• Keine Ergebnisse gefunden

BCS theory for finite size superconductors

N/A
N/A
Protected

Academic year: 2022

Aktie "BCS theory for finite size superconductors"

Copied!
4
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

arXiv:0710.2286v2 [cond-mat.supr-con] 5 May 2008

BCS theory for finite size superconductors

Antonio M. Garc´ıa-Garc´ıa,1Juan Diego Urbina,2, 3 Emil A. Yuzbashyan,4Klaus Richter,5and Boris L. Altshuler6, 7

1Physics Department, Princeton University, Princeton, New Jersey 08544, USA

2Institut F¨ur Theoretische Physik, Universit¨at Regensburg, 93040 Regensburg, Germany

3Department of Physics, Universidad Nacional de Colombia, Cll45 Cra 30, Bogota, Colombia

4Center for Materials Theory, Rutgers University, Piscataway, New Jersey 08854, USA

5Institut F¨ur Theoretische Physik, Universit¨at Regensburg, 93047 Regensburg, Germany

6Physics Department, Columbia University, 538 West 120th Street, New York, NY 10027, USA

7NEC-Laboratories America, Inc., 4 Independence Way, Princeton, NJ 085540, USA

We study finite size effects in superconducting metallic grains and determine the BCS order parameter and the low energy excitation spectrum in terms of size, and shape of the grain. Our approach combines the BCS self-consistency condition, a semiclassical expansion for the spectral density and interaction matrix elements, and corrections to the BCS mean-field. In chaotic grains mesoscopic fluctuations of the matrix elements lead to a smooth dependence of the order parameter on the excitation energy. In the integrable case we observe shell effects when e.g. a small change in the electron number leads to large changes in the energy gap.

PACS numbers: 74.20.Fg, 75.10.Jm, 71.10.Li, 73.21.La

Since experiments by Ralph, Black, and Tinkham [1]

on Al nanograins in mid nineties, there has been con- siderable interest in the theory of ultrasmall supercon- ductors (see [2, 3] for earlier studies). In particular, finite-size corrections to the predictions of the Bardeen, Cooper, and Schriffer (BCS) theory for bulk supercon- ductors [4] have been studied [5–10] within the exactly solvable Richardson model [11]. Pairing in specific po- tentials, such as a harmonic oscillator potential [12] and a rectangular box, [13, 14] and mesoscopic fluctuations of the energy gap [15, 16] have been explored as well.

Nevertheless, a comprehensive theoretical description of the combined effect of discrete energy spectrum and fluc- tuating interaction matrix elements has not yet emerged.

We note that the Richardson model alone cannot provide such a description as it does not allow for mesoscopic fluc- tuations of the matrix elements.

In the present paper we develop a framework based on the BCS theory and semiclassical techniques that permits a systematic analytical evaluation of the low energy spec- tral properties of superconducting nanograins in terms of their size and shape. Leading finite size corrections to the BCS mean-field can also be taken into account in our approach. Our main results are as follows. For chaotic grains, we show that the order parameter is energy de- pendent. The energy dependence is universal, i.e. its functional form is the same for all chaotic grains. The matrix elements are responsible for most of the deviation from the bulk limit. In integrable grains, we find that the superconducting gap is strongly sensitive to shell effects, namely, a small modification of the grain size or number of electrons can substantially affect its value.

We start with the BCS Hamiltonian, H = P

ǫncc−P

n,nIn,ncn↑cn↓cncn,wherec an-

nihilates an electron of spinσin staten,

In,n =I(ǫn, ǫn) =λV δ Z

ψn2(r)ψ2n(r)dV (1)

are matrix elements of a short-range electron-electron in- teraction, λ is the BCS coupling constant, and ψn and ǫnare eigenstates and eigenvalues of the one-body mean- field Hamiltonian of a free particle of massm in a clean grain of volumeV. Eigenvaluesǫnare measured from the Fermi levelǫF and the mean level spacingδ= 1/νTF(0), whereνTF(0) = 2V2 2m~2

3/2√ǫF is the spectral density at the Fermi level in the Thomas-Fermi approximation.

Our general strategy can be summarized as follows: a) use semiclassical techniques to compute the spectral den- sityν(ǫ) =P

nδ(ǫ−ǫn) and I(ǫ, ǫ) as series in a small parameter 1/kFL, wherekF is the Fermi wavevector and L≃V1/3 is the size of the grain b) solve the BCS gap equation in orders in 1/kFL c) evaluate the low energy spectral properties of the grain such as the energy gap, excitation energies, and Matveev-Larkin parameter [5]

including finite size corrections to the BCS mean-field.

The results thus obtained are strictly valid in the region, kFL ≫1 (limit of validity of the semiclassical approxi- mation),δ/∆0<1 (limit of validity of the BCS theory), andl≫ξ≫L(condition of quantum coherence) where ξ=~vF/∆0is the superconducting coherence length,vF

is the Fermi velocity,l is the coherence length of the sin- gle particle problem and ∆0 is the bulk gap. We note that in Al grains [9] ξ ≈1600nm andl > 10000nmfor temperaturesT ≤4K. Therefore the region l≫ξ≫L is accessible to experiments.

Since the matrix elementsI(ǫ, ǫ) are energy dependent the BCS order parameter ∆(ǫ) also depends on energy.

(2)

2 The self-consistency equation for ∆(ǫ) reads

∆(ǫ) = Z ED

−ED

∆(ǫ)I(ǫ, ǫ) 2

q

ǫ2+ ∆(ǫ)2

ν(ǫ)dǫ, (2)

where ED is the Debye energy. In the limit V → ∞, the spectral density in the 2ED energy window near the Fermi level can be taken to be energy indepen- dent and given by the Thomas-Fermi approximation, ν(ǫ) =νTF(0), matrix elements are also energy indepen- dent,I(ǫ, ǫ) =λδ, and the gap is equal to its bulk value,

0= 2EDeλ1. As the volume of the grain decreases the mean level spacing increases and eventually both ν(ǫ) andI(ǫ, ǫ) deviate from the bulk limit.

Semiclassical evaluation ofν(ǫ). The spectral density in a 3dgrain,

ν(ǫ)≃νTF(0) [1 + ¯g(0) + ˜gl)] (3) consists of a monotonous part, ¯g(0) = ±4kFV + k2C2

FV

and an oscillatory contribution ˜gl). Here S and C denote the surface area and mean curvature of the grain, respectively, and upper/lower signs stand for Neu- mann/Dirichlet boundary conditions. The oscillatory contribution, to leading order, is given by the Gutzwiller trace formula [17, 18],

˜

gl) =ℜ 2π k2FV

l

X

p

Ape i

kFLppkFLp

F

. (4) where both the amplitude Ap and the topological in- dex βp depend on classical quantities only [18]. The summation is over a set of classical periodic orbits pof lengthLp. For isolated grains Dirichlet is the most nat- ural choice, but we also include Neumann to illustrate the dependence of our results on boundary conditions.

Only orbits shorter than the quantum coherence length l of the single-particle problem are included. This effec- tively accounts for inelastic scattering and other factors that destroy quantum coherence. Here we focus on the limit l ≫ ξ, the case l ∼ξ will be discussed elsewhere [20]. In Eq. (4) classical actions ~k(ǫ)Lp are expanded as k(ǫ)≈ kFkF/2ǫF. The amplitude Ap increases by a factor (kFL)1/2≫1 for each of the symmetry axes of the grain.

Semiclassical evaluation ofI(ǫ, ǫ). For integrable sys- temsI(ǫ, ǫ) depends on details of the system. In a rect- angular box it is simplyI(ǫ, ǫ) =λδbut in most other ge- ometries an explicit expression in terms of classical quan- tities is not available. In the chaotic case the situation is different. As a result of the quantum ergodicity theorem [19] it is well justified to assume that for systems with time reversal symmetry (the only ones addressed in this paper),ψn2(~r) = V1(1 +O(1/kFL)). In order to explicitly determine deviations from the bulk limit we replaceψ2n(~r)

inIn,n withhψ2(~r)iǫn, whereh. . .iǫ stands for an energy average aroundǫ. The single-particle probability density is thus effectively averaged over a small energy window resembling the effect of a finite coherence length.

Substituting hψ2(~r)iǫ intoIn,n, we obtain I(ǫ, ǫ) = λ

V

"

1− Sπ

4kFV 2

+ ¯I(ǫF, ǫ, ǫ)

# , (5) where

I(ǫ¯ F, ǫ, ǫ) = ¯IshortF) + ¯IlongF, ǫ−ǫ) (6) can be split into two parts coming from short and long orbits. Short orbits involve a single reflection at the grain boundary and result in a monotonous contribution

shortF) = πS

4kFV, (7)

while the contribution of long orbits depends on the en- ergy differenceǫ−ǫ

longF, ǫ−ǫ) = 1 VΠl

ǫ−ǫ ǫF

, (8)

with Πl(w) =R Pl

γ(r)D2γcos [wkFLγ]dr, where the sum is over all non-zero classical paths (not periodic orbits) γ(r) starting and ending at a given point r inside the grain [21] and the amplitude Dγ is defined in Refs. [17, 20, 21]. The integral stands for an average over all points r inside the grain. The explicit evaluation of Πl(w) for a given geometry requires in principle the knowledge of all classical pathsLγ up to lengthl. However, forl≫L, one can use asum rule for classical closed orbits [23] to obtain

Πl(w) = 2π

kF

2

sin(wkFl) wkF

. (9)

Solution of the gap equation. First, let us consider chaotic grains. Here we present only the final answer for the 3dcase deferring a more detailed account, including the 2dcase, to Ref. [20]. Writing the gap function ∆(ǫ) formally as a series in 1/kFL,

∆(ǫ) = ∆0

h1 +f(1)+f(2)+f(3)(ǫ)i

, (10) substituting it into Eq. (2), and using the above expres- sions for the density of states and interaction matrix el- ements, we derive

f(1)= 1±1 λ

πS

4kFV, (11)

where ± stands for Neumann (+) and Dirichlet (−) boundary conditions. Note that to leading order the combined effect of the interaction matrix elements and

(3)

3

-4 -2 0 2 4

ε/∆0

1 1.2 1.4 1.6

∆(ε)/∆

0

FIG. 1: Superconducting order parameter ∆(ǫ) in units of the bulk gap ∆0 for chaotic Al grains (kF = 17.5nm1, δ = 7279/N,∆0 ≈ 0.24meV) as a function of energy ǫcounted from the Fermi level. Different curves correspond to grain sizes (top to bottom) L = 6nm, kFL = 105, δ/∆0 = 0.77) (Dirichlet and Neumann boundary conditions), L = 8nm, kFL = 140, δ/∆0 = 0.32 (Dirichlet), and L = 10nm, kFL= 175, δ/∆0 = 0.08 (Dirichlet). The leading con- tribution comes from the energy dependent matrix elements I(ǫ, ǫ) given by Eq. (5).

the density of states have very different consequences on the gap, depending on the kind of boundary conditions.

For Dirichlet the leading finite size corrections to the gap vanishes.

The second order (1/(kFL)2) correction reads λf(2)= 2C

k2FV + 2

∓1 +1±1 λ

πS 4kFV

2

+ ˜g(0), where,

˜

g(0) = 2π kF2V

l

X

p

ApW(Lp/ξ) cos(kFLpp) (12)

and,

W(Lp/ξ) = λ 2

Z

−∞

dtcos(Lpt/ξ)

√1 +t2 (13) exponentially suppresses periodic orbits longer thanξ.

The third order correction (included in the definition ofδ) is energy dependent,

f(3)(ǫ) = πλδ

0

"

0

2+ ∆20 +π 4

#

. (14)

Note that a)δ/∆0≪1 is an additional expansion param- eter, therefore the contribution (14) can be comparable to lower orders in the expansion in 1/kFLand b) the or- der parameter ∆(ǫ) has a maximum at the Fermi energy (ǫ= 0) and slowly decreases on an energy scaleǫ∼∆0as one moves away from the Fermi level. One can also show

that mesoscopic corrections given by Eqs. (11,12) and (14) always enhance ∆(0) as compared to the bulk value

0. Fig. 1 shows the gap function ∆(ǫ) for Al grains of different sizesL, where we used (see [1])kF ≈17.5nm−1, λ≈0.18, andδ≈7279/N meV, whereN the number of particles.

Several remarks are in order: a) the smoothing of the spectral density energy dependence in Eq. (12) caused by a cutoff functionW is a superconductivity effect not related to the destruction of quantum coherence, b) the energy dependence of the gap is universal in the sense that it does not depend on specific grain details, c) the matrix elementsI(ǫ, ǫ) play a crucial role, e.g. they are responsible for most of the deviation from the bulk limit in Fig. 1, d) the requirementξ≫Lused to derive Eq. (9) is well justified for nanograins since L ∼ 10nm, while ξ∼104nm.

We now turn to the integrable case. Probably the simplest example is that of a rectangular box, since in this case the interaction matrix elements are simply I(ǫ, ǫ) = λδ. The calculation is simplified as now the order parameter is energy independent. We have

∆ = ∆0

h1 +f(1)+f(3/2)+f(2)i

, (15)

wheref(n)∝(kFL)−nλ−1. We obtain λf(1)= ¯g(0) + ˜g(1)(0), λf(3/2)= X

i,j6=i

˜

gi,j(3/2)(0),

λf(2)=X

i

˜

gi(2)(0) +f(1)[f(1)−g(0)],¯

(16)

where ˜g(k)∝(kFL)−k denotes the oscillating part of the spectral density and indexes i and j take values 1, 2, and 3 in three dimensions. Explicit expressions for ˜g(k),

˜

g(k)i , and ˜g(k)i,j in terms of periodic orbits for a rectangu- lar box can be found in Ref. [18] (the cutoff function in our case is given by Eq. (13)). We note that: a) Eq.(15) is also obtained by expanding the standard expression of the bulk gap ∆ = 2EDexp(−νTF(0)/ν(0)λ) in pow- ers of (kFLλ)−1 with ν(0) given by Eq.(3). b) unlike the chaotic case, the leading smooth correction to the bulk limit does not vanish for any boundary condition, c) smooth and oscillating corrections are of comparable magnitudes.

Shell effects and fluctuations. Motivated by previous studies for other fermionic systems such as nuclei and atomic clusters (see e.g. Ref. [22]), we investigate shell effects in metallic nanograins. In particular, we are inter- ested in the fluctuations of the BCS gap with the num- ber of electrons on the grain. As an illustration let us consider a cubic geometry. To determine the gap, we solve the gap equation (2) numerically and determine the

(4)

4

5.1 5.2 5.3 5.4

0

0 0.5 1 1.5 2

/0

11.90 12.00 12.10 12.20

0

0.5 1 1.5

FIG. 2: Superconducting order parameter ∆ in units of the bulk gap ∆0 for a cubic Al grain as a function of the ratio

0/δ, whereδ is the mean level spacing. Black crosses cor- respond to the exact numerical solution of the gap equation (2), while the red circles represent the semiclassical analytical expression (15).

Fermi energy for a given number of electrons N by in- verting the relation 2RǫF

ν(ǫ)dǫ = N. We find a good agreement between numerical results and the semiclassi- cal expansion (15), see Fig. 2. We also observe that a slight modification of the grain size (or equivalently the number of electronsN or the mean level spacingδ) can result in substantial changes in the value of the gap, see Fig. 2. The typical magnitude of fluctuations of the gap,

˜

0 ≈q

πδ

4∆0 [16] is consistent with our results (see Fig.

2).

Low energy excitations. Having solved the gap equa- tion (2), one can evaluate low energy properties of the grain taking into account finite size corrections to the BCS mean-field approximation. For example, the energy cost for breaking a Cooper pair in an isolated grain is [24],

∆E= 2∆(0)−δ, (17)

where ∆(0) is the solution of equation (2) taken at the Fermi energy and is given by Eqs. (10) and (15) for chaotic and rectangular shapes, respectively. We note that the correction to the mean-field (−δ) has been evalu- ated [10] for constant interaction matrix elements. Never- theless, since the deviation of matrix elements from a con- stant energy independent value is itself of order (kFL)−1, Eq. (17) is accurate up to terms of order (δ/∆0)(kFL)−1, which are negligible as compared to the ones we kept in Eqs. (17), (10), and (15).

Similarly, the Matveev-Larkin parity parameter [5]

reads ∆p ≡ E2N+112 E2N +E2N+2

= ∆(0)− δ2, whereEN is the ground state energy for a superconduct- ing grain with N electrons. Quasiparticle energies are pǫ2+ ∆(ǫ)2 plus corrections to mean-field, which can be determined using the approach of Ref. [10].

We see that finite size corrections to the BCS mean-

field approximation are comparable to the energy depen- dent correction (10) obtained within mean-field, but have an opposite sign. We also note that our approach of ex- panding around the bulk BCS ground state is applicable only when δ ≪ ∆0, i.e. when corrections to the BCS mean-field approximation are small [25].

To conclude, we have determined the low energy ex- citation spectrum for small superconducting grains as a function of their size and shape by combining the BCS mean-field, semiclassical techniques and leading correc- tions to the mean-field. For chaotic grains the non-trivial energy dependence of the interaction matrix elements leads to a universal smooth dependence (14) of the gap function on excitation energy. In the integrable case we found that small changes in the number of electrons can substantially modify the superconducting gap.

A.M.G. thanks Jorge Dukelsky for fruitful conversa- tions. K.R. and J.D.U. acknowledge conversations with Jens Siewert and financial support from the Deutsche Forschungsgemeinschaft (GRK 638). E.A.Y. was sup- ported by Alfred P. Sloan Research Fellowship and NSF award NSF-DMR-0547769.

[1] D.C. Ralph, et al., Phys. Rev. Lett.74, 3241 (1995); C.T.

Black, et al., Phys. Rev. Lett.76, 688 (1996).

[2] R. Parmenter, Phys. Rev.167, 387 (1968).

[3] B. Muhlschlegel, et al, Phys. Rev.B 6, 1767 (1972).

[4] J. Bardeen, et al., Phys. Rev.108, 1175 (1957).

[5] K. A. Matveev et al., Phys. Rev. Lett.78, 3749 (1997).

[6] G. Sierra, et al., Phys. Rev. B61, R11890 (2000); J. von Delft,Annalen der Physik(Leipzig),10, 219 (2001).

[7] J. Dukelsky, et al., Phys. Rev. Lett.83, 172 (1999); Phys.

Rev.B 61, 12302 (2000); J.M. Roman, et al., Nucl. Phys.

B634, 483 [FS] (2002);R. A. Smith et al., Phys. Rev.

Lett.77, 4962 (1996).

[8] M. Schechter, et al.,Phys. Rev. B63, 214518 (2001).

[9] J.M. Gordon, et al., Phys. Rev. B35, 4909 (1987).

[10] E. A. Yuzbashyan, et al.,Phys. Rev. B71, 094505 (2005).

[11] R.W. Richardson, Phys. Lett.3, 277 (1963).

[12] H. Heiselberg, Phys. Rev. A68, 053616 (2003).

[13] V.N. Gladilin, et al., Solid State Comm.121, 519 (2002).

[14] A. A. Shanenko, et al., Phys. Rev.B 74, 052502 (2006);

M. D. Croitoru, et. al, Phys. Rev.B 76, 024511 (2007).

[15] M. Farine, et al. , Phys. Rev. B68, 024507 (2003).

[16] H. Olofsson, et al. Phys. Rev. Lett.100, 037005 (2008).

[17] M. Gutzwiller,Chaos in Classical and Quantum Mechan- ics, Springer, New York, 1990.

[18] M. Brack, and R.K. Bhaduri, Semiclassical Physics, (Addison-Wesley, New York, 1997).

[19] See Y. Colin de Verdiere inLes Houches Lecture Notes SeriesLII (1989), (North-Holland, Amsterdam, 1989).

[20] A. M. Garcia-Garcia, J.D. Urbina, et al., in preparation.

[21] J. D. Urbina et al., Phys. Rev. Lett.97, 214101 (2006).

[22] W.D. Knight, et.al., Phys. Rev. Lett.52, 2141 (1984).

[23] M. Sieber, J. Phys. A: Math. Gen32, 7679 (1999).

[24] B. Janko, et al., Phys. Rev.B 50, 1152 (1994).

[25] P. W. Anderson: J. Phys. Chem. Solids11, 26 (1959).

Referenzen

ÄHNLICHE DOKUMENTE

The depopulation kinetics of the luminescence states, e.g., S 1 (LE) and S 1 (CT), have been measured detecting the fluorescence decay curve exciting the fluorescent sample at the

In view of the slow geometrical relaxation of excitation energy excess in PVA polymers, the relaxation time τ R for thermal equilibration between the luminescent molecule and

Russia’s strategic objective vis-à-vis Belarus in the domain of nuclear energy is primarily to block the implementation of any nuclear power plant projects by third parties (Western

[r]

Although the penalty for tax evasion applicable in past rounds is not relevant for current payoffs, indi- viduals who were used to a high penalty declared a higher share of income,

(1998), people’s ex-ante willingness to pay to reduce risk will tend to vary with their perceptions of the attitudes towards the characteristics of different hazards, such as

Damage costs and external costs of severe accidents in different energy chains were estimated, based on the unit cost values for the various types of consequences. Unit values for

TU Wien – Institute for Production Engineering and Laser Technology (project management) / Institute for Energy Systems and Thermo- dynamics / Institute of Computer Aided