(1⫺2.4/kFl)/␭0. We used

the same values for kF,L, andDas in Fig. 1. But we found

that Tc0⫽24 K for pure Nb3Ge gives a better fitting, which

supports the conjecture that sputtered films may have not yet achieved the highest possible Tc’s.9

Poate et al.8 irradiated superconducting Nb-Ge films by 2-MeV ␣ particles and found a Tc-resistance correlation

similar to that as-grown films. Figure 2共b兲 shows the corre-lations both for 130 as-grown films9 and for

␣-particle-irradiated films.8They lie nicely within the corre-lation band. It indicates that the correcorre-lation of Tc of the

re-sistance ratio is universal irrespective of how disorder is caused, e.g., by irradiation or substitutional alloying. There-fore, our theory provides an explanation of both data. The correlation was also reported in V3Si and V3Ge.4

V. DISCUSSION

It is clear that the weak localization effect in supercon-ductors caused by impurity doping or radiation damage should be subjected to further experimental study. In particu-lar, since the same weak localization correction term occurs in both the conductivity and the phonon-mediated interac-tions, comparative study of the normal and superconducting properties of the samples will be beneficial. It is noteworthy that Fiory and Hebard69found that both the conductivity and

FIG. 2. 共a兲 Calculated Tc’s vs resistance ratio for Nb-Ge. The data points relate to about 130 films made with various sputtering voltage, deposition conditions, film thickness, crystal structure, and chemical composition. Data are from Testardi et al., Ref. 9. 共b兲

Tc-resistance-ratio correlation band for 130 as-grown films with the values for damaged films superimposed. Data are from Poate et al., Ref. 8.

(8)

the transition temperature vary as (kFl)⫺2 for bulk

amor-phous InOx.

The antilocalization effect of the spin-orbit interaction will provide more insights into the weak localization effect in superconductors. In fact, Miller et al.81found compensa-tion for the Tcdecrease in highly disordered superconductors

by adding impurities with large spin-orbit scattering. The loss of thermal electrical resistivity␳ph(T)

共electron-phonon interaction兲 with decreasing Tc needs more

experi-mental study. In particular, we may consider samples satis-fying Matthiessen’s rule, where correlation of Tc and the

resistance ratio is more physically transparent. We propose to investigate the usual low-Tc superconductors near the

superconductor-insulator transition.82 We expect to find the loss of thermal electrical resistivity as approaching the insu-lating regime. Unfortunately, no systematic study is available yet. Note that this behavior may provide a means of probing the phononmechanism in exotic superconductors, such as, heavy fermion superconductors, organic superconductors, and high-Tc cuprates.

VI. CONCLUSION

We have considered irradiation effects on A15 supercon-ductors. The universal large reduction of Tc and Ic due to

radiation damage has been explained by the weak localiza-tion of electrons. Using the weak localizalocaliza-tion correclocaliza-tion to the phonon-mediated interaction derived from the density correlation function, we calculated Tc values which are in good agreement with experimental data. It is shown that weak localization decreases significantly both the electron-phonon interaction and the electron-phonon-mediated interaction, and thereby gives rise to the universal correlation of Tcand the resistance ratio.

ACKNOWLEDGMENTS

Y.J.K. is grateful to Professor Yun Kyu Bang and Profes-sor Bilal Tanatar for discussions and encouragement. M.P. thanks the FOPI at the University of Puerto Rico–Humacao for release time.

1S. V. Vonsovsky, Y. A. Izyumov, and E. Z. Kurmaev, Supercon-ductivity of Transition Metals 共Springer-Verlag, Berlin, 1982兲,

Chap. 9.

2C. L. Snead, Jr. and T. Luhman, in Physics of Radiation Effects in Crystals, edited by R. A. Johnson and A. N. Orlov 共Elsevier,

New York, 1986兲, Chap. 6.

3A. R. Sweedler, C. L. Snead, Jr., and D. E. Cox, in Treatise on Materials Science and Technology, edited by Th. Luhman and

D. Dew-Hughes共Academic Press, New York, 1979兲, Vol. 14, pp. 349–426.

4L.R. Testardi, J.M. Poate, and H.J. Levinstein, Phys. Rev. B 15,

2570共1977兲.

5R. Bett, Cryogenics 14, 361共1974兲.

6A.R. Sweedler, D.S. Schweitzer, and G.W. Webb, Phys. Rev.

Lett. 33, 168共1974兲.

7A.R. Sweedler, D.E. Cox, and L. Newkirk, J. Electron. Mater. 4,

883共1975兲.

8J.M. Poate, L.R. Testardi, A.R. Storm, and W.M. Augustyniak,

Phys. Rev. Lett. 35, 1290共1975兲.

9L.R. Testardi, R.L. Meek, J.M. Poate, W.A. Royer, A.R. Storm,

and J.H. Wernick, Phys. Rev. B 11, 4304共1975兲.

10A.R. Sweedler and D.E. Cox, Phys. Rev. B 12, 147共1975兲. 11J.M. Poate, R.C. Dynes, L.R. Testardi, and R.H. Hammond, Phys.

Rev. Lett. 37, 1308共1976兲.

12G. Ischenko, H. Adrian, S. Klau¨munzer, M. Lehmann, P. Mu¨ller,

H. Neumu¨ller, and W. Szymczak, Phys. Rev. Lett. 39, 43

共1977兲.

13H. Wiesmann, M. Gurvitch, A.K. Ghosh, H. Lutz, K.W. Jones,

A.N. Goland, and M. Strongin, J. Low Temp. Phys. 30, 513

共1978兲.

14C.L. Snead, Jr., J. Nucl. Mater. 72, 192共1978兲.

15R. C. Dynes, J. M. Rowell, and P. H. Schmidt, in Ternary Super-conductors, edited by G. K. Shenoy, B. D. Dunlap, and F. Y.

Fradin共North-Holland, Amsterdam, 1981兲, p. 169.

16K.E. Kihlstrom, D. Mael, and T.H. Geballe, Phys. Rev. B 29, 150 共1984兲.

17D.A. Rudman and M.R. Beasley, Phys. Rev. B 30, 2590共1984兲.

18J. Geerk, H. Rietschel, and U. Schneider, Phys. Rev. B 30, 459 共1984兲.

19R.C. Dynes and C.M. Varma, J. Phys. F: Met. Phys. 6, L215 共1976兲.

20

M. Gurvitch, A.K. Ghosh, C.L. Snead, Jr., and M. Strongin, Phys. Rev. Lett. 39, 1102共1977兲

21H. Wiesmann, M. Gurvitch, A.K. Ghosh, H. Lutz, O.F.

Kam-merer, and M. Strongin, Phys. Rev. B 17, 122共1978兲.

22A.K. Ghosh, M. Gurvitch, H. Wiesmann, and M. Strongin, Phys.

Rev. B 18, 6116共1978兲.

23D.F. Farrell and B.S. Chandrasekar, Phys. Rev. Lett. 38, 788 共1977兲.

24A.R. Sweedler, D.E. Cox, and S. Moehlecke, J. Nucl. Mater. 72,

50共1978兲.

25P.W. Anderson, K.A. Muttalib, and T.V. Ramakrishnan, Phys.

Rev. B 28, 117共1983兲.

26Yong-Jihn Kim and A.W. Overhauser, Phys. Rev. B 47, 8025 共1993兲.

27

P.W. Anderson, J. Phys. Chem. Solids 11, 26共1959兲.

28Yong-Jihn Kim, Mod. Phys. Lett. B 10, 555共1996兲. 29Yong-Jihn Kim, Int. J. Mod. Phys. B 11, 1731共1997兲.

30Mi-Ae Park and Yong-Jihn Kim, Bull. Am. Phys. Soc. 44, 1589 共1999兲.

31M.F. Merriam, S.H. Liu, and D.P. Seraphim, Phys. Rev. 136, A17 共1964兲.

32G. Boato, G. Gallinaro, and C. Rizzuto, Phys. Rev. 148, 353 共1966兲.

33A. Hofmann, W. Bauriedl, and P. Ziemann, Z. Phys. B: Condens.

Matter 46, 117共1982兲.

34M. Hitzfeld and G. Heim, Solid State Commun. 29, 93共1979兲. 35Yong-Jihn Kim and A.W. Overhauser, Phys. Rev. B 49, 15 799

共1994兲.

36Mi-Ae Park, M.H. Lee, and Yong-Jihn Kim, Physica C 306, 96 共1998兲.

37J.M. Valles, Jr., A.E. White, K.T. Short, R.C. Dynes, J.P. Garno,

A.F. Levi, M. Anzlowar, and K. Baldwin, Phys. Rev. B 39, 11 599共1989兲.

(9)

38Y. Li, G. Xiong, and Z. Gan, Physica C 199, 269共1992兲. 39Y. Dalichaouch, M.S. Torikachvili, E.A. Early, B.W. Lee, C.L.

Seaman, K.N. Yang, H. Zhou, and M.B. Maple, Solid State Commun. 65, 1001共1988兲.

40J. Fink, N. Nu¨cker, H. Romberg, M. Alexander, M.B. Maple, J.J.

Neumeier, and J.W. Allen, Phys. Rev. B 42, 4823共1990兲.

41C.C. Tsuei, S. von. Molnar, and J.M. Coey, Phys. Rev. Lett. 41,

664共1978兲.

42R.A. Pollak, C.C. Tsuei, and R.W. Johnson, Solid State Commun. 23, 879共1977兲.

43A. R. Sweedler et al.共unpublished兲.

44B.S. Brown, J.W. Hafstrom, and T.E. Klippert, J. Appl. Phys. 48,

1759共1977兲.

45A.K. Ghosh, H. Weismann, M. Gurvitch, H. Lutz, O.F.

Kam-merer, C.L. Snead, A. Goland, and M. Strongin, J. Nucl. Mater.

72, 70共1978兲.

46C.C. Koch, J.O. Scarbrough, and D.M. Kroeger, Phys. Rev. B 9,

888共1974兲.

47L.R. Testardi, J.M. Poate, W. Weber, W.M. Augustyniak, and

J.H. Barret, Phys. Rev. Lett. 39, 716共1977兲.

48R. Viswanathan and R. Caton, Phys. Rev. B 18, 15共1978兲. 49B.S. Brown, T.H. Blewitt, D.G. Wozniak, and M. Suenaga, J.

Appl. Phys. 46, 5163共1975兲.

50B.S. Brown, T.H. Blewitt, T.L. Scott, and D.G. Wozniak, J. Appl.

Phys. 49, 4144共1978兲.

51D.G. Schweitzer and D.M. Parkin, Appl. Phys. Lett. 24, 333 共1974兲.

52D.M. Parkin, and D.G. Schweitzer, Nucl. Technol. 22, 108 共1974兲.

53C.L. Snead, Jr. and D.M. Parkin, Nucl. Technol. 29, 264共1976兲. 54M. Ma and P.A. Lee, Phys. Rev. B 32, 5658共1985兲.

55D.J. Thouless, Philos. Mag. 32, 877共1975兲.

56C. Kittel, Quantum Theory of Solids共Wiley, New York, 1963兲,

Eq.共53兲 in Chap. 8.

57M. Kaveh and N.F. Mott, J. Phys. C 14, L177共1981兲. 58N.F. Mott and M. Kaveh, Adv. Phys. 34, 329共1985兲. 59

A. D. Stone, in Physics of Nanostructures, edited by J. H. Davies and A. R. Long共IOP, Bristol, 1992兲, p. 65.

60P. L. Taylor, Quantum Approach to the Solid State共Prentice-Hall,

Englewood Cliffs, NJ, 1970兲, p. 286.

61G. Bergmann, Phys. Rep. 101, 1共1984兲.

62D.J. Thouless, Phys. Rep., Phys. Lett. 13C, 93共1974兲. 63M. Kaveh, Philos. Mag. B 51, 453共1985兲.

64

W.L. McMillan, Phys. Rev. B 24, 2739共1981兲.

65B.L. Altshuler and A.G. Aronov, Zh. E´ ksp. Teor. Fiz. 77, 2028 共1979兲 关Sov. Phys. JETP 50, 968 共1979兲兴.

66E. Abrahams, P.W. Anderson, P.A. Lee, and T.V. Ramakrishnan,

Phys. Rev. B 24, 6783共1981兲.

67J.P. Orlando, E.J. McNiff, Jr., S. Foner, and M.R. Beasley, Phys.

Rev. B 19, 4545共1979兲.

68S.J. Bending, M.R. Beasley, and C.C. Tsuei, Phys. Rev. B 30,

6342共1984兲.

69A.T. Fiory and A.F. Hebard, Phys. Rev. Lett. 52, 2057共1984兲. 70G.M. Eliashberg, Zh. E´ ksp. Teor. Fiz. 38, 966 共1960兲 关Sov. Phys.

JETP 11, 696共1960兲兴.

71W.L. McMillan, Phys. Rev. 167, 331共1968兲.

72A. A. Abrikosov, L. P. Gor’kov, and I. E. Dzyaloshinski, Meth-ods of Quantum Field Theory in Statistical Physics共Dover, New

York, 1975兲, p. 79.

73A. Fetter and J. D. Walecka, Quantum Theory of Many-Particle Systems共McGraw-Hill, New York, 1971兲, p. 401.

74S. V. Vonsovsky, Y. A. Izyumov, and E. Z. Kurmaev, Supercon-ductivity of Transition Metals 共Springer-Verlag, Berlin, 1982兲,

Chap. 2.

75G. Grimvall, The Electron-Phonon Interaction in Metals

共North-Holland, Amsterdam, 1981兲, p. 4.

76P.A. Lee and T.V. Ramakrishnan, Rev. Mod. Phys. 57, 287 共1985兲.

77H. Gutfreund, M. Weger, and O. Entin-Wohlman, Phys. Rev. B 31, 606共1985兲.

78Yong-Jihn Kim and K.J. Chang, Mod. Phys. Lett. B 12, 763 共1998兲.

79H. Lutz, H. Weismann, O.F. Kammerer, and M. Strongin, Phys.

Rev. Lett. 36, 1576共1976兲.

80H. Wiesmann, M. Gurvitch, H. Lutz, A. Ghosh, B. Schwarz, M.

Strongin, P.B. Allen, and J.W. Halley, Phys. Rev. Lett. 38, 782

共1977兲.

81T.A. Miller, M. Kunchur, Y.Z. Zhang, P. Lindenfeld, and W.L.

McLean, Phys. Rev. Lett. 61, 2717共1988兲.

Referanslar

Benzer Belgeler

Gulick wrote that “while our American nation includes in it representa- tives from most of the peoples under the sun, we pos- sess less of the folk music, folk dances,

In this respect, we investigated the impact of charge and substitutional atom doping on the electronic transport properties of the hybrid metallic-semiconducting lateral

Lantanit iyonları, bir yandan uzun ömürlü ve kararlı yayma bantları ile güçlü lüminesans özellik gösterirken [39], diğer taraftan yüksek spin ve güçlü manyetik

In addition, it is stated that the optimal solution to the channel switching problem results in channel switching between at most two different channels, and an approach is proposed

Debre (2008), “İlköğretim Sosyal Bilgiler Dersi Coğrafya Konularının Öğretiminde Ders Anlatım Stratejisi Olarak Dramatizasyonun Kullanılmasının Öğrencinin Başarı

Gallego and van Ryzin [ 6 ] also show that the optimal revenue for the deterministic problem in which the demand rate is a deterministic function of a given price constitutes an

Minitab İstatistik Yöntemlerine göre Kontrol Grubu-4 ile 18°C ve 28°C deki 1 dişi ve 1 erkek olmak üzere dörderli gruplara ayrılan ikinci deneme gruplarının

音波拉皮、美白保養、電波拉皮、除臉部泛紅血絲、腋下與手部多 汗症治療