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 Sharma Department of Physics and Astronomy, Texas A&\&M University, College Station, 77843, Texas, USA    Chun-Chia Chen Department of Physics, Duke University, Durham, 27701, North Carolina, USA    Jordan McCourt Department of Physics, Duke University, Durham, 27701, North Carolina, USA    Mingi Kim Department of Physics and Astronomy, Purdue University, West Lafayette, 47907, Indiana, USA    Kenji Watanabe Advanced Materials Laboratory, NIMS, Tsukuba, 305-0044, Japan    Takashi Taniguchi Advanced Materials Laboratory, NIMS, Tsukuba, 305-0044, Japan    Leonid Rokhinson Department of Physics and Astronomy, Purdue University, West Lafayette, 47907, Indiana, USA    Gleb Finkelstein Department of Physics, Duke University, Durham, 27701, North Carolina, USA    Ivan Borzenets Department of Physics and Astronomy, Texas A&\&M University, College Station, 77843, Texas, USA borzenets@tamu.edu
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.

keywords:
Graphene, Bilayer Graphene, Josephson Junctions, Fermi Velocity, Andreev Levels.

Ballistic graphene Josephson junctions (GJJs) have been widely utilized as a platform to study novel quantum physics phenomena1, 2 and devices3, including: entangled pair generation4, 5, topological states arising from the mixing of superconductivity and quantum Hall states6, as well as photon sensing via bolometry/calorimetry7. Superconductor-normal metal-superconductor Josephson junction (SNSJJ) hosts Andreev bound states (ABS), which carry supercurrents across the normal region of the JJ; in order to enter the ballistic regime, a disorder-free weak link and high transparency at the SN interface are necessary. Hexagonal Boron-Nitride (hBN) encapsulated graphene as a weak link enables highly transparent contacts at the interface whilst keeping graphene clean throughout the fabrication process 8. 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 1.

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 the ABS level spacing 2, 9, 10, 11, 12, 13. Note that in the intermediate regime (Lξ0L\approx\xi_{0}) δE\delta E is found to be suppressed 5. 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) 2, 13. 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 the 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 14. 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 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. 15, 16, 5, 2, 17, 6.

Refer to caption
Figure 1: (a) Differential resistance (dV/dIdV/dI) 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 up (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=15V_{G}=15 V, T=1.37T=1.37 K, showing device’s dV/dIdV/dI versus bias current. Blue line corresponds to IBiasI_{Bias} swept up, with red line swept down (positive to negative).

Figure. 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 the 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 the switching current. Figure. 1(a) shows that the junction can sustain a larger region of critical current as we modulate the carrier density to higher values via VGV_{G}. Fig. 1(b) displays line traces 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, 15, or can also be attributed to self-heating 16, 19, 17. The measured switching current ISI_{S} is slightly suppressed compared to the junction’s “true” critical current ICI_{C}. However, previous measurements on the statistical distribution of ISI_{S} in similar graphene devices found that ISI_{S} is suppressed from ICI_{C} by no more than 10%10\% for critical currents up to a few μA\mu A 2, 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 (Figure. 2c) We also extract the conductance of the junction in the normal regime (IBiasICI_{Bias}\gg I_{C}). Figure. 2(b) shows this conductance (GG) for the 400400 nm junction device. Due to the significant contact resistance (RCR_{C}) of the device, the measured conductance GG is uniformly suppressed compared to the ballistic limit expectation. However, when accounting for RCR_{C} within the fit, we find that the conductance GG scales as the square-root (as opposed to linearly) of VGV_{G} (blue curve of Figure 2(b)). This is consistent with ballistic transport23, 2. To further demonstrate the ballistic nature of the device, we present normal resistances (RNR_{N}) of junctions of length 500500 nm, 600600 nm, and 700700 nm with the fitted, constant contact resistance RCR_{C} subtracted (Figure. 2(b) inset). The inset plot shows that the values of RNRCR_{N}-R_{C} are independent of the junction length, demonstrating 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 blue curve corresponds to the fit for ballistic devices, with an addition of a contact resistance. 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. the temperature trends in Figure. 2(c). Here, the y-axis is plotted in logarithmic scale. From the slope of the 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 Figure. 3a). Unlike for the case of monolayer graphene, a clear dependence on VGV_{G} is seen (The observed trend further supports the view that our devices operate in the long ballistic regime. Diffusive Josephson junctions are governed by the Thouless energy ETh1/[(RNRC)VGVD]E_{Th}\propto 1/[(R_{N}-R_{C})\sqrt{V_{G}-V_{D}}] 22, 24 which does not match the trend with respect to VGV_{G} seen in Figure. 3(a) ). The energy δE\delta E scales linearly with the Fermi velocity vFv_{F} (Figure. 3(b) ). 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 the fit discussed below, we use ξ\xi’s dependence in vFv_{F}.

We now compare the experimentally obtained δE\delta E (and vFv_{F}) to the theoretical expectation. With the dispersion relation for bilayer graphene written as : =12γ1(1+22k2/γ12m1)\mathcal{E}=\frac{1}{2}\gamma_{1}(\sqrt{1+2\hbar^{2}k^{2}/\gamma_{1}^{2}m^{*}}-1), we get the expression for the Fermi velocity : vF=2Fγ1(F+γ1)(2F+γ1)2mv_{F}=\sqrt{\frac{2\mathcal{E}_{F}\gamma_{1}(\mathcal{E}_{F}+\gamma_{1})}{(2\mathcal{E}_{F}+\gamma_{1})^{2}m^{*}}} 25, 26, 27. Here, γ1=0.39eV\gamma_{1}=0.39~{}eV a parameter describing the interlayer coupling 25, kk is the momentum wavevector, mm^{*} is the effective mass of electrons. Moreover, 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^{*}}. 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}}, where ee is the electron charge. The gate oxide capacitance per unit area is Cox=ϵ0ϵrdC_{ox}=\frac{\epsilon_{0}\epsilon_{r}}{d}, where ϵ0\epsilon_{0} is the vacuum permittivity, ϵr\epsilon_{r} is the relative permittivity of the oxide, and dd is the thickness of the oxide layer. 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=2πeϵ0ϵrγ1(VGVD)(2de2γ1m+πϵ0ϵr2(e(VGVD)+γ1))m(2de2γ1m+πϵ0ϵr2(2e(VGVD)+γ1))2\displaystyle v_{F}=\hbar\sqrt{\frac{2\pi e\epsilon_{0}\epsilon_{r}\gamma_{1}(V_{G}-V_{D})(2de^{2}\gamma_{1}m^{*}+\pi\epsilon_{0}\epsilon_{r}\hbar^{2}(e(V_{G}-V_{D})+\gamma_{1}))}{m^{*}(2de^{2}\gamma_{1}m^{*}+\pi\epsilon_{0}\epsilon_{r}\hbar^{2}(2e(V_{G}-V_{D})+\gamma_{1}))^{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} 28, where mem_{e} is the electron rest mass. 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}. We also note that ξ\xi has a dependence on vFv_{F} and the superconducting gap Δ\Delta: ξ=vF/2Δ\xi=\hbar v_{F}/2\Delta13. To fit δE\delta E, the model is set as: δE(VG)=(m,Δ,VD,d)\delta E(V_{G})=\mathcal{F}(m^{*},\Delta,V_{D},d) where m,Δ,VD,dm^{*},\Delta,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}.)

Parameter Fitted Value Expected Value
Δ\Delta 0.990.99 meV 0.81.20.8\sim 1.2 meV
dd 323323 nm 280330280\sim 330 nm
mm^{*} 0.028me0.028~{}m_{e} 0.020.06me0.02\sim 0.06~{}m_{e}
VDV_{D} 2.042.04 V +2\approx+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 obtained from the resistance map. The expectations for superconducting gap Δ\Delta and the effective mass mim_{i} are obtained from previous works28, 5.

The resulting fits of the data from the 400nm400~{}nm junction for δE\delta E and vFv_{F} are plotted as solid lines in Figure. 3(a) and Figure. 3(b) respectively. Moreover, taking the fitted parameters from Table 1, we calculate the Fermi velocity vFv_{F} for the available data points of all other junctions on the same substrate. As seen from Figure. 3(b), the calculated vFv_{F} of all devices is in good agreement with the fit obtained from the 400400 nm junction (This is as expected for devices on the same substrate; as long as they have consistent parasitic doping and superconductor-graphene contact interface). The fitted parameters are summarized in Table 1. All fall within the range of expected values, with Δ\Delta being consistent with previously measured values for graphene/MoRe junctions2. Furthermore, using the values obtained from the model, we find that vFv_{F} saturates to the value of 1.11061.1*10^{6} m/s as VGV_{G} tends to infinity.

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 parameters 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 parameters obtained from the L=400L=400 nm fit.

In conclusion, we study the evolution of the critical current with respect to the 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 bilayer graphene, we are able to match the observed gate dependence to the theoretical expectation. Our observation is contrasted with monolayer graphene JJs, which do not have a gate-dependent δE\delta E. This result showcases the greater tunability of BGJJs, and offers additional avenues for device characterization. Although not observed here, it should be possible to engineer Josephson junctions that transition from the short to the intermediate/long ballistic regimes in-situ via gate voltage. The ability to tune ABS level spacing could have applications in self-calibrating sensors, or for matching resonance conditions in multi-terminal superconducting devices.

L.R. acknowledges support from NSF (DMR-2005092 award) for contact depositon. G.F. acknowledges support from Duke University. I.V.B. and A.S. acknowledge the support from Texas A&\&M University.

References

  • Park et al. 2024 Park, G. H.; Lee, W.; Park, S.; Watanabe, K.; Taniguchi, T.; Cho, G. Y.; Lee, G. H. Controllable Andreev Bound States in Bilayer Graphene Josephson Junctions from Short to Long Junction Limits. Physical Review Letters 2024, 132
  • Borzenets et al. 2016 Borzenets, I. V.; Amet, F.; Ke, C. T.; Draelos, A. W.; Wei, M. T.; Seredinski, A.; Watanabe, K.; Taniguchi, T.; Bomze, Y.; Yamamoto, M.; Tarucha, S.; Finkelstein, G. Ballistic Graphene Josephson Junctions from the Short to the Long Junction Regimes. Phys. Rev. Lett. 2016, 117, 237002
  • Kroll et al. 2018 Kroll, J. G.; Uilhoorn, W.; van der Enden, K. L.; de Jong, D.; Watanabe, K.; Taniguchi, T.; Goswami, S.; Cassidy, M. C.; Kouwenhoven, L. P. Magnetic field compatible circuit quantum electrodynamics with graphene Josephson junctions. Nature Communications 2018, 9
  • Chen et al. 2015 Chen, W.; Shi, D. N.; Xing, D. Y. Long-range Cooper pair splitter with high entanglement production rate. Scientific Reports 2015, 5
  • Borzenets et al. 2016 Borzenets, I. V.; Shimazaki, Y.; Jones, G. F.; Craciun, M. F.; Russo, S.; Yamamoto, M.; Tarucha, S. High Efficiency CVD Graphene-lead (Pb) Cooper Pair Splitter. Scientific Reports 2016, 6
  • Amet et al. 2016 Amet, F.; Ke, C. T.; Borzenets, I. V.; Wang, J.; Watanabe, K.; Taniguchi, T.; Deacon, R. S.; Yamamoto, M.; Bomze, Y.; Tarucha, S.; Finkelstein, G. Supercurrent in the quantum Hall regime. Science 2016, 352, 966–969
  • Lee et al. 2020 Lee, G.-H.; Efetov, D. K.; Jung, W.; Ranzani, L.; Walsh, E. D.; Ohki, T. A.; Taniguchi, T.; Watanabe, K.; Kim, P.; Englund, D.; Fong, K. C. Graphene-based Josephson junction microwave bolometer. Nature 2020, 586, 42–46
  • Dean et al. 2010 Dean, C. R.; Young, A. F.; Meric, I.; Lee, C.; Wang, L.; Sorgenfrei, S.; Watanabe, K.; Taniguchi, T.; Kim, P.; Shepard, K. L.; Hone, J. Boron nitride substrates for high-quality graphene electronics. Nature Nanotechnology 2010, 5, 722–726
  • Kulik 1970 Kulik, I. O. Macroscopic Quantization and the Proximity Effect in S-N-S Junctions. Soviet Physics JETP 1970, 30, 944
  • Bardeen and Johnson 1972 Bardeen, J.; Johnson, J. L. Josephson Current Flow in Pure Superconducting-Normal-Superconducting Junctions. Physical Review B 1972, 5, 72–78
  • Svidzinsky et al. 1972 Svidzinsky, A. V.; Antsygina, T. N.; Bratus, E. N. Superconducting current in wide sns junctions. Soviet Physics JETP 1972, 34, 860
  • Svidzinsky et al. 1973 Svidzinsky, A. V.; Antsygina, T. N.; Bratus, E. N. Concerning the theory of the Josephson effect in pureSNS junctions. Journal of low Temperature Physics 1973, 10, 131–136
  • Bagwell 1992 Bagwell, P. F. Suppression of the Josephson current through a narrow, mesoscopic, semiconductor channel by a single impurity. Phys. Rev. B 1992, 46, 12573–12586
  • Wang et al. 2013 Wang, L.; Meric, I.; Huang, P. Y.; Gao, Q.; Gao, Y.; Tran, H.; Taniguchi, T.; Watanabe, K.; Campos, L. M.; Muller, D. A.; Guo, J.; Kim, P.; Hone, J.; Shepard, K. L.; Dean, C. R. One-Dimensional Electrical Contact to a Two-Dimensional Material. Science 2013, 342, 614–617
  • Borzenets et al. 2011 Borzenets, I. V.; Coskun, U. C.; Jones, S. J.; Finkelstein, G. Phase Diffusion in Graphene-Based Josephson Junctions. Physical Review Letters 2011, 107
  • Borzenets et al. 2013 Borzenets, I. V.; Coskun, U. C.; Mebrahtu, H. T.; Bomze, Y. V.; Smirnov, A. I.; Finkelstein, G. Phonon Bottleneck in Graphene-Based Josephson Junctions at Millikelvin Temperatures. Physical Review Letters 2013, 111
  • Tang et al. 2022 Tang, J.; Wei, M. T.; Sharma, A.; Arnault, E. G.; Seredinski, A.; Mehta, Y.; Watanabe, K.; Taniguchi, T.; Amet, F.; Borzenets, I. Overdamped phase diffusion in hBN encapsulated graphene Josephson junctions. Phys. Rev. Research 2022, 4, 023203
  • Tinkham 2004 Tinkham, M. Introduction to Superconductivity; Dover Publications, 2004
  • Courtois et al. 2008 Courtois, H.; Meschke, M.; Peltonen, J. T.; Pekola, J. P. Origin of Hysteresis in a Proximity Josephson Junction. Physical Review Letters 2008, 101
  • Coskun et al. 2012 Coskun, U. C.; Brenner, M.; Hymel, T.; Vakaryuk, V.; Levchenko, A.; Bezryadin, A. Distribution of Supercurrent Switching in Graphene under the Proximity Effect. Physical Review Letters 2012, 108
  • Lee et al. 2011 Lee, G. H.; Jeong, D.; Choi, J. H.; Doh, Y. J.; Lee, H. J. Electrically Tunable Macroscopic Quantum Tunneling in a Graphene-Based Josephson Junction. Physical Review Letters 2011, 107
  • Ke et al. 2016 Ke, C. T.; Borzenets, I. V.; Draelos, A. W.; Amet, F.; Bomze, Y.; Jones, G.; Craciun, M.; Russo, S.; Yamamoto, M.; Tarucha, S.; Finkelstein, G. Critical Current Scaling in Long Diffusive Graphene-Based Josephson Junctions. Nano Letters 2016, 16, 4788–4791
  • Castro Neto et al. 2009 Castro Neto, A. H.; Guinea, F.; Peres, N. M. R.; Novoselov, K. S.; Geim, A. K. The electronic properties of graphene. Rev. Mod. Phys. 2009, 81, 109–162
  • Dubos et al. 2001 Dubos, P.; Courtois, H.; Pannetier, B.; Wilhelm, F. K.; Zaikin, A. D.; Schön, G. Josephson critical current in a long mesoscopic S-N-S junction. Phys. Rev. B 2001, 63, 064502
  • McCann and Koshino 2013 McCann, E.; Koshino, M. The electronic properties of bilayer graphene. Reports on Progress in Physics 2013, 76, 056503
  • Fang et al. 2007 Fang, T.; Konar, A.; Xing, H.; Jena, D. Carrier statistics and quantum capacitance of graphene sheets and ribbons. Applied Physics Letters 2007, 91
  • Fates et al. 2019 Fates, R.; Bouridah, H.; Raskin, J. P. Probing carrier concentration in gated single, bi- and tri-layer CVD graphene using Raman spectroscopy. Carbon 2019, 149, 390–399
  • Zou et al. 2011 Zou, K.; Hong, X.; Zhu, J. Effective mass of electrons and holes in bilayer graphene: Electron-hole asymmetry and electron-electron interaction. Physical Review B 2011, 84