• Aucun résultat trouvé

Chaotic dot-superconductor analog of the Hanbury Brown Twiss effect

N/A
N/A
Protected

Academic year: 2022

Partager "Chaotic dot-superconductor analog of the Hanbury Brown Twiss effect"

Copied!
5
0
0

Texte intégral

(1)

Article

Reference

Chaotic dot-superconductor analog of the Hanbury Brown Twiss effect

SAMUELSSON, Martin Peter, BUTTIKER, Markus

Abstract

As an electrical analog of the optical Hanbury Brown Twiss effect, we study current cross-correlations in a chaotic quantum dot-superconductor junction. One superconducting and two normal reservoirs are connected via point contacts to a chaotic quantum dot. For a wide range of contact widths and transparencies, we find large positive current correlations.

The positive correlations are generally enhanced by normal backscattering in the contacts.

Moreover, for normal backscattering in the contacts, the positive correlations survive when suppressing the proximity effect in the dot with a weak magnetic field.

SAMUELSSON, Martin Peter, BUTTIKER, Markus. Chaotic dot-superconductor analog of the Hanbury Brown Twiss effect. Physical Review Letters , 2002, vol. 89, no. 4, p.

046601.1-046601.4

DOI : 10.1103/PhysRevLett.89.046601 arxiv : cond-mat/0203188

Available at:

http://archive-ouverte.unige.ch/unige:4251

Disclaimer: layout of this document may differ from the published version.

(2)

arXiv:cond-mat/0203188v1 [cond-mat.mes-hall] 8 Mar 2002

Chaotic dot-superconductor analog of the Hanbury Brown Twiss effect

P. Samuelsson and M. B¨uttiker

D´epartement de Physique Theorique, Universit´e de Gen`eve, CH-1211 Gen´eve 4, Switzerland.

(February 1, 2008)

As an electrical analog of the optical Hanbury Brown Twiss effect, we study current cross- correlations in a chaotic quantum dot-superconductor junction. One superconducting and two normal reservoirs are connected via point contacts to a chaotic quantum dot. For a wide range of contact widths and transparencies, we find large positive current correlations. The positive cor- relations are generally enhanced by normal backscattering in the contacts. Moreover, for normal backscattering in the contacts, the positive correlations survive when suppressing the proximity ef- fect in the dot with a weak magnetic field.

In quantum theory, identical particles are indistin- guishable. Under exchange of any pair of particles, the many-body wavefunction remains invariant up to a sign, positive for bosons and negative for fermions. This ex- change symmetry was used in the pioneering interferom- eter experiment with photons, by Hanbury Brown and Twiss [1]. Several theoretical works [2–4] have suggested different analogs of this experiment with electrons in mesoscopic multiterminal conductors. It has been shown [3] that the fermionic statistics of electrons leads to neg- ative correlations between currents flowing in different terminals. Such negative correlations were also recently observed experimentally [5].

When normal conductors are connected to a super- conductor, correlations are introduced between electrons and holes due to Andreev reflections at the normal- superconductor interface, a phenomenon known as the proximity effect. The influence of the proximity effect on the current auto-correlations, i.e. the shotnoise, in a two-terminal diffusive normal-superconductor junctions was recently studied [6,7].

In multiterminal conductors, Andreev reflection can lead to positive cross correlations between currents flow- ing in the contacts to the normal reservoirs [8,9]. So far, positive correlations have been predicted only for single mode junctions [8–10]. Moreover, in multiterminal dif- fusive junctions it was found that cross correlations are negative in the absence of the proximity effect [11].

This raises two important questions: i) are the positive correlations in normal-superconducting junctions a large effect, of the order of the number of modes in multimode junctions, and if this is the case ii) is the proximity effect necessary to obtain these positive correlations? In this paper we give an answer to these two questions. The posi- tive correlations are large, and surprisingly, get enhanced by normal backscattering at the normal-superconducting interface. Moreover, positive correlations can exist even in the absence of the proximity effect, if the normal- superconductor interface is nonideal.

We study the current correlations in a system consist- ing of a chaotic quantum dot connected via point con- tacts to one superconducting and two normal reservoirs.

Systems consisting of chaotic dots coupled to supercon- ductors have recently [12] attracted a lot of interest. The generic properties of the model makes our result qualita- tively relevant for multiterminal normal-superconducting structures with random scattering.

ΓN N, ΓN

M, ΓS

V =0S V =V1

V =V2

N,

S N

N

1

2

FIG. 1. A chaotic quantum dot (grey shaded), acting as a beam-splitter, is connected to two normal reservoirs (N1 and N2) and one superconducting reservoir (S) via quantum point contacts.

A schematic picture of the system is shown in Fig. 1.

A quantum dot is connected to two normal reservoirs (N1 andN2) and one superconducting reservoir (S) via quantum point contacts. The contacts to the normal and superconducting reservoirs have mode independent transparency ΓN and ΓS respectively and supportNand M transverse modes. The conductances of the point contacts are much larger than the conductance quanta 2e2/h, i.e. NΓN, MΓS≫1, so Coulomb blockade effects in the dot can be neglected. The two normal reservoirs are held at the same potentialV and the potential of the superconducting reservoir is zero.

We consider the case where the classical motion in the dot is chaotic on time scales longer than the er- godic time τerg. The quasiparticle dwell time in the dot, ¯h/ET h, is assumed to be much larger thanτerg, but much smaller than the inelastic scattering time. Here ET h = (2NΓN +MΓS)δ/π, where δ is the mean level spacing in the dot. Under these conditions random ma- trix theory [14] can be used to describe the statistical properties of the scattering matrix Sd of the dot. The

(3)

scattering matrixS of the combined system, dot and su- perconductor, can be expressed [14] in terms of the scat- tering matrix Sd of the dot and the Andreev reflection amplitude at the contact-superconductor interface.

Due to the random scattering in the dot, the cur- rent Ii(t) in contact i fluctuates around its quantum statistical average ¯Ii. We study the zero-frequency spectral density of the current cross-correlationsP12 = 2R

dt∆I1(t)∆I2(0) where ∆Ij(t) =Ii(t)−I¯i. The corre- lation can be expressed in terms of the scattering matrix S. We consider the limit of zero temperature and a po- tentialeV much lower thanET hand ∆, where the energy dependence of the scattering matrixS can be neglected.

Moreover, at |E|<∆ no quasiparticles can escape into the superconductor and the scattering matrixSdescribes only scattering between the normal reservoirs,

S=

See Seh She Shh

, Sαβ=

Sαβ11 S12αβ

Sαβ21 S22αβ

, (1) where Sijαβ are matrix amplitudes (N ×N) for injected quasiparticles (e or h) of type β in lead j to be back reflected as quasiparticles of typeαin leadi.

Noting that the current fluctuation is just the sum of the fluctuations of electron and hole currents, the noise power can be conveniently written [8],

P12=P12ee+P12hh+P12eh+P12he (2) where the noise powerP12αβfor correlation between quasi- particle currents, is given by

P12αβ=σV4e3 h

X

i,j=1,2

trh

(Sαe1i)S1αhj(S2βhj)S2βei

i, (3)

with σ = +(−) for α = β (α 6= β). The expression for the correlations in Eqs. (2)-(3) is an extension of the result [3] for a purely normal conductor, taking into account both electron and hole quasiparticles. However, unlike the normal cross correlations, which are manifestly negative, the cross-correlationP12, can be positive, be- cause the correlations between different types of quasi- particles, P12eh+P12he, are positive. Note however that P12ee+P12hh−(P12eh+P12he), the cross-correlator between probability currents, is manifestly negative [8], a conse- quence of the ferminonic statistics of the quasiparticles.

The ensemble averaged correlations hP12i are calcu- lated using the statistical properties of the scattering ma- trixSd of the dot. We first consider the case with ideal contacts ΓNS = 1 and no magnetic field in the dot, i.e. conserved time reversal symmetry. In this case it is useful to decomposeSd as (see e.g. Ref. [14])

Sd=

U 0 0 U

r2N,2N t2N,M

tM,2N rM,M

UT 0 0 U′T

, (4) where U(U) is a unitary matrix of dimension 2N × 2N (M × M), uniformly distributed in the unitary

group. The diagonal matrices r2N,2N and rM,M have min(2N, M) elements√

1−Tnand the rest unity (t2N,M

andtM,2N follow from the unitarity ofSd). HereTn are the transmission eigenvalues, which have a density [16]

ρ(T) = (2N +M)/(2π)√

T−Tmin/(T√

1−T), where Tmin = [(2N−M)/(2N+M)]2 is the smallest possible eigenvalue.

Inserting the decomposition into the expression for the total scattering matrix S we find [14] the quasi- particle scattering amplitudes Sαβ in Eq. (1) as e.g See=U√

1−T /(2−T)U(T = 1−r22N,2N). Summing over injection contacts in the expression for the individual quasiparticle current correlators in Eq. (3) and inserting the scattering amplitudes we get

P12ee=V4e3

h tr[U(1− R)UC1UTRUC2] (5) and similar for the other termsP12αβ. HereR=T2/(2− T)2are eigenvalues of the matrix product (Seh)Sehand the matrixC1 is diagonal with elements (C1)n = 1 for n≤N and 0 otherwise. The matrixC2 = 1−C1. The ensemble average of Eq. (5) is carried out in two steps.

First,P12αβ is averaged over the unitary matrix U, using the diagrammatic technique in Ref. [15]. This gives to leading order in N, M (i.e neglecting weak localization corrections) for the total current corellations in Eq. (2),

hP12iU =V4e3 h

tr[R(1− R)]−tr(R)tr(1− R) 4N

. (6) We then perform the average over transmission eigenval- ues by integratinghP12iU weighted by the densityρ(T).

Using that to leading order in N we have htr(R)tr(1− R)iT =htr(R)iThtr(1−R)iT, we get (< .. >=< .. >U,T)

hP12i P0

= (1 +γ)22

1

1 +γ −2γ

f −f2−16γ3 f5/2

(7) wheref(γ) = 1 + 6γ+γ2 andγ= 2N/M, the ratio be- tween the conductances of the point contacts connected to the normal and the superconducting reservoirs. The correlation is normalized withP0 = 4V N e3/h, i.e. it is large, of orderN. The correlationhP12i, plotted in Fig.

2, is positive for a dominating coupling to the supercon- ductor, but crosses over (atγ ≈0.5) to negative values upon increasing the coupling to the normal reservoirs.

In the general case, it is difficult to provide a simple explanation for the sign and magnitude of the correla- tionshP12i, since they result from a competition between hP12eei+hP12hhiandhP12ehi+hP12hei, which have opposite sign and generally are of the same magnitude. It is however possible to get a qualitative picture in certain limits by studying the different contributing scattering processes.

(4)

10−2 10−1 100 101 102 103

−0.06

−0.04

−0.02 0 0.02 0.04 0.06 0.08 0.1 0.12

10 1 P /P0

100

12

0.1 2N/M 0.01

-0.04 0 0.04 0.08

FIG. 2. The ensemble averaged current cross-correlation hP12i with (solid) and without (dashed) proximity effect in the dot, as a function ofγ= 2N/M. The correlations in the presence of the proximity effect are positive for a dominating coupling to the superconductor (γ < 0.5), but are negative for all γ when suppressing the proximity effect with a weak magnetic field in the dot.

In the limit γ →0, the coupling to the normal reser- voirs is negligible and a gap in the spectrum opens up around Fermi energy in the dot [12]. As a consequence, quasiparticles injected from one normal reservoir are An- dreev reflected, effectively direct at the contact-dot inter- face [13], with unity probability back to the same reser- voir. There is thus no partition of incoming quasiparticles and hence no noise,P12= 0.

Increasing the coupling to the normal reservoir, the probability of normal reflection(∼γ) as well as cross An- dreev reflection (∼γ2), from one reservoir to the other, becomes finite. Since normal reflection is the dominant process, we can neglect the terms inP12 in Eq. (3) con- taining the cross Andreev reflection amplitude (e.g. S12eh).

This giveshP12i= 8V(e3/h)htr(S12ee(S12ee)S11eh(S11eh))i= 2P0γ, positive since only terms in hP12ehi+hP12hei con- tribute. This expression shows that the correlations can be explained as partition noise of injected electrons, which have probability<

∼1 to be Andreev reflected (effec- tively at the contact-dot interface) and probability ∼γ to be normally reflected.

In the opposite limit,γ≫1, the coupling to the super- conductor is weak and the Andreev reflection probability is small, of order 1/γ. The correlation can be written as hP12i=−16V(e3/h)htr(S11ee(S11ee)S21eh(Seh21))i=−P0/γ, which is negative since we find that in this limit only terms in hP12eei+hP12hhicontribute. This shows that the correlation can be explained as partition noise of elec- trons which have probability <

∼ 1/2 to be normally re- flected (without reaching the dot-superconductor con- tact) and probability∼1/γ to be Andreev reflected.

When the proximity effect is suppressed by a weak magnetic field in the dot, [13] it is no longer possible to express the correlations directly in terms of the trans-

mission eigenvalues T, as in Eq. (6). We instead use the fact that the scattering matrixSd of the dot itself is uniformly distributed in the unitary group, without con- straints on symmetry ofSd. Since the scattering matrix amplitudesSαβin Eq. (3) can be expressed [14] in terms ofSd, the correlationsP12αβcan be directly averaged over Sd using the diagrammatic technique in Ref. [15]. This gives for the total correlations

hP12i

P0 =−γ2(1 +γ)

(2 +γ)4 , (8)

plotted in Fig. 2. The gap in the spectrum is suppressed and the scattering processes responsible for the positive correlations in the presence of the proximity effect are strongly modified, most notably in the limitγ≪1. As a consequence, the correlations are manifestly negative for allγ, i.e. there are no positive correlations in the absence of the proximity effect, similar to what was found for a metallic diffusive system [11].

Until now we have only considered ideal point contacts, with ΓNS = 1. In an experimental situation, it is often difficult to obtain an ideal contact between the dot and the superconductor. It is therefore of interest to study the situation with a nonideal dot-superconductor contact. We consider first the case with proximity effect in the dot. To calculate the current correlations in this case we note [14] that a nonideal interface changes the density of transmission eigenvalues, ρ(T), but not the distributions of the unitary matrix U in Eq. (5). The transmission eigenvalue density is calculated numerically [16,15] for different contact transparencies ΓS < 1 and the integrals in Eq. (6) are then performed.

The resulting correlations are plotted in Fig. 3. Sur- prisingly, the main effect of normal backscattering at the dot-superconductor contact is to cause a crossover from negative to positive correlation for a strong coupling to the normal reservoirs. In this limit, γ ≫ 1, injected quasiparticles undergo at the most one scattering event at the dot-superconductor contact before leaving the junc- tion. The expression for the scattering matricesSijαβ in Eq. (3) simplify considerably and we can derive an ana- lytical expression for the correlations, giving [17]

hP12i P0

= 1

γReh(1−2Reh) (9) where Reh = Γ2S/(2 − ΓS)2 is the Andreev reflec- tion probability of quasiparticles incident in the dot- superconductor contact. There is a crossover from neg- ative to positive correlations already for Reh = 1/2, i.e ΓS = 2(√

2−1) ≈0.83, in agreement with the full nu- merics in Fig. 3.

Since ΓS<1, nowRehis smaller than one. As a conse- quence, there is additional partition due to the possibility of normal reflection at the dot-superconductor contact.

(5)

The partition noise discussed above, from electrons be- ing either Andreev reflected or normally reflected (with- out reaching the dot-superconductor contact), is thus re- duced to −P0Reh/γ. However, the additional normal reflection at the dot-superconductor contact give rise to a noise term 2P0Reh(1−Reh)/γ, with opposite sign [to- gether they give Eq. (9)]. For Reh < 1/2, the second term is dominating, causing the crossover to positive cor- relations.

10−2 10−1 100 101 102

−0.06

−0.04

−0.02 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14

Γ

N

S

2

N1 S

P /P0

2N/MΓS

12

0 0.12

1

0.1 10 100

-0.04 0.04 0.08

FIG. 3. Ensemble averaged current cross correlations P12 in the presence of normal backscattering at the nor- mal-superconducting interface (see inset). The contact trans- parencies ΓSare 1 (solid), 0.8 (dash-dotted) 0.6 (dashed) and 0.4 (dotted). The correlations are plotted as a function of 2N/(MΓS). Note that the correlations for a dominating cou- pling to the normal reservoirs crosses over from negative to positive onincreasingthe normal backscattering.

Interestingly, in the absence of a proximity effect in the dot, we find in the same way that the correlation for γ≫ 1 is also given by Eq. (9), the argument being the same as in the presence of the proximity effect. The proximity effect thus plays no role in this limit, where the quasiparticles undergo at most one Andreev reflection.

This shows that the proximity effect is not a necessary condition for positive correlations.

Finally, we note that the effect of normal backscatter- ing in the contacts between the dot and the normal reser- voirs is to enhance the positive correlations for a domi- nating coupling to the superconducting reservoir. In the limit with tunnel barriers in all contacts, the transmis- sion eigenvalue density is known analytically [16] and the correlationhP12ifollows from Eq. (6) [18]

hP12i P0ΓN

= ¯γ

(1 + ¯γ2)3/2

1−5 ¯γ2 (1 + ¯γ2)2

, (10) where ¯γ= 2NΓN/(MΓS). The correlations are thus pos- itive for ¯γ <(√

5−1)/2 and ¯γ >(√

5+1)/2, and negative for intermediate values.

In conclusion, we have studied the current cross- correlations in a three terminal superconducting-chaotic

dot analog of the Hanbury Brown Twiss interferometer.

We find that the correlations are positive for a wide range of junction parameters, and survive even in the absence of a proximity effect in the dot. The magnitude of the positive correlations is large, proportional to the num- ber of transport modes in the contacts to the dot, which should simplify an experimental observation.

We acknowledge discussions with C.W.J. Beenakker, W. Belzig, E. Sukhorukov, and G. Johansson. The work was supported by the Swiss National Science founda- tion and the program for Materials with Novel Electronic Properties. In the final stages of this work we became aware of Boerlin et al. [19], who investigate a three ter- minal junction in the tunnel limit. In this limit, our results coincide.

[1] R. Hanbury Brown and R.Q. Twiss, Nature 177, 27 (1956).

[2] M. B¨uttiker, Phys. Rev. Lett.652901 (1990), Th. Martin and R. Landauer, Phys. Rev. B451742 (1992).

[3] M. B¨uttiker, Phys. Rev. B4612485 (1992).

[4] For a review, see Ya. M. Blanter and M. B¨uttiker, Phys.

Rep.3361 (2000).

[5] M. Henny et al. Science284, 296 (1999), W.D. Oliwer et al. Science284299 (1999); S. Oberholzer et al., Physica E.6314 (2000).

[6] X. Jehl et al., Nature (London) 405 50 (2000), A.A.

Kozhevnikov, R.J. Shoelkopf, and D.E. Prober, Phys.

Rev. Lett.843398 (2000).

[7] W. Belzig and Y. Nazarov, Phys. Rev. Lett.87 067006 (2001).

[8] M. P. Anatram and S. Datta, Phys. Rev. B 53 16390 (1996).

[9] J. Torr´es and Th. Martin, Eur. Phys. J. B12319 (1999).

[10] T. Gramespacher and M. B¨uttiker, Phys. Rev. B61, 8125 (2000).

[11] K. Nagaev and M. B¨uttiker, Phys. Rev. B 63 081301, (2001).

[12] J. Melsen et al. Europhys. Lett.35, 7 (1996), D. Taras- Semchuk and A. Altland, Phys. Rev. B64014512 (2001).

[13] A. A. Clerk, P.W. Brouwer and V. Ambegaokar, Phys.

Rev. B6210226 (2000).

[14] C. W. J. Beenakker, Rev. Mod. Phys,69, 731 (1997).

[15] P.W. Brouwer, and C.W.J Beenakker, C.W.J. J. Math.

Phys.37, 4904. (1997)

[16] Yu. V. Nazarov, in Quantum Dynamics of Submicron Structures, ed. H.A. Cerdeira et al, NATO ASI Series E291, 687 (1995).

[17] We note that this result coincides with what is found for the single mode model of Ref. [9], upon substituting ΓS= 2ǫ.

[18] The same result was obtained in Ref. [19], see comment at the end of the paper.

[19] J. Boerlin, W. Belzig, and C. Bruder, cond-mat/0201579

Références

Documents relatifs

Firstly Kondo effect is associated with the strong coupling fixed point of Anderson impurity model and is very stable to perturbations and hence to show that Majorana fermions

Ce qui est int´ eressant dans cette mesure, c’est que c’est pour nous la mesure la plus facile qu’on puisse faire avec le nouveau d´ etecteur que avons install´ e.. C’est un

Although the effect is easily explained in the context of classical wave optics, as in fig- ure 1, the HBT experiment stimulated deep questions about the quantum description of

For intellectual curiosity, the same PL measurements of DCM dissolved in various deuterated form of the select solvents were performed. The peak PL emissions

Le mouvement d’un bras du corps humain est complexe puisqu’ il est constitué d’un assemblage d’éléments en mouvement tridimensionnel. La modélisation de ce bras est

En effet le flux de métaux transportés par les eaux de surface drainant des districts miniers ne sont pas sans un impact sur les sites ‘avals’ et peuvent

Dans ce cadre, et en supposant l’existence d’une taxonomie sur tous les items de la base de donn´ ees, nous utilisons une mesure de similarit´ e Overall-Relatedness permet- tant

Our work consists to calculate the binding energy of an impurity in CdSe quantum dot, placed at the center, with the absence and the presence of the confined LO- phonon interaction as