Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Patterson, Bailey - Solid State Physics Introduction to theory

.pdf
Скачиваний:
1117
Добавлен:
08.01.2014
Размер:
7.07 Mб
Скачать

8.5 The Theory of Superconductivity (A)

497

 

 

and G can be put in the form

 

G = k (εk Ek + k

 

 

(8.198)

bk ) .

Note by Fig. 8.20 and (8.193) how Ek predicts a gap, for clearly Ek

0. Continu-

ing

 

bk = −vkukαkαk + uk vk βk βkvk2αkβk+ uk2βkαk .

(8.199)

¯

 

 

 

 

 

 

 

 

 

 

 

 

 

But bk involves only diagonal terms, so using an appropriate anticommutation re-

lation

 

 

 

 

 

 

 

 

 

 

 

 

bk = uk vk 1

 

 

 

,

 

 

αkαk

βkβk

(8.200)

so

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

(8.201)

 

 

bk = ukvk (12nk ) ,

where

 

 

 

 

 

 

 

 

 

 

 

 

 

nk

=

 

1

 

 

= f (Ek ) .

(8.202)

 

 

e

βEk +1

 

 

 

 

 

 

 

 

 

 

f(Ek) is of course the Fermi function but it looks strange without the chemical potential. This is because α, βdo not change the particle number. See Marder [8.22]. Therefore,

k = −Vk,k

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

bk

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

= −Vk,kukvk[12 f (Ek)]

(8.203)

= −Vk,k

 

 

 

k

[12 f (Ek)].

 

 

 

 

 

 

 

k

 

 

 

2Ek

 

 

 

 

 

 

 

 

 

Now assume that (not using = 1)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

k=

 

 

when

 

 

εk

 

<

ωD ,

(8.204)

 

 

 

 

 

Vk,k= −V

when

 

 

 

εk

 

<

ωD ,

(8.205)

 

 

 

 

k = 0

 

when

 

εk

 

 

>

ωD ,

(8.206)

 

 

 

and

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Vk,k= 0

when

 

 

εk

 

 

>

ωD ,

(8.207)

 

 

 

where ωD is the Debye frequency (see (8.138)) So,

 

 

 

=V

|ε

 

|< ωD

 

 

[12 f (Ek)]

.

(8.208)

 

 

 

 

 

k

 

 

 

 

 

 

2Ek

 

 

 

 

 

 

 

 

 

 

 

 

 

 

498 8 Superconductivity

Fig. 8.20. Gap in single-particle excitations near the Fermi energy

For T = 0, then

 

 

 

 

 

 

 

 

 

 

= kV

 

εk2+

 

,

 

(8.209)

 

 

 

2

2

 

 

 

and for T 0, then

 

 

 

 

 

 

 

 

 

= k

V

 

 

 

E

 

 

(8.210)

2

 

tanh

 

k.

 

 

+

2

 

2kT

 

 

 

2 εk

 

 

 

 

 

 

We can then write

 

 

 

 

 

 

 

 

 

ωD N (E)V

 

 

 

 

dE .

(8.211)

0

 

 

E2 +

 

2

 

 

 

 

 

 

 

 

 

 

 

If we further suppose that N(E) constant N(0) ≡ the density of states at the Fermi level, then (8.211) becomes

1

=

ω

D

 

dE

 

 

 

 

E

2

+

2

 

ω

D

N (0)V

 

 

2

 

= ln E +

 

 

 

0

0

 

 

E

+

2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

(8.212)

 

 

ωD +

(

ωD )

2

+

2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

= ln

 

 

 

 

 

 

 

 

.

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

This equation can be written as

 

1

 

=

 

 

 

 

 

 

 

exp

 

 

 

 

 

 

 

 

 

 

 

 

 

 

ωD +

( ωD )2 +

 

 

 

 

 

 

N (0)V

 

2

 

 

(8.213)

 

 

 

 

=

 

 

 

 

=

 

 

 

 

 

ω0

 

1 + ( / ωD )

2

2 ω

D

 

 

 

 

 

 

 

+1

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

8.5 The Theory of Superconductivity

(A) 499

 

 

 

 

in the weak coupling limit (when <<

ωD). Thus, in the weak coupling limit, we

obtain

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

1

 

 

 

 

 

 

2 ω

D

exp

 

 

 

 

 

.

 

(8.214)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

N (0)V

 

 

From (8.210) by similar reasoning

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

ε

2

+

2

 

 

 

 

1

 

ωD

tanh

 

 

 

2kT

 

 

=

 

 

 

 

 

 

 

 

 

dε ,

(8.215)

N (0)V

 

 

 

 

2

 

2

 

 

0

 

 

 

ε

+

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

where, again, N(0) is the density of states at the Fermi energy.

For T greater than some critical temperature there are no solutions for , i.e. the energy gap no longer exists. We can determine Tc by using the fact that at T = Tc,

= 0. This says that

1

=

ωD

tanh(ε / 2kTc )

dε .

(8.216)

N (0)V

 

0

 

ε

 

In the weak coupling approximation, when N(0)V << 1, we obtain from (8.216) that

kTc =1.14 ωD exp(1/ N (0)V ) .

(8.217)

Equation (8.217) is a very important equation. It depends on three material properties:

a)The Debye frequency ωD

b)V that measures the strength of the electron–phonon coupling and

c)N(0) that measures the number of electrons available at the Fermi energy.

Note that typically ωD (m)–1/2, where m is the mass of atoms. This leads directly to the isotope effect. Note also the energy gap Eg = 2 (0) at absolute zero.

We can combine this result with our result for the energy gap parameter in the ground state to derive a relation between the energy gap at absolute zero and the critical superconducting transition temperature with no magnetic field. By (8.217) and (8.214), we have that

(0) = 2

ω

D

exp(1/ N (0)V ) =

 

2

kT ,

(8.218)

 

 

 

 

 

1.14

c

 

 

 

 

 

 

 

or

 

 

 

 

 

 

 

 

 

2

(0) = 3.52kTc .

 

 

 

(8.219)

Note that our expression for

(0) and Tc both involve the factor exp(−1/N(0)V); that

is, a power series (in V) expansion for both (0) and Tc have an essential singu larity in V. We could not have obtained reasonable results if we had tried ordinary

500 8 Superconductivity

perturbation theory because with ordinary perturbation theory, we cannot reproduce the effect of an essential singularity in the perturbation. This is similar to what happened when we discussed a single Cooper pair.

Our discussion has only been valid for weakly coupled superconductors.

Roughly speaking, these have (T /θ )2 > (500)−2. Pb, Hg, and Nb are strongly cou-

c D ~

pled, and for them (Tc /θD)2 ≥ (300)−2. Alternatively, the electron–phonon coupling parameter is about three times larger than is a typical weak coupling superconductor. A result for the strong coupling approximation is given below.

8.5.4Remarks on the Nambu Formalism and Strong Coupling Superconductivity (A)

The Nambu approach to superconductivity is presented by matrices and diagrams. The Nambu formalism includes the possibility of Cooper pairs in the calculation from the beginning via two component field operators. This approach allows for the treatment of retardation effects that need to be included for the strong (electron lattice) coupling regime. An essential step in the development was taken by Eliashberg and this leads to his equations. The Eliashberg strong coupling calculation of the superconducting transition temperature gives with a computer fit (via McMillan):

 

 

θD

 

1.04(1

+ λ)

 

T

=

exp

.

 

 

 

 

 

 

c

1.45

 

λ μ

(1

+

 

 

 

 

 

0.62λ)

θD is the Debye temperature, and for definitions of λ (the coupling constant) and μ* (the Coulomb pseudopotential term) see Jones and March [8.17]. They also

give a nice summary of the calculation. Briefly λ = N(0)Vphonon, μ = N(0)Vcoulomb where V in (8.218) is Vphonon Vcoulomb, and

 

 

EF

 

−1

μ = μ 1

+ μ ln

.

k θ

 

 

 

 

 

 

B D

 

Usually λ empirically turns out to be not much larger than 5/4 (or smaller).

q

k q

k

Fig. 8.21. Lowest-order correction to self-energy Feynman diagram (for electrons due to phonons)

8.7 Heavy-Electron Superconductors (EE, MS, MET) 501

The calculation includes the self-energy terms. The lowest-order correction to self-energy for electrons due to phonons is indicated in Fig. 8.21. The BCS theory with the extension of Eliashberg and McMillan has been very successful for many superconductors.

A nice reference to consult is Mattuck [8.23 pp. 267-272].

8.6 Magnesium Diboride (EE, MS, MET)

For a review of the new superconductor magnesium diboride, see, e.g., Physics Today, March 2003, p. 34ff. The discovery of the superconductor MgB2, with a transition temperature of 39 K, was announced by Akimitsu in early 2001. At first sight this might not appear to be a particularly interesting discovery, compared to that of the high-temperature superconductors, but MgB2 has several interesting properties:

1.It appears to be a conventional BCS superconductor with electron–phonon coupling driving the formation of pairs. It shows a strong isotope effect.

2.It does not appear to have the difficulty that the high-Tc cuprate ceramics have of having grain boundaries that inhibit current.

3.It is a widely available material that comes right off the shelf.

4.MgB2 is an intermetallic (two metals forming a crystal structure at a welldefined stoichiometry) compound with a transition temperature near double that of Nb3Ge.

Possibly, the transition temperature can be driven higher by tailoring the properties of magnesium diboride. At this writing, several groups are working intensely on this material, with several interesting results including the fact that it has two superconducting gaps arising from two weakly interacting bands.

8.7 Heavy-Electron Superconductors (EE, MS, MET)

UBe13 (Tc = 0.85 K), CeCu2Si2 (Tc = 0.65 K), and UPt3 (Tc = 0.54 K) are heavyelectron superconductors. They are characterized by having large low-temperature specific heats due to effective mass being two or three orders of magnitude larger than in normal metals (because of f band electrons). Heavy-electron superconductors do not appear to have a singlet state s-wave pairing, but perhaps can be characterized as d-wave pairing or p-wave pairing (d and p referring to orbital symmetry). It is also questionable whether the pairing is due to the exchange of virtual phonons—it may be due, e.g., to the exchange of virtual magnons. See, e.g., Burns [8.9 p51]. We have already mentioned these in Sect. 5.7.

502 8 Superconductivity

8.8 High-Temperature Superconductors (EE, MS, MET)

It has been said that Brazil is the country of the future and always has been as well as always will be. A similar comment has been made about superconductors. The problem is that superconductivity applications have been limited by the fact that liquid helium temperatures (of order 4 K) have been necessary to retain superconductivity. Liquid nitrogen (which boils at 77 K) is much cheaper and materials that superconduct at or above the boiling temperature of liquid nitrogen would open a large range of practical applications. Particularly important would be the transport of electrical power.

Just finding a high superconducting transition temperature Tc, however, does not solve all problems. The critical current can be an important limiting factor. Thermally activated creep of fluxoids (due to J × B) can lower Jc (the critical current) as the current interacts with the fluxoid and causes energy loss when the fluxoid becomes unpinned and thus creeps (can move). This is important in the high-Tc superconductors considered in this section.

Until 1986, the highest transition temperature for a superconductor was Tc = 23.2 K for Nb3Ge. Then Bednortz and Müller found a ceramic oxide (product of clay) of lanthanum, barium, and copper became superconducting at about 35 K. For this work they won the Nobel prize for Physics in 1987. Since Bednortz’s pioneering work several other high-Tc superconductors have been found.

The “1-2-3” compound YBa2Cu3O7, has a Tc of 92 K. The “2-1-4” compound (e.g. BaxLa2–xCuO4–y) are another class of high-Tc superconductors.

Tl2Ba2Ca2Cu3O10 has a remarkably high Tc of 125 K.

The high-Tc materials are type II and typically have a penetration depth to coherence length ratio K 100 and typically have a very large upper critical field. As we have mentioned, thermally activated creep of fluxoids due to the J × B force may cause energy dissipation and limit useable current values. For real materials, the critical current (Jc), critical temperature (Tc), and critical magnetic field (Bc) vary, but can be conveniently represented as shown in Fig. 8.22. As mentioned, the high-temperature superconductors (HTSs) are typically type II and also their Jc parallel to the copper oxide sheets (mentioned below) 107 A/cm2, while perpendicular to the sheets Jc can be about 107 A/cm2. A schematic of J, Bc, and Tc is shown in Fig. 8.22 for type I materials. For HTS, the representation of Fig. 8.22 is not complex enough. In Table 8.1 we list selected superconductor elements and compounds along with their transition temperature.

For HTS, we are faced with a puzzle as to what causes some ceramic copper oxide materials to be superconductors at temperatures well above 100 K. In conventional superconductors, we talk about electrons paired into spherically symmetric wave functions (s-waves) due to exchange of virtual phonons. Apparently, lattice vibrations cannot produce a strong enough coupling to produce such high critical temperatures. It appears parallel Cu-O planes in these materials play some very significant but not yet fully understood role. Hole conduction in these planes is important. As mentioned, there is also a strong anisotropy in electrical conduction. Although there seems to be increasing evidence for d-wave pairing, the exchange

8.8 High-Temperature Superconductors (EE, MS, MET) 503

mechanism necessary to produce the pair is still not clear as of this writing. It could be due to magnetic interactions or there may be new physics. See, e.g., Burns [8.9].

Table 8.1. Superconductors and their transition temperatures

Selected elements*

Transition temperature Tc (K)

Al

1

.17

Hg

4

.15

Nb

9

.25

Sn

3

.72

Pb

7

.2

 

 

 

Selected compounds*

 

 

 

 

 

Nb3Ge

23

.2

Nb3Sn

18.

Nb3Au

10

.8

NbSe2

7

.2

MgB2**

39

 

Copper oxide (HTS)*

 

 

 

 

 

Bi2Sr2Ca2Cu3O10

~110

 

YBa2Cu3O7

~92

 

Tl2Ba2Ca3Cu4O11

~122

 

 

 

 

Heavy fermion*

 

 

 

 

 

UBe13

0

.85

CeCu2Si2

0

.65

UPt3

0

.54

Fullerenes***

 

 

 

 

 

K3C60

19

.2

RbCs2C60

33

 

*Reprinted from Burns G, High Temperature Superconductivity Table 2-1 p. 8 and Table 3-1 p. 57, Academic Press, Copyright 1992, with permission from Elsevier. On p. 52 Burns also briefly discusses organic superconductors.

**Canfield PC and Crabtree GW, Physics Today 56(3), 34 (2003).

***Huffmann DR, “Solid C60,” Physics Today 41(11), 22 (1991).

504 8 Superconductivity

J

Bc

B

Tc

T

Fig. 8.22. J, B, T surface separating superconducting and normal regions

8.9 Summary Comments on Superconductivity (B)

1.In the superconducting state E = 0 (superconductivity implies the resistivity ρ vanishes, ρ → 0).

2.The superconducting state is more than vanishing resistivity since this would imply B was constant, whereas B = 0 in the superconducting state (flux is excluded as we drop below the transition temperature).

3.For “normal” BCS theory:

a)An attractive interaction between electrons can lead to a ground state separated from the excited states by an energy gap. Most of the important properties of superconductors follow from this energy gap.

b)The electron–lattice interaction, which can lead to an effective attractive interaction, causes the energy gap.

c)The ideas of the penetration depth (and, hence, the Meissner effect—flux exclusion) and the coherence length follow from the theory of superconductivity.

4.Type II superconductors have upper and lower critical fields and are technically important because of their high upper critical fields. Magnetic flux can penetrate between the upper and lower critical fields, and the penetration is quantized in units of hc/|2e|, just as is the magnetic flux through a superconducting ring. Using a unit of charge of 2e is consistent with Cooper pairs.

8.9 Summary Comments on Superconductivity (B) 505

5.In zero magnetic fields, for weak superconductors, superconductivity occurs at the transition temperature:

k

T

1.14 ω

D

exp(1/ N V ) ,

(8.220)

 

B c

 

0 0

 

where N0 is half the density of single-electron states, V0 is the effective interaction between electron pairs near the Fermi surface, and ωD θD, where θD is the Debye temperature.

6. The energy gap (2 ) is determined by (weak coupling):

(0) 2 ωD exp(1/ N0V0 ) =1.76kTc

(8.221)

 

 

 

 

 

T 1/ 2

 

 

.

(8.222)

(T )

 

 

 

 

 

 

T Tc

(0) 1

T

 

;

 

 

 

 

 

 

 

c

 

 

 

 

 

7. The critical field is fairly close to the empirical law (for weak coupling):

 

H

c

(T )

 

 

T

 

2

 

 

 

 

 

 

 

 

 

 

.

 

(8.223)

 

H

c

(0)

 

1T

 

 

 

 

 

 

 

 

 

 

c

 

 

 

 

8.The coherent motion of the electrons results in a resistanceless flow because a small perturbation cannot disturb one pair of electrons without disturbing all of them. Thus, even a small energy gap can inhibit scattering.

9.The central properties of superconductors are the penetration depth λ (of magnetic fields) and the coherence length ξ (or “size” of Cooper pairs). Small λ/ξ ratios lead to type I superconductors, and large λ/ξ ratios lead to type II behavior. ξ can be decreased by alloying.

10.The Ginzburg–Landau theory is used for superconductors in a magnetic field where one has inhomogeneities in spatial behavior.

11.We should also mention that one way to think about the superconducting transition is a Bose–Einstein condensation, as modified by their interaction, of bosonic Cooper pairs.

12.See the comment on spontaneously broken symmetry in the chapter on magnetism. Superconductivity can be viewed as a broken symmetry.

13.In the paired electrons of superconductivity, in s and d waves, the spins are antiparallel, and so one understands why ferromagnetism and superconductivity don’t appear to coexist, at least normally. However, even p-wave superconductors (e.g. Strontium Ruthenate) with parallel spins the magnetic fields are commonly expelled in the superconducting state. Recently, however, two materials have been discovered in which ferromagnetism and superconductivity coexist.

They are UGe2 (under pressure) and ZrZn2 (at ambient pressure). One idea is that these two materials are p-wave superconductors. The issues about these materials are far from settled, however. See Physics Today, p. 16, Sept. 2001.

14.Also, high-Tc (over 100 K) superconductors have been discovered and much work remains to understand them.

506 8 Superconductivity

In Table 8.2 we give a subjective “Top Ten” of superconductivity research.

Table 8.2 Top 10 of superconductivity (subjective)

Person

Achievement

Date/comments

 

 

 

 

 

– Started low-T physics

1.

H. Kammerlingh Onnes

Liquefied He

1908

 

 

Found resistance of Hg → 0

1911

– Discovered supercon-

 

 

at 4.19 K

 

 

ducting state

 

 

 

1911

– Nobel Prize

2. W. Meissner and

Perfect diamagnetism

1933

– Flux exclusion

 

R. Ochsenfeld

 

 

 

 

3.

F. and H. London

London equations and flux

1935

B proportional to curl of

 

 

expulsion

 

 

J

4.

V.L. Ginzburg and L.D.

Phenomenological equations

1950

– Eventually GLAG

 

Landau

 

 

 

equations

 

 

 

1962

– Nobel Prize, Landau

 

 

 

2003

– Nobel Prize, Ginzburg

 

A. A. Abrikosov

Improvement to GL equa-

1957

– Negative surface energy

 

 

tions, Type II

2003

– Nobel Prize

 

L. P. Gor’kov

GL limit of BCS and order

1959

– Order parameter pro-

 

 

parameter

 

 

portional to gap pa-

 

 

 

 

 

rameter

5.

A. B. Pippard

Nonlocal electrodynamics

1953

x and l dependent on

 

 

 

 

 

mean free path in alloys

6.

J. Bardeen, L. Cooper,

Theory of superconductivity

1957

– e.g. see (8.217)

 

and J. Schrieffer

 

1972

– Nobel Prize (all three)

7.

I. Giaver

Single-particle tunneling

1960

– Get gap energy

 

 

 

1973

– Nobel Prize

8.

B. D. Josephson

Pair tunneling

1962

– SQUIDS and metrology

 

 

 

1973

– Nobel Prize

9.

Z. Fisk, et al

Heavy fermion “exotic” su-

1985

– Pairing different than

 

 

perconductors

 

 

BCS, probably

10. J.G. Bednorz and K. A. High-temperature supercon-

1986

– Now, Tcs are over

 

Muller

ductivity

 

 

100 K

 

 

 

1987

– Nobel Prize (both)

Соседние файлы в предмете Химия