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

Dressel.Gruner.Electrodynamics of Solids.2003

.pdf
Скачиваний:
66
Добавлен:
27.08.2013
Размер:
3.91 Mб
Скачать

178

7 Broken symmetry states of metals

Fig. 7.1. (a) Temperature dependence of the superconducting penetration depth λ(T ) in comparison with the order parameter (T ). According to Eq. (7.2.10), (λ(T )/λ(0))2 is proportional to the density of superconducting charge carriers Ns. (b) Temperature dependence of the penetration depth λ(T ) calculated by Eq. (7.4.25).

(retarded) electron–phonon interaction and the cutoff is associated with the energy of the relevant phonons h¯ ωP. In case of density wave states such retardation does not play a role, and the cutoff energy is the bandwidth of the metallic state. As this is usually significantly larger than the phonon energies, the transition temperatures for density waves are, in general, also larger than the superconducting transition temperatures. Another difference lies in the character of the state: if we calculate the electronic density for superconductors, we find that it is constant and independent of position, this is due to the observation that the total momentum of the Cooper pairs is zero – the superconducting gap opens at zero wavevector. In

7.2 The response of the condensates

179

contrast, electron–hole condensates of the density states have a total momentum of 2kF (or 2kF); this leads to a spatial variation of the charge density given by Eq. (7.1.4). There is, however, one important difference: in the latter case, the period of the charge density is the same as the period of the underlying lattice. For density waves, this period π/ kF may be different from the lattice period; in fact this period is often incommensurate with the underlying lattice. If this is the case, the energy associated with these condensates is independent of their positions or of the phase φ – with interesting implications as far as the dynamics of density waves is concerned.

7.2 The response of the condensates

The ground state is a coherent superposition of Cooper pairs, all in the same quantum state with the same wavefunction, and as such the condensate can be written as a macroscopic wavefunction with an amplitude and phase

(r, t) = [Ns(r, t)]1/2 exp{iφ (r, t)} ,

(7.2.1)

where Ns(r, t) is the condensate density. The phase φ (r, t) is that of our entire macroscopic system and is therefore a physical observable – beautifully demonstrated by the Josephson effects. In order to see the significance of the phase, we calculate the current density due to the condensate1

J = −

e

 

p + p

;

(7.2.2)

 

2m

in one dimension we find

 

 

 

 

Nseh

 

dφ

 

 

 

 

J

 

 

 

 

,

 

(7.2.3)

 

=

¯

 

 

 

s

 

 

 

 

 

 

 

 

m dr

 

 

 

i.e. the current is determined by the spatial derivative of the phase. For density waves, a somewhat different handwaving argument applies. The current density is

JDW = −NDWev = −NDWe

dr

(7.2.4)

 

.

dt

We write the phase as φ = 2kF · r, and there are two electrons per density wave

period π/ kF; this leads to

 

 

 

 

JDW = −

e dφ

(7.2.5)

 

 

 

π dt

and NDW is the (two-dimensional) density in the plane perpendicular to the density wave modulation. The continuity equation gives the electron density

dJDW

+ e

dNDW

= 0 .

(7.2.6)

dr

dt

1 Note that the pairs have a charge 2e and mass 2m.

180

7 Broken symmetry states of metals

These differences between the superconducting and density wave ground state lead to responses to the electromagnetic fields which are different for the various ground states.

7.2.1 London equations

We write for the acceleration of the superconducting current density Js in the presence of an external field E

d m

Js = E ,

(7.2.7)

dt Nse2

where Ns is the superfluid particle density, which can be taken as equal to the particle density in the normal state N . In addition, using Maxwell’s equation (2.2.7a), this relation can be written as

d

×

m

1

B = 0 .

 

 

Js +

 

dt

Nse2

c

Flux expulsion, the so-called Meissner effect, is accounted for by assuming that not only the time derivative in the previous equation, but the function in the bracket itself, is zero, × (mc/Nse2)Js + B = 0. With Eq. (2.1.2) the expression reduces to

Nse2

 

Js = − mc A .

(7.2.8)

Thus, the superconducting current density is proportional to the vector potential A instead of being proportional to E as in the case of normal metals. We can also utilize Maxwell’s equation × B = 4cπ Js (neglecting the displacement term and the normal current) to obtain for the magnetic induction B

2B =

4π Nse2

 

mc2 B = λL2B ,

(7.2.9)

with the so-called London penetration depth

λL =

mc2

 

1/2

c

 

 

=

,

(7.2.10)

4π Nse2

ωp

where ωp is the well known plasma frequency. As this equation describes the spatial variation of B, λL characterizes the exponential decay of the electromagnetic field. As such, λL is the equivalent to the skin depth δ0 we encountered in metals (Eq. (2.3.16)). There are, however, important differences. δ0 is inversely proportional to ω1/2, and thus it diverges at zero frequency. In contrast, the penetration depth is frequency independent; there is a decay for B, even for dc applied fields. The temperature dependence of the penetration depth follows the

7.2 The response of the condensates

181

temperature dependence of the condensate density. Ns(T ) can be also calculated using the BCS formalism, and it is displayed in Fig. 7.1. For time-varying fields, Eq. (7.2.7) also gives the 1dependence of the imaginary part of the conductivity as

σ2(ω) =

Nse2

=

 

c2

(7.2.11a)

 

 

 

.

mω

4π λL2 ω

Through the Kramers–Kronig relation (3.2.11a) the real part of the conductivity is given by

 

π

 

Nse2

c2

 

σ1(ω) =

 

 

 

δ{ω = 0} =

 

δ{ω = 0} .

(7.2.11b)

2

m

8λL2

7.2.2 Equation of motion for incommensurate density waves

The energy related to the spatial and temporal fluctuations of the phase of the density waves is described by the Lagrangian density, which also includes the potential energy due to the applied electric field. The equation of motion of the phase condensate is

d2φ

vF2

m d2φ

 

e

 

 

 

 

 

=

 

kF · E(q, ω) ,

(7.2.12)

dt2

m

dr2

m

where m is the mass ascribed to the dynamic response of the density wave condensates. If the interaction between electrons is solely responsible for the formation of density wave states, m is the free-electron mass – or the bandmass in the metallic state, out of which these density wave states develop. This is simply because the kinetic energy of the moving condensate is 12 N mv2, with no additional terms included. This is the case for spin density waves. For charge density waves, however, the translational motion of the condensate leads also to oscillations of the underlying lattice; this happens because of electron–phonon coupling. The motion of the underlying ions also contributes to the overall kinetic energy, and this can be expressed in terms of the increased effective mass m . The appropriate expression is [Gru94]

m

= 1 +

 

4 2

,

(7.2.13)

mb

λ h2

ω2

 

 

 

P ¯

P

 

where λP is the dimensionless electron–phonon coupling constant, and ωP is the relevant phonon frequency, i.e. the frequency of the phonon mode which couples the electrons together with an electron–hole condensate. As the single-particle gap 2 ωP as a rule, the effective mass is significantly larger than the bandmass mb.

182

7 Broken symmetry states of metals

The frequency and wavevector dependent conductivity of the collective mode σˆ coll(q, ω) is obtained by a straightforward calculation, and one finds that

σˆ coll(q, ω) =

Js(q, ω)

=

 

1 m

 

iωωp2

(7.2.14)

 

 

 

 

 

 

 

 

 

.

E(q, ω)

8π m

ω2 vF2

m

q2

 

m

 

 

In general, non-local effects are not important and the q dependence can be neglected. The real part of the optical conductivity is

σ1coll(ω) =

π Nse2

 

2m δ{ω = 0} ;

(7.2.15a)

and the imaginary part is evaluated from Eq. (7.2.14) as

σ2 (ω) = −

2ω

σ coll(ω)

 

Nse2

(7.2.15b)

π

−∞ ω 2

ω2

dω = m ω .

coll

 

1

 

 

 

 

 

These are the same expressions as those we derived above for the superconducting case; this becomes evident by inserting the expression for the penetration depth λL into Eqs (7.2.11). Note, however, that because of the (potentially) large effective mass, the total spectral weight associated with the density wave condensate can be small.

7.3Coherence factors and transition probabilities

7.3.1Coherence factors

One of the important early results which followed from the BCS theory was the unusual features of the transition probabilities which result as the consequence of an external perturbation. Let us assume that the external probe which induces transitions between the various quasi-particle states has the form

k

 

|Hint|k, σ ak+ak.

 

Hint =

k , σ

(7.3.1)

,k ,σ,σ

In a metal or semiconductor, we will find that the various transitions between the states k and k proceed independently; as a consequence in order to obtain the total transition probability we have to add the squared matrix elements. In superconductors – and also for the other broken symmetry ground states – the situation is different. Using the transformation (7.1.10) we have utilized before, one can show that two transitions, (ak+ak) and (a+k,σ ak ,σ ), connect the same quasi-particle states. In the case where σ = σ , for example,

+

= +

v v

k

γ

k ,1

γ

+

 

 

 

γ

+

γ

k,0

 

ak ak

k

 

 

k,1

+ uk uk

 

k ,0

 

 

 

+

v

 

γ

k ,1

γ

k,0

+ uk

v

k

γ

+

γ

+

(7.3.2a)

 

k uk

 

 

 

 

k ,0

 

k,1

7.3 Coherence factors and transition probabilities

183

and

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

+

 

 

γ +

γ

k,1

+

v v

k

γ

k ,0

γ

+

 

ak,σ ak ,σ

= +ukuk

 

k ,1

 

k

 

 

 

k,0

 

 

+

v

γ

k ,0

γ

k,1

+ uk

v

k

γ

+

γ

+ .

(7.3.2b)

 

kuk

 

 

 

 

k ,1

 

k,0

The last two terms in both equations describe the creation of (or destruction of) pairs of quasi-particles, and the process requires an energy which exceeds the gap; here the combinations vkuk ηukvk appear in the transition probabilities. The first two terms describe the scattering of the quasi-particles, processes which are important for small energies h¯ ω < . For these processes the coefficients which determine the transition probabilities are uk uk + ηvk vk . These coefficients are the so-called coherence factors, with η = +1 constructive and with η = −1 destructive interference between the two processes. Whether η is positive or negative depends upon whether the interaction changes sign by going from k to k .

For electromagnetic waves, the absorption Hamiltonian is

 

 

 

 

 

 

eh

 

A(k k ) (k

k ) a+

 

a

 

,

(7.3.3)

 

 

¯

k

 

 

int

=

− · +

 

 

 

k

 

 

H

 

2mc ,k

k

 

 

 

 

and the interaction depends on the direction of the momentum. In this case, the transition probabilities add, and η = +1. The coherence factors can be expressed in terms of quasi-particle energies and gap value: after some algebra for scattering processes one obtains the form [Tin96]

2

 

1

1

 

ζkζk

2

.

 

ukuk + ηvkvk

=

 

+

 

+ η

 

(7.3.4)

2

EkEk

EkEk

When summing over the k values, the second term on the right hand side becomes zero, and the coherence factor can be simply defined as

 

 

 

 

1

 

 

 

 

2

 

 

0

if η

1

called case 1

 

 

 

F(E

, E ) =

 

1 +

η

 

 

 

1

if η

= −1

called case 2

(7.3.5)

 

 

2

EE

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

of hω

 

= +

 

 

if h

ω

2

 

. In the opposite case

2 , quasi-particles are created or

 

 

2

 

 

 

¯

 

 

 

 

 

 

 

 

 

 

 

 

 

 

¯

 

 

 

 

annihilated and with vkuk ηukvk

η the situation is reversed, and thus we find

 

 

 

 

F(

E

,

E

)

1

if

= −1

case 1

 

(7.3.6)

 

 

 

 

 

 

 

 

0

if η = +1

case 2 .

 

In the case of density waves a transformation similar to that performed for the superconducting case defines the quasi-particles of the system separated into two categories: the rightand left-going carriers. After some algebra, along the lines performed for the superconducting case, one finds that the case 1 and case 2 coherence factors for density waves are the opposite to those which apply for the superconducting ground state.

184

7 Broken symmetry states of metals

Similar arguments apply for other external probes. example, is described by the interaction Hamiltonian

H =

PωP)1/2

|P|

 

 

bP bP

 

 

 

 

P

 

 

k

 

 

Ultrasonic attenuation, for

ak+Pak,

(7.3.7)

where bP and bP refer to the longitudinal phonons of frequency ωP and density ρP. Here the change of the sign of P does not lead to the change of the sign of interactions. Consequently the transition probabilities add differently, and case 1 coherence factors apply. For processes which reverse the spin σ , the arguments are similar; such process, like nuclear magnetic resonance (NMR) relaxation, are characterized by case 2 coherences.

7.3.2 Transition probabilities

In order to derive an expression for the absorption rate W (per unit time and per unit volume) we start from Fermi’s golden rule. The transition rate WEE of absorbed energy in a transition from an occupied to an unoccupied state is proportional to the number of occupied quasi-particle states f (Es)Ds(Es) and to the number of unoccupied quasi-particle states [1 f (Es + h¯ ω)]Ds(Es + h¯ ω). Thus we obtain for the rate of absorbed energy:

Ws

| pEE |2 F(E, E + h¯ ω)Ds(E)Ds(E + h¯ ω) [ f (E) f (E + h¯ ω)] dE ,

 

(7.3.8)

where | pEE |2 is the matrix element of the transition from the state with energy E to the state with energy (E E + h¯ ω), h¯ ω is the energy of the external probe, Ds(E) is the density of states given by Eq. (7.1.16), f (E) = [1 + exp {E/ kBT }]1 is the Fermi distribution function, and we have also included the coherence factor F(E, E ). We normalize the transition probability to the value in the normal state – just above where the broken symmetry state develops – this is Wn =

|

p

 

|

2 Ds2

(0)hω. Then

 

 

 

 

 

 

 

 

 

 

 

 

 

 

EE

 

 

¯

 

 

 

hω) η 2 [ f ( )

 

 

 

 

 

 

 

 

 

 

 

 

W

 

1

(

(E2

 

f (

 

hω)]

 

E

 

 

 

 

 

 

Wn

= hω

−∞

 

2)1/2

[(E + hω)2

2

]1/2

 

 

 

 

 

 

 

 

s

 

 

 

|E E + ¯

+

|

E

E +

¯

d

 

.

(7.3.9)

 

 

 

 

 

 

¯

 

 

 

 

 

 

¯

 

 

 

 

 

 

 

 

As there are no states in the gap, the regions |E| or |E + h¯ ω| < are excluded from the integration. First we consider T = 0, then there are no thermally excited quasi-particles, and therefore the absorption rate is zero for ω < ωg = 2 /h¯ . The process allowing absorption of energy is pair breaking by creation of two quasiparticles with h¯ ω 2 ; the normalized rate takes the form:

Ws

 

1

[1 2 f (E + hω)] E(E + hω) + η 2

d

 

(7.3.10)

Wn

= hω

h¯ ω

 

E

(E2 2)1/2¯

[(E + hω)2

¯2

]1/2

 

 

 

¯

 

 

¯

 

 

 

 

 

Ws (ω) / Wn

1.5

(a)

1.0

0.5

0.00

7.3

Coherence factors and transition probabilities

 

 

 

 

 

185

 

 

 

 

Case 1

 

(b)

 

 

 

 

 

 

2.5

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Case 2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

(SDW)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

(Superconductor)

 

 

 

 

2.0

 

 

 

 

 

 

 

 

 

 

 

 

W

 

 

 

 

 

 

 

 

 

 

 

 

 

1.5

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

s

 

 

 

 

 

 

 

 

 

 

 

 

 

1.0

W / (T)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

n

 

 

Case 2

 

 

 

 

 

 

 

 

0.5

 

 

 

(Superconductor)

 

 

 

 

Case 1

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

(SDW)

 

 

0.0

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

1

2

 

0.0

0.5

1.0

 

 

 

 

 

 

 

Frequency ω / 2

h

 

 

Temperature T / Tc

 

 

 

 

Fig. 7.2. (a) Frequency dependence of the T 0 absorption rate Ws(ω)/ Wn for case 1 and 2 coherence factors. (b) Temperature dependence of the low frequency (h¯ ω 0.1 ) absorption rate Ws(T )/ Wn for case 1 and 2 coherence factors evaluated from Eq. (7.3.10). Case 1 applies for ultrasonic attenuation in superconductors and case 2 describes electromagnetic absorption or nuclear relaxation in the superconducting ground state. Case 1 also applies to the electromagnetic absorption in the spin density wave ground state.

for h¯ ω 2 . The frequency dependence of the absorption is different for case 1 and case 2 coherence factors. In the latter case, the transition probability (close to zero at the gap) effectively cancels the singularity of the density of states, and the absorption smoothly increases from zero at h¯ ω = 2 to the value equal to the normal state absorption. In contrast, F(E, E ) = 1 for frequencies h¯ ω = 2 and is weakly frequency dependent; this leads to a peak in the absorption for case 1 coherence.

As mentioned above, one does not observe an absorption at zero temperature for frequencies h¯ ω < 2 because pair breaking is not possible. Scattering of thermally excited single-particle states, however, becomes possible even at low frequencies for T > 0; for these the coherence effects have a different influence. At these low frequencies, the Fermi function difference f (E)f (E+h¯ ω) ≈ −f /∂E, the coherence factor, F(E, E ) 1, and the appropriate expression for ω = 0 is

Ws

= 2

 

 

E

2

+

2

f

dE ,

(7.3.11)

Wn

0

 

E

2

2

E

 

 

 

 

 

 

 

 

 

 

which diverges at Tc. At finite frequencies this divergence is removed, but one still recovers a maximum somewhat below Tc. In contrast, for case 1 coherence factors, it is easy to show that the transition rate is suppressed when compared with the normal state transition. Fig. 7.2 displays the absorption as a function of energy and as a function of temperature for both case 1 and 2 coherence factors.

186

7 Broken symmetry states of metals

7.4 The electrodynamics of the superconducting state

The electrodynamics of superconductors involves different issues and also varies from material to material. The reason for this is that different length scales play important roles, and their relative magnitude determines the nature of the superconducting state, and also the response to electromagnetic fields. The first length scale is the London penetration depth

λL =

c

(7.4.1)

ωp

determined through the plasma frequency ωp by the parameters of the metallic state and derived in Section 7.2.1. The second length scale is the correlation length

 

 

hv

F

 

ξ0

=

¯

(7.4.2)

 

 

π

 

describing, crudely speaking, the spatial extension of the Cooper pairs. This length scale can be estimated if we consider the pair wavefunctions to be the superpositions of one-electron states within the energy region around the Fermi level; then the corresponding spread of momenta δp is approximately | | = δ( p2/2m) vFδp/2, where vF is the Fermi velocity. This corresponds to a spatial range of ξ0 h¯ p. The correct expression, the BCS coherence length, is given by Eq. (7.4.2) at zero temperature. This is the length scale over which the wavefunction can be regarded as rigid and unable to respond to a spatially varying electromagnetic field. This parameter is determined through the gap , by the strength of the coupling parameter. The third length scale is the mean free path of the uncondensed electrons

= vFτ

(7.4.3)

set by the impurities and lattice imperfections at low temperatures; vF is the Fermi velocity, and τ is the time between two scattering events. For typical metals, the three length scales are of the same order of magnitude, and, depending on the relative magnitude of their ratios, various types of superconductors are observed; these are referred to as the various limits. The local limit is where ξ, λ. More commonly it is referred to as the dirty limit defined by 0. In the opposite, so-called clean, limit → ∞, it is necessary to distinguish the following two cases: the Pippard or anomalous limit, defined by the inequality λ ξ, (type I superconductor); and the London limit for which ξ λ, (type II superconductor).2

We first address the issues related to these length scales and describe the modifications brought about by the short mean free path; these are discussed using

2 See Fig. E.6 and Appendix E.4 for further details.

7.4 The electrodynamics of the superconducting state

187

spectral weight arguments. The relationship between λL and ξ can be cast into the form which is similar to the relationship between the skin depth δ0 and the mean free path in metals; here also we encounter non-local electrodynamics in certain limits. Finally we recall certain expressions for the conductivity and surface impedance as calculated by Mattis and Bardeen [Mat58]; the formalism is valid in various limits, as will be pointed out.

7.4.1 Clean and dirty limit superconductors, and the spectral weight

The relative magnitude of the mean free path with respect to the coherence length has important consequences for the penetration depth, and this can be described using spectral weight arguments. The mean free path is given by vFτ , whereas the coherence length ξ vF/ , and consequently in the clean limit 1/τ < and 1/τ > is the so-called dirty limit. In the former case the width of the Drude response in the metallic state, just above Tc, is smaller than the frequency corresponding to the gap 2 /h¯ , and the opposite is true in the dirty limit.

The spectral weight associated with the excitations is conserved by going from the normal to the broken symmetry states. While we have to integrate the real part of the normal state conductivity spectrum σ1n(ω), in the superconducting phase there are two contributions: one from the collective mode of the Cooper pairs σ1coll(ω), and one from the single-particle excitations σ1sp(ω); thus

sp

π

 

N e2

 

−∞

σ1coll + σ1

dω = −∞ σ1ndω =

 

 

 

(7.4.4)

2

 

m

assuming that all the normal carriers condense. Because of the difference in the coherence factors for the superconducting and density wave ground states, the conservation of the spectral weight has different consequences. The arguments for superconductors were advanced by Tinkham, Glover, and Ferrell [Glo56, Fer58], who noted that the area A which has been removed from the integral upon going through the superconducting transition,

[σ1n σ1s] dω = A

(7.4.5)

0+

is redistributed to give the spectral weight of the collective mode with σ1= 0) = Aδ{ω}. A comparison between this expression and Eq. (7.2.11b) leads to

c

 

(7.4.6)

λ = √8A

,

connecting the penetration depth λ to the missing spectral weight A. This relationship is expected to hold also at finite temperatures and for various values of the mean free path and coherence length – as long as the arguments leading to the