Crossed Andreev Reflection Tuned by Dynamical Coulomb Blockade

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Crossed Andreev reflection tuned by dynamical Coulomb blockade

A. Kleine,∗ A. Baumgartner, J. Trbovic, and C. Schönenberger


Institute of Physics, University of Basel,
Klingelbergstrasse 82, 4056 Basel, Switzerland

(Dated: December 19, 2008)


We show experimentally that in nanometer scaled superconductor/normal metal hybrid devices
and in a small window of contact resistances, crossed Andreev reflection (CAR) can dominate the
nonlocal transport for all energies below the superconducting gap. Besides CAR, elastic cotun-
neling (EC) and nonlocal charge imbalance (CI) can be identified as competing subgap transport
mechanisms in temperature dependent four-terminal nonlocal measurements. We demonstrate a
systematic change of the nonlocal resistance vs. bias characteristics with increasing contact resis-
arXiv:0812.3553v1 [cond-mat.mes-hall] 18 Dec 2008

tances, which can be tuned in the fabrication process. For samples with higher contact resistances,
CAR is weakened relative to EC in the midgap regime, possibly due to dynamical Coulomb blockade.
Gaining control of CAR is an important step towards the realization of a solid state entangler.

Quantum mechanically entangled pairs of particles are transmissions [7], by spin-active interfaces or ferromag-
a major building block of quantum computation and netic contacts [8], by disorder, or by electron-electron
information processing. A natural source of entangled interactions [9, 10, 11, 12]. It has been suggested that
electron pairs is a BCS-type superconductor where the Coulomb interaction and the electromagnetic environ-
Cooper pairs (CP) form spin singlet states. The two ment are crucial for CAR to be the dominant process
electrons of a Cooper pair can be spatially separated in [13].
two different metallic leads in a nonlocal process called Experimentally, first signatures of CAR have been
crossed Andreev reflection (CAR) [1, 2, 3]. reported only recently for finite subgap bias in multi-
At temperatures (T ) well below the superconducting terminal hybrid structures [15, 16, 17]. Other reports
transition temperature, Tc , and for bias potentials be- emphasize the importance of EC and CI, or even that CI
low the superconducting energy gap ∆, electrons from a dominates [18, 19].
normal metal contact (N) can enter the superconductor Here we present the results of systematic experiments
(S) only as Cooper pairs by a process known as Andreev on a series of samples with different barrier resistances.
reflection (AR) [4]. In this local process a hole is re- We show that the earlier, seemingly contradicting results
flected into the same N to conserve momentum. If two can be reconciled in one coherent picture. In a small
normal metal contacts, N1 and N2, are spatially sepa- window of contact resistances CAR can dominate the
rated by less than the Ginzburg-Landau coherence length nonlocal transport for all subgap bias voltages. EC is
ξ (ξ ∼ 100 − 200 nm for Al), the two electrons form- favored for larger barrier resistances, while nonlocal CI
ing a CP can originate from different normal contacts, is suppressed.
see Fig. 1(a). This process opens an additional nonlo- In our experiments we use a four-terminal nonlocal
cal conduction path known as CAR. An inverse process measurement geometry: a bias U between the normal
was proposed as the basis of a solid-state entangler: the metal contact N1 (injector) and the grounded supercon-
electrons of a CP are split between the two leads while ductor lead S1 causes a current I as shown in Fig. 1(c).
retaining their entanglement from the superconductor. We measure both, I and the nonlocal potential differ-
However, this method of creating entangled particles can
be accompanied by two additional processes that lead to
a correlated signal on both, N1 and N2, but not to entan-
glement. In the first, a single electron from N1 can reach
the other contact N2 by elastic cotunneling (EC) [5, 6, 8],
see Fig. 1(b). In the second, called nonlocal charge im-
balance (CI), electrical charge can be transferred to the
second contact by the diffusion of quasi-particles gener-
ated by finite temperatures or finite bias.
Recently, the relative strength of these subgap pro-
cesses was the subject of extensive theoretical work. FIG. 1: (Color online) Schematics of (a) crossed Andreev
Standard BCS theory predicts that to lowest order in the reflection and (b) elastic cotunneling. (c) SEM image of a
tunneling rates, CAR and EC exactly cancel in normal typical sample. A current I is applied between normal contact
metal/insulator/superconductor (NIS) systems at low T N1 and the superconducting contact S1, while the nonlocal
and bias [6]. This cancelation can be lifted for higher voltage Unl is detected between N2 and S2.
2

ence Unl between the normal contact N2 (detector) and


TABLE I: Properties of samples A, B and C. ρ: specific re-
the second superconducting contact S2. In a CAR pro- sistance and ℓel : elastic mean free path of the Al strip at
cess a particle is injected from N1 and one of opposite T = 4.2 K; ξ: superconducting coherence length; Rinj , Rdet :
charge is generated in N2. In Fig. 1(a) such a process injector and detector barrier resistance at T = 1.5 K; ∆∗ :
is shown schematically for U < 0, where electrons are effective energy gap of the injector; d: injector-detector dis-
injected from N1 and holes are created in N2. For each tance.
U a potential of opposite sign builds up on N2 to ensure ρ ℓel ξ Rinj Rdet ∆∗ d
the zero current boundary condition of a voltage mea- [µΩcm] [nm] [nm] [kΩ] [kΩ] [meV] [nm]
surement. In EC and CI processes, however, the charges A 2.63 23 184 0.39 0.51 0.29 55
on the injector and detector have the same sign and Unl B 2.47 24 189 0.80 5.51 0.26 155
has the same polarity as the bias. In practice, we use
C 1.20 50 273 2.60 11.20 0.32 220
ac modulation techniques and investigate the nonlocal
ac ac
differential resistance Rnl = dUnl /dI ≈ Udet /Iinj , where
’ac’ denotes the amplitude of the corresponding variable.
Rnl is negative for CAR and positive for all other subgap 10 s delay between points (not shown). The dc coincide
processes. with the ac results, but suffer from a much worse signal-
The samples are planar multi-terminal hybrid struc- to-noise ratio. Another source of spurious Unl signals is
tures, see Fig. 1(c). They were fabricated on thermally resistive leakage into the detector. We reduce this leakage
oxidized Si wafers by e-beam lithography and using the by using voltage amplifiers with 1 GΩ input impedance
angle-evaporation technique in ultra-high vacuum with a and check for each device that a change to 100 MΩ has
base pressure of 10−10 mbar. Because of the low spin- no effect on our results. We note that for T > Tc all
orbit interaction, we chose Al as superconducting mate- nonlocal signals vanish, implying that they are related to
rial. The Al layer has a thickness of 50 nm and a width the superconductor and not to the measurement setup or
of 200-300 nm. We oxidized the Al between 3 and 15 min inhomogeneous current paths.
in situ in an oxygen atmosphere with a pressure in the In the following we present experiments on three sam-
range of 0.1 to 12 mbar to obtain the required tunnel bar- ples A, B, and C, which have a similar geometry, but
riers. Without breaking the vacuum, we tilted the sample differ in the contact resistances. The characteristics of
and deposited Pd as normal metal contacts with a typ- each sample are summarized in Table I.
ical thickness of 30-40 nm and a width of 90 nm. In our
Al/AlOx /Pd-samples we obtained edge-to-edge distances
d between the N electrodes of ∼ 50 nm.
We estimate the energy gap ∆∗ of the superconduc-
tor from the the voltage of maximum conductance in
superconducting tunneling spectroscopy (STS) measure-
ments on the individual barriers, see Table I. We note
that ∆∗ depends also on the barrier characteristics, which
might cause the deviation from the bulk energy gap ∆ ≈
200 µeV. We measured a critical temperature Tc ≈ 1.2 K
and a resistivity ρAl = 1.2 − 2.6 µΩcm of the Al lay-
ers at T = 4.2 K. We find an elastic mean free path
ℓel ≈ 20 − 50 nm and ξ ≈ 180 − 270 nm at cryogenic
temperatures [20]. Therefore the Al is in the ’dirty limit’
of superconductivity.
We performed nonlocal ac measurements using lock-in
amplifiers at a frequency of ∼ 10 Hz in a He3 -cryostat
with a base temperature of 0.23 K. The dc bias applied
ac
to N1, Udc , is modulated by Uinj ≈ 12 µVrms. Since the
junction resistances vary strongly with the applied bias,
leakage currents and capacitive cross-talk can produce
spurious signals which might resemble the expected char-
acteristics. To exclude cross-talk we repeated all mea- FIG. 2: (Color online) Results of sample A. (a) Rnl vs. bias for
surements at various frequencies. For all samples shown various temperatures with N1 as injector and N2 as detector.
here we find that the in-phase part of Unl is indepen- (b) STS characterization of contacts N1 and N2 at 0.23 K.
dent of the measurement frequency, while the capacitive The dashed lines label the bias corresponding to ∆∗ of the
part Ynl is negligible, see Fig. 2(c). For comparison we injector. (c) Frequency dependence of the in-phase nonlocal
simultaneously measure Unl in a DC configuration with signal Unl and the capacitive part Ynl .
3

velops with a positive maximum of Rnl ≈ 30 Ω. For


0.13 mV . |Udc | . ∆∗ /e with ∆∗ ≈ 0.31 meV, Rnl is
negative with minima at ±0.23 mV with Rnl ≈ −13 Ω.
The sign changes at Udc ≈ ±0.13 mV, independently of
temperature. For |Udc | > ∆∗ /e the signal is positive and
tends to zero for increasing bias, in contrast to samples A
and B. With increasing temperature all nonlocal signals
tend to zero, independently of the bias, with Rnl ≈ 0 for
T > Tc . At zero bias the signal decreases monotonically
with increasing T and vanishes already for T > 0.5 K,
see Fig. 4(c). A finite nonlocal signal only develops for
T well below Tc . From the STS measurements shown
in Fig. 4(b) we find that the two contact resistances are
FIG. 3: (Color online) Results of sample B. Rnl vs. bias quite different and much larger than in the previous sam-
voltage at various T with N1 as injector and N2 as detector. ples (Rinj = 2.6 kΩ and Rdet = 11.2 kΩ).
The inset shows the STS characterization of N1 and N2 at Based on the discussion above we interpret Rnl < 0
0.23 K. The dashed lines indicate the bias corresponding to
in the superconducting energy gap as CAR being the
∆∗ of the injector.
dominant process, while Rnl > 0 is attributed to EC or
CI. The latter two processes can be distinguished by the
temperature dependence of Rnl at a given bias. EC is
Figure 2(a) shows Rnl as a function of the bias voltage a second order coherent process which is reduced with
Udc at various temperatures for sample A. At base tem- temperature due to the loss of coherence by phonon in-
perature we observe a negative nonlocal resistance Rnl for teractions. CI, however, increases exponentially up to Tc
all subgap voltages Udc < 0.29 mV with Rnl ≈ −1.0 Ω. due to the decreased superconducting energy gap and the
This value does not change with temperature up to broadened thermal distribution and vanishes for T > Tc
T ≈ 500 mK. When increasing T further Rnl increases [18, 19]. At zero bias and low enough temperatures CI
and becomes positive for T ≥ 1.05 K. For voltages larger can be neglected.
than ∆∗ /e, Rnl is positive at all temperatures and in-
creases with T . At 1.6 K, i.e. for T > Tc , Rnl ≈ 0 for
all biases. The contact characterization of sample A at
base temperature is shown in Fig. 2(b) and exhibits a sig-
nificant subgap conductance for both, N1 and N2. The
dashed lines represent the bias corresponding to the en-
ergy gap ∆∗ of the injector, which coincides with the sign
change of Rnl . Within the energy gap both barriers ex-
hibit nearly the same normalized conductance values and
both contacts have similar resistances above Tc , 0.4 kΩ
and 0.5 kΩ. Figure 2(c) shows the frequency dependence
of the nonlocal signal as discussed above.
In Fig. 3 the bias dependence of Rnl of sample B is
plotted for several temperatures. The shape and ampli-
tudes of the curves are very similar to those of sample
A in Fig. 2(a), as is the temperature dependence. How-
ever, in sample B at base temperature we observe a local
maximum for Udc ≈ 0 with Rnl ≈ 0, whereas for fi-
nite subgap biases we find Rnl < 0 with two minima of
Rnl ≈ −1.1 Ω at Udc ≈ ±0.235 mV. Above Tc the nonlo-
cal signals vanish for all bias voltages. The inset shows
the STS characterization of the individual contacts at
0.23 K. In contrast to sample A, the normalized injec-
tor and detector conductances differ considerably in the
gap (∆∗ ≈ 0.255 meV) and are smaller than in sample A FIG. 4: (Color online) Results of sample C. (a) Rnl vs. bias
voltage for various T with N1 as injector and N2 as detector.
(Rinj = 0.8 kΩ and Rdet = 5.5 kΩ). The solid lines are guides to the eye and the dashed lines
Bias dependent measurements for sample C are shown show the voltage corresponding to ∆∗ of the injector. (b)
in Fig. 4(a). At base temperature a prominent local STS-measurements of contact N1 and N2 at 0.23 K. (c) Rnl
maximum in the middle of the superconducting gap de- at zero bias as a function of T .
4

For sample A we conclude that CAR is the dominant resistances, DCB sets in and shifts the balance to EC,
nonlocal transport process at base temperature and at all which compensates CAR at zero bias. At a finite subgap
subgap biases. From the characteristic strong increase of bias EC is weaker because the process can not be elastic
Rnl to positive values with T , we conclude that nonlo- anymore. With resistances of a few kΩ as in sample C,
cal CI becomes the dominant process at higher tempera- DCB becomes strong and inhibits CAR relative to EC.
tures. These findings also apply for sample B, except Since also CI is blocked by DCB, Rnl is reduced relative
around zero bias, where CAR and EC approximately to the other samples for |Udc | > ∆∗ /e. A possible ex-
compensate each other at low temperatures. At finite planation of the much larger signals in sample C is that
subgap bias, EC is weakened as reported in [16] and the DCB also affects AR and thus the local current stronger
characteristics become similar to sample A. In sample C, than the nonlocal processes, which leads to an increased
EC not only compensates CAR, but even dominates the nonlocal resistance.
nonlocal midgap transport. As in A and B, CAR domi- In conclusion we find in our experiments that for a
nates at finite subgap bias. In all samples CI dominates small window of contact resistances CAR can dominate
for bias potentials above the energy gap. In contrast to the transport characteristics for all subgap biases. Above
the first two samples, CI is strongly suppressed in sample the superconducting energy gap the CI rates are impor-
C for |Udc | >> ∆∗ /e. tant. We have shown that it is possible to tune CAR and
Our experiments suggest that the resistances of the EC with the help of the NIS interface resistances. The
injector and detector contacts determine the relative contribution of CAR to the subgap transport decreases
strength of the CAR and EC rates and whether CI plays around zero bias with raising the injector and detector
a significant role [21]. Based on this observation, we put contact resistances. We qualitatively explain the various
forward a qualitative explanation of our results which subgap transport characteristics by different dependences
might also clarify the variety of experimental findings in of the relevant processes on dynamical Coulomb block-
the literature. The possibility of measuring negative Rnl ade, which depends crucially on the contacts resistances.
is described by a model in [13], where the two normal However, to elucidate the exact mechanisms that lead to
contacts are electromagnetically coupled. This coupling our results, DCB has to be taken into account in further
can cause correlations between fluctuations on the nor- theoretical and experimental work.
mal metal contacts, favoring CAR or EC, depending on
We thank A. D. Zaikin, D. S. Golubev and A. Levy
how the modes of the contacts are coupled. However, on
Yeyati for fruitful discussions. This work is financially
a superconducting island EC is favored over CAR, be-
supported by the NCCR on Nanoscale Science and the
cause the Coulomb energy of an added Cooper pair has
Swiss-NSF.
to be supplied for CAR, while EC is only weakly affected
[13]. We expect a similar effect for dynamical Coulomb
blockade (DCB), where the response of the electromag-
netic environment of a normal contact to a charge on the
superconducting contact leads to a blockade for the next ∗
Electronic address: andreas.kleine@unibas.ch
charge trying to enter the superconductor. This block- [1] P. Recher, E.V. Sukhorukov, and D. Loss, Phys. Rev. B
ade strongly depends on the resistance of the contacts 63, 165314 (2001).
and becomes significant for resistances on the order of [2] J.M. Byers, and M.E. Flatté, Phys. Rev. Lett. 74, 306
(1995).
h/2e2 ≈ 12 kΩ. For much smaller resistances no DCB
[3] G. Deutscher, and D. Feinberg, Appl. Phys. Lett. 76, 487
occurs [14]. While CAR vanishes for completely trans- (2000).
parent contacts [7], AR is stronger suppressed than CAR [4] A.F. Andreev, Sov. Phys. JETP 19, 1228 (1964).
in the DCB regime on two independent (but spatially [5] D.V. Averin, and Y.V. Nazarov, Phys. Rev. Lett. 65,
close) contacts [14]. Though it has not been discussed 2446 (1990).
in the literature, we expect EC compared to CAR to be [6] G. Falci, D. Feinberg, and F.W.J. Hekking, Europhys.
less affected by DCB (similar as with classical Coulomb Lett. 54, 255 (2001).
[7] M.S. Kalenkov, and A.D. Zaikin, Phys. Rev. B 75,
blockade) since no charge is added on the superconduct-
172503 (2007).
ing contact. We also expect CI to be considerably af- [8] R. Mélin, and D. Feinberg, Phys. Rev. B 70, 174509
fected by DCB for analogous reasons as for CAR. (2004).
The systematic change of the measured nonlocal resis- [9] J.P. Morten, A. Brataas, and W. Belzig, Phys. Rev. B
tance vs. bias characteristics between our samples can 74, 214510 (2006).
thus be understood as follows: the contacts of sample A [10] S. Duhot, and R. Mélin, Phys. Rev. B 75, 184531 (2007).
[11] D.S. Golubev, and A.D. Zaikin, Phys. Rev. B 76, 184510
have the lowest resistances and DCB can be neglected,
(2007).
which allows CI and CAR to develop. The latter is fa- [12] M.S. Kalenkov, and A.D. Zaikin, Phys. Rev. B 76,
vored over EC possibly due to an electromagnetic cou- 224506 (2007).
pling of the injector and detector characteristic for all [13] A. Levy Yeyati, F.S. Bergeret, A. Martin-Rodero, and
our samples. For sample B with slightly larger contact T.M. Klapwijk, Nature Phys. 3, 455 (2007).
5

[14] P. Recher, and D. Loss, Phys. Rev. Lett. 91, 267003 [20] We used the following expressions: ξ ∼ = ξ0 ℓel , ξ0 =
(2003). ~vF /π∆(0), ℓel = 3D/vF and the diffusion constant
[15] D. Beckmann, H.B. Weber, and H. v. Löhneysen, Phys. D = σAl /e2 NAl . NAl = 2.4 · 1028 1/eVm3 represents the
Rev. Lett. 93, 197003 (2004). density of states of Al at the Fermi energy [Jedema et al.,
[16] S. Russo, M. Kroug, T.M. Klapwijk, and A.F. Morpurgo, Nature 416, 713 (2002)], vF = 1.3 · 106 m/s the Fermi ve-
Phys. Rev. Lett. 95, 027002 (2005). locity [Strunk et al., Phys. Rev. B 57, 10854 (1998)], σAl
[17] D. Beckmann, and H. v. Löhneysen, Appl. Phys. A 89, the conductivity of Al and e the electron charge.
603 (2007). [21] We exclude d as the crucial parameter as long as d < ξ
[18] P. Cadden-Zimansky, and V. Chandrasekhar, Phys. Rev. because sample C with the largest d exhibits the largest
Lett. 97, 237003 (2006). nonlocal signals, in contrast to what one would expect in
[19] P. Cadden-Zimansky, Z. Jiang, and V. Chandrasekhar, this case.
New J. Phys. 9, 116 (2007).

You might also like