This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Fermi Velocity Dependent Critical Current in Ballistic Bilayer Graphene Josephson Junctions

Amis Sharma1, Chun-Chia Chen2, Jordan McCourt2,
Kenji Watanabe3, Takashi Taniguchi3, François Amet4, Gleb Finkelstein2,
Ivan Borzenets1,∗
1Department of Physics and Astronomy, Texas A&\&M University, College Station, 77843, TX, USA
2Department of Physics, Duke University, Durham, 27701, NC, USA
3Advanced Materials Laboratory, NIMS, Tsukuba, 305-0044, Japan
4Department of Physics and Astronomy, Appalachian State University, Boone, 28607, NC, USA
Corresponding Author. E-mail: [email protected]
Abstract

We perform transport measurements on proximitized, ballistic, bilayer graphene Josephson junctions (BGJJs) in the intermediate-to-long junction regime (L>ξL>\xi). We measure the device’s differential resistance as a function of bias current and gate voltage for a range of different temperatures. The extracted critical current ICI_{C} follows an exponential trend with temperature: exp(kBT/δE)\exp(-k_{B}T/\delta E). Here δE=νF/2πL\delta E=\hbar\nu_{F}/2\pi L: an expected trend for intermediate-to-long junctions. From δE\delta E, we determine the Fermi velocity of the bilayer graphene, which is found to increase with gate voltage. Simultaneously, we show the carrier density dependence of δE\delta E, which is attributed to the quadratic dispersion of bilayer graphene. This is in contrast to single layer graphene Josephson junctions, where δE\delta E and the Fermi velocity are independent of the carrier density. The carrier density dependence in BGJJs allows for additional tuning parameters in graphene-based Josephson Junction devices.

Ballistic graphene Josephson junctions (GJJs) have been widely utilized as a platform to study various interesting physics that arise at low temperatures from interactions between superconductors and normal metals [1, 2, 3, 4, 5, 6]. The ballistic superconductor-normal metal-superconductor Josephson junction (SNSJJ) hosts Andreev bound states (ABS) which carry supercurrents across the normal region of the JJ; a disorder-free weak link and high transparency at the SN interface are necessary. Hexagonal Boron-Nitride (hBN) encapsulated graphene as the weak link enables highly transparent contacts at the interface whilst keeping graphene clean throughout the fabrication process [7]. Here, we study proximitized, ballistic, bilayer graphene Josephson junctions (BGJJs). Bilayer graphene devices (in contrast to monolayer) allow extra potential tunability via a non-linear dispersion relation, applied displacement field, or lattice rotation [6].
The critical current (ICI_{C}) of SNSJJ in the intermediate-to-long regime, where the junction length (L) \geq superconducting coherence length (ξ0\xi_{0}), scales with temperature (T) as IC=exp(kBT/δE)I_{C}=exp(-k_{B}T/\delta E). Here, δE=νF/2πL\delta E=\hbar\nu_{F}/2\pi L, an energy scale related to ABS level spacing [8, 9, 10, 11, 12]. Note that in the intermediate regime (Lξ0L\approx\xi_{0}) δE\delta E is found to be suppressed [4]. A previous study of GJJs found that in this regime, the relation is held more precisely when ξ\xi was taken into account along with L, that is: δE=νF/2π(L+ξ)\delta E=\hbar\nu_{F}/2\pi(L+\xi) [8]. Monolayer graphene displays a linear dispersion relation, which results in a constant fermi velocity (νF0\nu_{F0}). Thus, in ballistic GJJs, δE\delta E remains independent of the carrier density. In comparison, bilayer graphene displays a quadratic dispersion relation at low energies. In BGJJs we studied, a back-gate voltage (VGV_{G}) controls the carrier density; and δE\delta E dependence on VGV_{G} is observed. Using δE\delta E, we extract Fermi velocity in bilayer graphene: it is seen that νF\nu_{F} increases with VGV_{G}, and saturates to the constant value, νF0\nu_{F0}, of the monolayer graphene.
Our device consists of a series of four terminal Josephson junctions (on SiO2/Si\mathrm{S}iO_{2}/Si substrate) made with hBN encapsulated bilayer graphene contacted by Molybdenum-Rhenium (MoRe) electrodes. Bilayer graphene is obtained via the standard exfoliation method. It is then encapsulated in hexagonal Boron-Nitride using the dry transfer method [13]. MoRe of 8080 nm thickness is deposited via DC magnetron sputtering. The resulting device has four junctions of lengths 400400 nm, 500500 nm, 600600 nm, and 700700 nm. The width of the junctions is 4μ4~{}\mum. The device is cooled in a Leiden cryogenics dilution refrigerator operated at temperatures above 11 K, and the measurements were performed using the standard four-probe lock-in method. A gate voltage VGV_{G} is applied to the SiSi substrate with the oxide layer acting as a dielectric, which allows modulation of the carrier density.[14, 15, 4, 8, 16, 17].

Refer to caption
Figure 1: (a) Differential resistance versus gate voltage (VGV_{G}) and bias current IBiasI_{Bias} taken at T=1.37T=1.37 K. The black region around zero bias corresponds to the superconducting state. IBiasI_{Bias} is swept from negative to positive. Thus, the transition at negative bias corresponds to the re-trapping current IRI_{R}, while the transition at positive bias is the switching current ICI_{C}. (b) Vertical line cut of the resistance map taken at VG=5V_{G}=5 V, T=1.37T=1.37 K, showing device’s differential resistance versus bias current.

Fig. 1(a) displays the differential resistance (dV/dIdV/dI) map of the 400400 nm junction at T=1.37T=1.37 K; we see zero resistance (black region) across all applied VGV_{G} indicating the presence of supercurrent. As the bias current IbiasI_{bias} is swept from negative to positive values, the junction first reaches its superconducting state at a value |Ibias|=IR|I_{bias}|=I_{R}, known as re-trapping current. Then, as |Ibias||I_{bias}| is increased to higher positive values, the junction transitions to the normal state at |Ibias|=IS|I_{bias}|=I_{S}, known as switching current. Fig. 1(a) shows that the junction can sustain larger region of critical current as we modulate the carrier density to higher values via VGV_{G}. Fig. 1(b) displays a line plot extracted from the dV/dIdV/dI map which shows hysteresis in IRI_{R} and ISI_{S}. This is a commonly observed phenomenon in underdamped junctions [18, 14], or can also be attributed self-heating [15, 19, 16]. The critical current ICI_{C} of the junction is approximately ISI_{S} as found in switching statistics measurements, discussed in previous publications [8, 20, 21, 22].
Extracting the critical current ICI_{C} from the differential maps for different temperatures, we can see that ICI_{C} falls exponentially with inverse TT (Fig. 2c) We also extract conductance of the junction in the normal regime (IBiasICI_{Bias}\gg I_{C}). Fig. 2(b) shows this conductance (GG) for 400400 nm junction device. We find that conductance GG scales as the square-root of VGV_{G} as seen from the fit (blue curve). However, due to large contact resistance (RCR_{C}) of the device, the measured conductance GG is suppressed compared to the ballistic limit expectation. Therefore, to demonstrate the ballistic nature of the device, we present normal resistances (RNR_{N}) of junctions of length 500500 nm, 600600 nm, and 700700 nm (Fig. 2(c) inset). Taking into account RCR_{C}, we plot extracted values versus VGV_{G} from the Dirac point. The inset plot shows that the RNRCR_{N}-R_{C} values are independent of junction length, which demonstrates the ballistic nature of the devices.

Refer to caption
Figure 2: (a) Device picture. Image shows series of junctions with different Lengths: 400400 nm, 500500 nm, 600600 nm and 700700 nm. (b) The ballistic conductance vs Gate voltage for L=400L=400 nm junction. The inset shows junction resistance minus the parasitic contact resistance plotted against gate voltage from the Dirac point for all our devices. (c) Critical currents ICI_{C} of L=400L=400 nm junction plotted against temperature TT, for various gate voltages, on a semi-log scale. The plots show VGV_{G} dependence of ICI_{C}: the gray lines show that the slope of the curve for the lowest plotted gate VG=8V_{G}=8 V, is smaller than the slope of the highest plotted gate VG=21V_{G}=21 V.

To extract δE\delta E of the junction, we go to the discussion of ICI_{C} vs temperature trends in Fig. 2(c). Here the y-axis is plotted in log scale. From the slope of curves Log(IC)=(kB/δE)T\mathrm{Log}(I_{C})=-(k_{B}/\delta E)T for each gate, one can extract δE\delta E versus VGV_{G} (plotted in Fig. 3a). Unlike for the case of monolayer graphene, a clear dependence on VGV_{G} is seen. The energy δE\delta E scales linearly with Fermi Velocity vFv_{F} (Fig. 3b). Note that calculating vFv_{F} from δE\delta E for junctions in the intermediate regime requires knowledge of the superconducting coherence length ξ\xi. In our case ξ\xi is obtained from the fit discussed below.
We now compare the experimentally obtained δE\delta E (and vFv_{F}) to the theoretical expectation. Assuming a quadratic dispersion relation F=2kF22m\mathcal{E}_{F}=\frac{\hbar^{2}k^{2}_{F}}{2m^{*}}, it follows that the expression for the Fermi velocity is: vF=2EFmv_{F}=\sqrt{\frac{2E_{F}}{m^{*}}} [23, 24, 25]. The Fermi Energy F\mathcal{E}_{F} for bilayer graphene scales as: F=2π|n|2m\mathcal{E}_{F}=\frac{\hbar^{2}\pi|n|}{2m^{*}}. With mm^{*} being the effective mass of electrons in graphene. The carrier concentration nn, controlled by the applied gate voltage VGV_{G}, is given by n=VGVdeCTotaln=\frac{V_{G}-V_{d}}{e}C_{Total} with VdV_{d} as the gate voltage at the Dirac point. The total capacitance CTotalC_{Total} is a combination of quantum capacitance CqC_{q} and gate oxide capacitance CoxC_{ox}: CTotal=[1Cox+1Cq]1C_{Total}=\left[\frac{1}{C_{ox}}+\frac{1}{C{q}}\right]^{-1}. The quantum capacitance CqC_{q} for bilayer graphene is determined by Cq=2e2mπ2C_{q}=\frac{2e^{2}m^{*}}{\pi\hbar^{2}}. The gate oxide capacitance per unit area is Cox=ϵ0ϵrdC_{ox}=\frac{\epsilon_{0}\epsilon_{r}}{d}, where ϵ0\epsilon_{0} is vacuum permittivity, ϵr\epsilon_{r} is the relative permittivity of the oxide, and dd is oxide layer thickness. For Silicon oxide gate with d=300d=300 nm we get Cox115μF/m2C_{ox}\approx 115\mu F/m^{2}. Thus, the full expression for the Fermi velocity vFv_{F} is:

vF=2πϵ0ϵre|VGVd|m(2de2m+πϵ0ϵr2)\displaystyle v_{F}=\hbar\sqrt{\frac{2\pi\epsilon_{0}\epsilon_{r}e|V_{G}-V_{d}|}{m^{*}(2de^{2}m^{*}+\pi\epsilon_{0}\epsilon_{r}\hbar^{2})}} (1)

Note that the effective mass mm* typically ranges from 0.024me0.024~{}m_{e} to 0.058me0.058~{}m_{e} for 11012410121*10^{12}\sim 4*10^{12} carriers/cm2cm^{2} [26], where mem_{e} is the electron rest mass. Moreover, since mm* has a carrier concentration dependence, we assume a linear shift of mm* with the gate voltage as: m=mi+dm(VGVd)m^{*}=m_{i}+dm(V_{G}-V_{d}) [26].
Experimental data provides us with the following: δE(VG)=2π(L+ξ)vF\delta E(V_{G})=\frac{\hbar}{2\pi(L+\xi)}v_{F}. To fit this data, the model is set as :δE(VG)=(mi,dm,ξ,Vd,d)\delta E(V_{G})=\mathcal{F}(m_{i},dm,\xi,V_{d},d) where mi,dm,ξ,Vd,dm_{i},dm,\xi,V_{d},d are the fitting parameters, and VGV_{G} is the independent variable. (We use the as-designed length of the device LL, and take ϵr=3.9\epsilon_{r}=3.9 for SiO2SiO_{2}.)
The resulting fit of the data from the 400nm400~{}nm junction for both δE\delta E and vFv_{F} is plotted as solid lines in Fig. 3(a) and Fig. 3(b) respectively. Moreover, taking the fitted ξ\xi, we calculate the Fermi velocity vFv_{F} for all other junctions on the same substrate. As seen from Fig. 3(b), the calculated vFv_{F} of all devices is in good agreement with the fit obtained from the 400400 nm junction. The fitted parameters are summarized in Table 1. All fall within the range of expected values, with ξ\xi being consistent with previously measured values for graphene/MoRe junctions.

Parameter Fitted value Expected Value
ξ\xi 460.34460.34 nm 300500300\sim 500 nm
dd 317.71317.71 nm 300330300\sim 330 nm
mim_{i} 0.020me0.020~{}m_{e} 0.020.06me0.02\sim 0.06~{}m_{e}
dmdm 0.0003me0.0003~{}m_{e}/V 0.00020.0005me0.0002\sim 0.0005~{}m_{e}/V
VdV_{d} 1.99-1.99 V ±2\approx\pm 2 V
Table 1: The fitting parameters used to match the measured δE\delta E, and consequently the Fermi velocity vFv_{F}, versus gate to the theoretical expectation described in Equation 1. We see that resulting fitted values match closely to what is expected. The expected gate dielectric thickness dd is estimated from the substrate specifications plus the bottom hBN thickness. The expected Dirac point voltage VDV_{D} is ontained from the resistance map. The expectations for the superconducting coherence length ξ\xi and the effective mass mim_{i} are obtained from previous works[26, 4].
Refer to caption
Figure 3: (a) Energy δE\delta E extracted from the slope of log(ICI_{C}) vs T plotted against the gate voltage VGV_{G} from the Dirac point of the junction with L=400L=400 nm. We see δE\delta E dependence on the carrier density modulated via the gate voltage for the junction. (b) Fermi velocity (vfv_{f}) calculated from δE\delta E using the device dimensions, and the superconducting coherence length ξ\xi obtained from the fit to theory. The solid line represents the theoretical trend as fitted to the data for the L=400L=400 nm junction. In addition, panel (b) shows calculated vFv_{F} for the other junctions using the ξ\xi from the L=400L=400 nm fit.

In conclusion, we study the evolution of critical current with respect to gate in bilayer graphene Josephson Junctions (BGJJs). Using the critical current-temperature relation expected for intermediate-to-long junctions, we extract the relevant energy scale δE\delta E and find that it has a clear gate dependence. As δE\delta E is proportional to the Fermi velocity vFv_{F} in the bilayer graphene, we are able to match the observed gate dependence to the theoretical expectation. Our observation is contrasted with monolayer graphene JJs, that do not have a gate dependent δE\delta E. This result showcases the greater tunability of BGJJs, and offers additional avenues for device characterization.

References

  • Chen et al. [2015] W. Chen, D. N. Shi, and D. Y. Xing, Long-range cooper pair splitter with high entanglement production rate, Scientific Reports 510.1038/srep07607 (2015).
  • Feigel’man et al. [2008] M. V. Feigel’man, M. A. Skvortsov, and K. S. Tikhonov, Proximity-induced superconductivity in graphene, JETP Letters 88, 747 (2008).
  • Shimazaki et al. [2015] Y. Shimazaki, M. Yamamoto, I. V. Borzenets, K. Watanabe, T. Taniguchi, and S. Tarucha, Generation and detection of pure valley current by electrically induced berry curvature in bilayer graphene, Nature Physics 11, 1032–1036 (2015).
  • Borzenets et al. [2016a] I. V. Borzenets, Y. Shimazaki, G. F. Jones, M. F. Craciun, S. Russo, M. Yamamoto, and S. Tarucha, High efficiency CVD graphene-lead (pb) cooper pair splitter, Scientific Reports 610.1038/srep23051 (2016a).
  • Kroll et al. [2018] J. G. Kroll, W. Uilhoorn, K. L. van der Enden, D. de Jong, K. Watanabe, T. Taniguchi, S. Goswami, M. C. Cassidy, and L. P. Kouwenhoven, Magnetic field compatible circuit quantum electrodynamics with graphene josephson junctions, Nature Communications 910.1038/s41467-018-07124-x (2018).
  • Park et al. [2024] G. H. Park, W. Lee, S. Park, K. Watanabe, T. Taniguchi, G. Y. Cho, and G. H. Lee, Controllable andreev bound states in bilayer graphene josephson junctions from short to long junction limits, Physical Review Letters 13210.1103/physrevlett.132.226301 (2024).
  • Dean et al. [2010] C. R. Dean, A. F. Young, I. Meric, C. Lee, L. Wang, S. Sorgenfrei, K. Watanabe, T. Taniguchi, P. Kim, K. L. Shepard, and J. Hone, Boron nitride substrates for high-quality graphene electronics, Nature Nanotechnology 5, 722 (2010).
  • Borzenets et al. [2016b] I. V. Borzenets, F. Amet, C. T. Ke, A. W. Draelos, M. T. Wei, A. Seredinski, K. Watanabe, T. Taniguchi, Y. Bomze, M. Yamamoto, S. Tarucha, and G. Finkelstein, Ballistic graphene josephson junctions from the short to the long junction regimes, Phys. Rev. Lett. 117, 237002 (2016b).
  • Kulik [1970] I. O. Kulik, Macroscopic quantization and the proximity effect in s-n-s junctions, Soviet Physics JETP 30, 944 (1970).
  • Bardeen and Johnson [1972] J. Bardeen and J. L. Johnson, Josephson current flow in pure superconducting-normal-superconducting junctions, Physical Review B 5, 72–78 (1972).
  • Svidzinsky et al. [1972] A. V. Svidzinsky, T. N. Antsygina, and E. N. Bratus, Soviet Physics JETP 3, 860 (1972).
  • Svidzinsky et al. [1973] A. V. Svidzinsky, T. N. Antsygina, and E. N. Bratus, Journal of low Temperature Physics 10, 131 (1973).
  • Wang et al. [2013] L. Wang, I. Meric, P. Y. Huang, Q. Gao, Y. Gao, H. Tran, T. Taniguchi, K. Watanabe, L. M. Campos, D. A. Muller, J. Guo, P. Kim, J. Hone, K. L. Shepard, and C. R. Dean, One-dimensional electrical contact to a two-dimensional material, Science 342, 614 (2013).
  • Borzenets et al. [2011] I. V. Borzenets, U. C. Coskun, S. J. Jones, and G. Finkelstein, Phase diffusion in graphene-based josephson junctions, Physical Review Letters 10710.1103/physrevlett.107.137005 (2011).
  • Borzenets et al. [2013] I. V. Borzenets, U. C. Coskun, H. T. Mebrahtu, Y. V. Bomze, A. I. Smirnov, and G. Finkelstein, Phonon bottleneck in graphene-based josephson junctions at millikelvin temperatures, Physical Review Letters 11110.1103/physrevlett.111.027001 (2013).
  • Tang et al. [2022] J. Tang, M. T. Wei, A. Sharma, E. G. Arnault, A. Seredinski, Y. Mehta, K. Watanabe, T. Taniguchi, F. Amet, and I. Borzenets, Overdamped phase diffusion in hbn encapsulated graphene josephson junctions, Phys. Rev. Research 4, 023203 (2022).
  • Amet et al. [2016] F. Amet, C. T. Ke, I. V. Borzenets, J. Wang, K. Watanabe, T. Taniguchi, R. S. Deacon, M. Yamamoto, Y. Bomze, S. Tarucha, and G. Finkelstein, Supercurrent in the quantum hall regime, Science 352, 966 (2016).
  • Tinkham [2004] M. Tinkham, Introduction to Superconductivity (Dover Publications, 2004).
  • Courtois et al. [2008] H. Courtois, M. Meschke, J. T. Peltonen, and J. P. Pekola, Origin of hysteresis in a proximity josephson junction, Physical Review Letters 10110.1103/physrevlett.101.067002 (2008).
  • Coskun et al. [2012] U. C. Coskun, M. Brenner, T. Hymel, V. Vakaryuk, A. Levchenko, and A. Bezryadin, Distribution of supercurrent switching in graphene under the proximity effect, Physical Review Letters 10810.1103/physrevlett.108.097003 (2012).
  • Lee et al. [2011] G. H. Lee, D. Jeong, J. H. Choi, Y. J. Doh, and H. J. Lee, Electrically tunable macroscopic quantum tunneling in a graphene-based josephson junction, Physical Review Letters 10710.1103/physrevlett.107.146605 (2011).
  • Ke et al. [2016] C. T. Ke, I. V. Borzenets, A. W. Draelos, F. Amet, Y. Bomze, G. Jones, M. Craciun, S. Russo, M. Yamamoto, S. Tarucha, and G. Finkelstein, Critical current scaling in long diffusive graphene-based josephson junctions, Nano Letters 16, 4788–4791 (2016).
  • Zhu et al. [2009] W. Zhu, V. Perebeinos, M. Freitag, and P. Avouris, Carrier scattering, mobilities, and electrostatic potential in monolayer, bilayer, and trilayer graphene, Physical Review B 8010.1103/physrevb.80.235402 (2009).
  • Fang et al. [2007] T. Fang, A. Konar, H. Xing, and D. Jena, Carrier statistics and quantum capacitance of graphene sheets and ribbons, Applied Physics Letters 9110.1063/1.2776887 (2007).
  • Fates et al. [2019] R. Fates, H. Bouridah, and J. P. Raskin, Probing carrier concentration in gated single, bi- and tri-layer cvd graphene using raman spectroscopy, Carbon 149, 390–399 (2019).
  • Zou et al. [2011] K. Zou, X. Hong, and J. Zhu, Effective mass of electrons and holes in bilayer graphene: Electron-hole asymmetry and electron-electron interaction, Physical Review B 8410.1103/physrevb.84.085408 (2011).