Introduction

A graphene layer1,2, a one-atom-thick planar sheet of carbon atoms, attached to two closely spaced superconducting electrodes forms a Josephson junction (JJ)3, in which the supercurrent flows through graphene via coherent electron–hole Andreev pairs4. For earlier discussions on the proximity-induced superconductivity, please refer to ref. 5 and references therein. As a JJ can be described in terms of motion of superconducting phase particles in a potential well6, macroscopic quantum tunnelling (MQT)7 and energy level quantization8, which have already been observed in a JJ of graphene (JJG)9, provide essential requisites for superconducting quantum bits6 in graphene. In a JJG, however, complete switching off of the junction supercurrent by the gate voltages has not yet been achieved. This incomplete gate control of the supercurrent in a JJG4,9,10,11 poses an obstacle in the way of quantum-state engineering.

From a fundamental viewpoint, a JJG is rich in physics, providing the opportunity to study the Josephson effect combined with relativistic Dirac nature, as low-energy quasiparticles in graphene are described by massless Dirac fermions12. A recent theoretical work13 showed that a finite supercurrent flows through a JJG in the ballistic regime in the absence of an external magnetic field; this occurs even when the Fermi energy is tuned to the charge neutrality point (CNP; that is, the centre of a Dirac cone) with nominal zero carrier density. In this case, the product of the Josephson critical current Ic and the normal-state resistance Rn is within an order of Δ/e near the CNP, satisfying the conventional Ambegaokar–Baratoff relation14. It has also been shown15 that specular Andreev reflections occur at the graphene-superconductor interface near the CNP, due to the Dirac-cone energy band structure. However, the effects of intrinsic scattering sources in graphene on the supercurrent of a JJG have not been studied in detail16. The role of the valley degrees of freedom on scattering of supercurrent flow in graphene has not been explored. In addition, it would be interesting to investigate a JJG with an electrostatic potential barrier in its normal region; the presence of a potential step is known to induce the negative refraction and Klein tunnelling of Dirac fermions17,18, with the latter being analogous to the behaviour of relativistic particles. Understanding such scattering effects would be highly useful for graphene-based supercurrent device development.

Here we report complete switching off of the supercurrent in a p–n junction of graphene by dual operation of the top and bottom gates. This supercurrent OFF state does not necessarily require an infinite junction resistance by reduced carrier density in the normal state, which is in marked contrast to other hybrid JJs of nanomaterials19,20,21,22. Gate-driven crossover of the escape processes of superconducting phase particles, from MQT to the classical regime, is also demonstrated as a novel route for manipulating the macroscopic quantum states. We suggest that the spatially random pseudomagnetic field generated by the ripples (spatially non-uniform strain) in graphene is mainly responsible for the observed supercurrent OFF state. The ripples act as intravalley scattering sources. The random pseudomagnetic field can cause dephasing of coherent Andreev electron–hole pairs in the normal region of the junction, possibly leading to full suppression of the junction supercurrent near the CNP and in the bipolar junction region of the gate voltages, where scattering by ripples becomes most significant. The evidence of ripples in our junction was supported by weak antilocalization (WAL) analysis, which showed that the scattering rate by ripples was substantially enhanced near the CNP. Our study provides a key step towards developing superconducting quantum information devices23,24 based on graphene technology12.

Results

Gate-controlled complete suppression of Ic in a bipolar JJG

A schematic view of the bipolar JJG is shown in Fig. 1a, where the JJG is sandwiched between the local top-gate (Vtg) and global back-gate (Vbg) electrodes (see Methods for device fabrication). In the superconducting state at the base temperature (T~50 mK), the current−voltage (IV) characteristics in Fig. 1b, with Vtg=0 V, clearly show supercurrent branches; here the critical (or switching) current (Ic) reaches 500 nA in the highly p-doped region. Complete Ic suppression is obtained at the CNP, corresponding to Vbg=12.5 V. Varying Vtg and Vbg in a stepwise manner, we obtain the colour plot of Ic in Fig. 1c, which resembles a quadrant map of the normal-state conductance G (=Rn−1) showing four distinct regions of junction states: p–p′, p–n, n–p and n–n′. These states depend on the respective carrier types under the influence of Vbg and Vtg (see Supplementary Figs S1 and S2). It should be emphasized that Ic vanishes completely at the CNP and inner parts of the bipolar region as well. The slice plots in Fig. 1d show that the supercurrent OFF state is obtained even with relatively low values of Rn (~0.3 kΩ); this corresponds to a sheet conductance of Gsq~13 e2/h, where e is the elementary charge and h is Planck’s constant.

Figure 1: Gate-controlled complete suppression of supercurrent in JA.
figure 1

(a) Schematic configuration of the JJG p–n barrier (JA) established by the dual-gate operation of Vbg and Vtg. (b) Vbg dependence of the junction critical current (Ic) and retrapping current (Ir) at the base temperature and B=0, with Vtg fixed at 0 V. Highly doped states exhibit a clear Ic with large hysteresis. (c) Colour-coded plot of Ic taken at B=0. The quadrant map of Ic was obtained from the I–V curves by varying Vbg and Vtg by steps of 5 and 1 V, respectively. The overall CNP forms at Vbg=12.6 V and Vtg=0 V. (d) Vbg dependence of Ic (symbol) and the normal-state resistance Rn (line) for Vtg=0, −2, −4, −6 and −8 V.

When a microwave field with frequency fmw is applied, the I–V curve exhibits quantized voltage plateaus, or Shapiro steps (SSs)14, occurring at Vn=nhfmw/(2e) with integer n, due to phase locking between the JJG and microwave (Supplementary Fig. S3). Typical standard Josephson effects, including SSs and periodic modulation of the Ic with a perpendicular magnetic field (that is, Fraunhofer pattern14), are obtained in the overdoped unipolar region (Supplementary Fig. S3). When Vbg is varied from the overdoped n–n′ region to the CNP with Vtg=0 V, progressive evolution of the dI/dV peaks, equivalent to the SSs, is evident as shown in Fig. 2a for a fixed microwave power (P). Figure 2a clearly shows that the SSs are absent near the CNP, in contrast to the previous report of significantly reduced but still finite SSs in the plain JJG11. The vanishing SSs are attributed to complete Ic suppression at the CNP in this experiment. The SSs also vanish in the bipolar n–p and p–n region (Supplementary Fig. S4).

Figure 2: Suppression of SSs and MARs near CNP of graphene.
figure 2

(a) Progressive evolution of the differential conductance of the junction JA, under the irradiation of a microwave of frequency fmw, with varying Vbg for Vtg=0 V, with peaks occurring at the bias voltage V for integral multiples of V1 (=hfmw/2e=10.3 μV). The dI/dV peaks, equivalent to SSs, are completely suppressed as the CNP (Vbg=12.5 V for Vtg=0 V) is approached. (b) Progressive change of dI/dV versus V curve of the junction JB as the CNP is approached. The dI/dV value is normalized by the normal-state resistance Rn for each Vbg. The conductance enhancement at the subharmonic voltages, reflecting the MARs, is suppressed and the supercurrent at zero-bias voltage converts into a dip of dI/dV near the CNP.

Additional experimental evidence for complete Ic suppression is shown in Fig. 2b, displaying the dI/dV versus V curve with respect to Vbg. With Vbg=−8.4 V in the p–p′ region, we observe multiple dI/dV peaks occurring at V=2Δ/ne, where 2Δ=210 μeV is the superconducting energy gap of Al. These subgap conductance peaks are due to the multiple Andreev reflection (MAR) effect4,10,25, where the electron-like quasiparticle in a normal metal (N), incident to the superconductor (S), is retro-reflected as a phase-conjugated hole-like quasiparticle at the N–S interface, leaving Cooper pairs in S and vice versa. When Vbg approaches the CNP (Vbg=2.4 V for device JB), the infinitely high dI/dV peak, corresponding to a finite supercurrent, converts into a zero-bias dI/dV dip; this is consistent with the observations of the supercurrent OFF state formed near the CNP. We note that the MAR is also suppressed, resulting in a substantial decrease of the dI/dV peak heights near the CNP, in contrast with a previous report10 of MAR peaks that were insensitive to Vbg in a plain JJG. Similar gradual suppression of MAR peaks is seen as Vtg approaches the bipolar junction region for fixed Vbg=5 V (Supplementary Fig. S4e). Our observation suggests that the phase-coherent transport of the Andreev-reflected quasiparticle pairs is hindered at the Ic suppression region.

Theoretically, it is expected13,26 that a JJG can carry a non-zero supercurrent even at the CNP, where the charge density is tuned to zero. In this case, the IcRn product is nearly constant as ~2Δ/e with Vbg in a ballistic short junction13, which is in essence the Ambegaokar–Baratoff relation14; in a diffusive long junction, it is given by αETh/e (refs 11, 27), where ETh is the Thouless energy and α is a coefficient depending on ETh/Δ. Previous experiments4,9,10,11 of JJG in a diffusive junction regime have shown that the IcRn product at the CNP is almost half the maximum value obtained in the overdoped region. As our bipolar JJG exhibits the maximum IcRn=40 μV in the overdoped region, the minimum Ic is expected to be ~44 nA from Rn=450 Ω at the CNP. This discrepancy, between the expected value of Ic and the observed completely suppressed value, is not explained by thermal fluctuations11. Thermal suppression of Ic by 2 nA is estimated for T=50 mK, by comparing the Josephson coupling energy (EJ=ħIc/2e; ħ=h/2π) and thermal energy14 (kBT). Environmental noise or low transparency at the junction interface is ruled out by considering a maximum Ic~500 nA in the overdoped region, sufficiently larger than previously reported values4, and clear existence of the MAR peaks. A possible mechanism for the supercurrent OFF state is discussed below.

Temperature- and gate-dependent current-switching behaviour

Complete gate control of the supercurrent in the bipolar JJG would significantly facilitate the gate-tuneable operation of macroscopic quantum states in the JJ, where the confinement energy of superconducting phase particles is determined by EJ. As the MQT of the phase particle, a direct evidence of macroscopic quantum state existing in the JJ, is determined by competition between EJ and kBT, the gate tuneability of Ic in the bipolar JJG would allow the escaping process of the phase particle to be controlled using the gate voltages only. The switching distribution of Ic is used to reveal the escaping process in the JJ (see Fig. 3a). Temperature-dependent Ic distributions of the bipolar JJG with fixed Vbg and Vtg are fitted well to the theoretical models of phase diffusion (PD), thermal activation (TA) and MQT in sequence (see Fig. 3b and Supplementary Fig. S5), which are quite similar to the previous observations using the plain homogeneous JJG9. Surprisingly, the progressive change of the Ic distribution is also obtained with varying Vbg for fixed temperature, as shown in Fig. 3c. The theoretical models of PD, TA and MQT fit well the experimental data, revealing that sequential crossover of the escaping processes is tuned by Vbg instead of temperature. To the best of our knowledge, this work is the first to demonstrate full control of the phase particle’s escaping process using the gate voltage only. Similar behaviour of the gate-tuned Ic distribution curve is also observed with different Vtg (see Supplementary Fig. S6). This achievement is attributed to the complete gate tuneability of the Ic to render a significant variation of the ratio between kBT and EJ at a fixed temperature.

Figure 3: T - and Vbg-driven current switching of bipolar JJG ( JA).
figure 3

(a) Schematic of washboard potential U(φ) normalized by the Josephson coupling constant EJ (φ is the phase difference across the JJ) and dynamics of a phase particle escaping from the local potential minimum (PD, phase diffusion; TA, thermal activation; and MQT, macroscopic quantum tunnelling). (b) T dependence of the switching current distribution P(Ic) of the JJG, JA, for fixed Vbg (=−30 V) and Vtg (=0.5 V). The solid lines are the best fits to the PD (T=245~340 mK), TA (T=125~215 mK) and MQT (T=60, 80 mK) model. (c) Vbg dependence of the switching current distribution for fixed Vtg (0.5 V) at T=60 mK. The solid lines are the best fits for PD (Vbg=−6~−12 V), TA (Vbg=−15~−24 V) and MQT (Vbg=−27~−33 V) models, respectively. (d) Normalized s.d. versus Vbg plot for different Vtg at T=60 mK, extracted from the distributions in Fig. 3c and Supplementary Fig. S6. Lines are guides for the eye. Upward and downward arrows represent the crossover back-gate voltages between the MQT and TA regions (Vbg,q) and between the TA and PD regions (Vbg,p), respectively, for the given fixed values of Vtg. (e) The quadrant conductance map overlaid with the MQT, TA and PD regions of the switching-current distribution in the p–p' state of graphene of JA. Cross symbols represent the crossover gate voltages determined from Fig. 3d. Dashed lines are guides for the eye. The filled circular symbols denote the points where the T dependence of the s.d. given in Fig. 7. Cyan (red, yellow)-filled circles represent the MQT (TA and PD) regime at the base temperature, which is determined by the T dependence of normalized s.d.

Three distinctive regimes of the switching process can be identified using the standard deviation (s.d.) of the Ic distribution normalized by the mean value Ic, as shown in Fig. 3d, depending on Vbg and Vtg. The switching rate in the TA regime, ΓTA~exp(−Ic), is more sensitive to the Ic than that of the MQT where ΓMQT~exp(−Ic0.5). Thus, we can define the crossover gate voltage of Vbg,q, between the MQT and TA regions, as the value of Vbg at which the slope of the normalized s.d. changes abruptly9. A second crossover gate voltage (Vbg,p) between the TA and PD regions is defined as Vbg when the slope changes its sign. An overlay of the Ic switching mechanism phase diagram on the quadrant map of G is shown in Fig. 3e, where the crossovers between the switching processes are fully gate driven in contrast to the plain JJG9. We note that the phase boundary lines are shifted from the boundary of the quandrant map, forming the contour line of Ic for each crossover. This also indicates that the competition between EJ and kBT is responsible for the crossovers.

Supercurrent suppression mechanism

A typical turning-off process of the supercurrent in other nanohybrid JJs19,20,21,22 utilizes an infinite Rn induced by carrier depletion in the nanostructures. Ic suppression in the bipolar JJG, however, is not accompanied by a divergent Rn, alluding the unusual turning-off mechanism for the supercurrent. This could be caused by selective transmission of carriers in Klein tunnelling17,18, which hinders the phase-coherent propagation of the Andreev-reflected quasiparticle pairs at the p–n boundary, or by specular-like reflection at the interface between charge puddles16. The latter effect by charge puddles changes the dynamical phase of the charge carriers in the normal graphene region; hence, it can give rise to dephasing of the Andreev-reflected quasiparticle pairs. However, the change in the dynamical phase is not expected to be sufficient to cause dephasing in the short junction limit, where the superconducting coherence length, ξs (=(ħD/Δ)1/2=260 nm), is comparable to the junction spacing, L (300 nm), with a diffusion constant of D=110 cm2 s−1 in this study. Moreover, the former effect by Klein tunnelling alone is not sufficient to explain complete Ic suppression at the CNP (see Supplementary Figs S7 and S8).

Here we propose a new Ic suppression mechanism, based on ripples and other scattering sources of Andreev quasiparticle pairs, which may exist in our JJG. Our mechanism is summarized as follows. An electron and its Andreev-reflected hole acquire a different amount of phase shift in the normal n–p region, due to spatially random pseudomagnetic fields induced by ripples28,29. The resulting random phase shifts can cause dephasing of the Andreev pairs, and hence a large reduction in the phase-coherent effect of the Josephson current near the CNP, where scattering by ripples becomes significant. Scattering sources such as puddles30, specular Andreev reflection15 and the interface between n-type and p-type regions17,18 further enhance the reduction, possibly leading to full Ic suppression. This suppression in Ic occurs, because an electron and its Andreev-reflected hole propagate along different paths that become further deviated from each other due to the scattering sources, and acquire ripple-induced random phase shifts with respect to each other.

Ripples, spatially non-uniform strains, are intrinsic structures of graphene31 and affect the electron properties by inducing pseudomagnetic fields28,29. In contrast to real magnetic fields, the pseudomagnetic fields are opposite in their direction with respect to the other between the K and K′ valleys. Thus, although these pseudomagnetic fields break the effective time reversal symmetry defined within a single valley, the actual time reversal symmetry remains intact32. A ripple typically has a diameter of d~10 nm and a height hR~0.2 nm33,34, generating a pseudomagnetic flux, Φ, in the region of the ripple, defined by28

where Φ0=h/2e is the flux quantum and a=0.14 nm is the interatomic distance in graphene. An Andreev quasiparticle propagating through a ripple acquires a phase shift of 2πΝΦ due to the pseudomagnetic flux, where ΝΦ=Φ/Φ0.

Our JJG (whose junction length and width were ~300 and ~2,000 nm, respectively) could accommodate ~6,000 ripples in its normal region. The ripples were assumed to be randomly distributed spatially in the normal n–p region. Hence, the total phase shift acquired by an Andreev quasiparticle, resulting from multiple scattering events by the randomly distributed ripples over the junction length, L, follows a Gaussian distribution, with variance σ2 given by

where vF is the Fermi velocity, ~vF/(|kF(F′)|d2) is the scattering rate35 by ripples in the n(p) region and kF(F′) is the electron wave vector in the n(p) region. Here, L/(|kF|d2) represents the number of ripples in the area of L/|kF|. If the ripple-induced phase shift of an electron propagating through the normal region is uncorrelated with that of its Andreev partner (the Andreev-reflected hole), then the random distribution of the total phase shift results in severe reduction of the Josephson supercurrent. It is described by the current-phase relation (see Supplementary Note 7 for the detailed derivation)

where φ is the phase difference between the two superconducting JJG electrodes. We notice that the Josephson current is significantly suppressed near the CNP, because increases dramatically as kF(F′) vanishes near the CNP. In contrast to the Josephson supercurrent, the quasiparticle propagation in the normal state is not hindered by the ripple-induced dephasing effect, retaining a finite Rn even at the CNP, which is consistent with our observations.

As described above, equation (3) is derived under the condition of uncorrelated ripple-induced phase shifts between Andreev partners (an electron and its Andreev-reflected hole). This condition is satisfied, because the Andreev partners usually propagate along different paths from each other in the normal regions, acquiring phase shifts that are uncorrelated due to the random spatial distribution of the ripples. Although extremely rare, it is possible that Andreev partners have exactly the same paths (for example, the case with zero excitation energy). This case contributes to Ic, as the ripple-induced phase shifts of the Andreev partners cancel each other, because the Andreev partners are in different valleys (for example, an electron in the K-valley and a hole in the K′-valley) and are affected by the pseudomagnetic fields with directions opposite to each other32. In this case, other scattering sources such as puddles30, specular Andreev reflection15 and the interface between n and p17,18 further reduce and may fully suppress Ic, because they make the Andreev partners propagate along different paths that deviate from each other36. Thus, the ripple-induced phase shifts become uncorrelated between Andreev partners. For example, near the CNP, a puddle results in the deviation of the propagating paths between the Andreev partners. In this case, an electron (hole) undergoes negative refraction at the interface between the p-type (n-type) region and a zero-doping region formed near the puddle, whereas a hole (an electron) shows normal refraction at the same interface.

Our calculation results based on equation (3), along with the experimental parameters of this work, are in good agreement with the experimental data, as shown in Fig. 4a. The Ic suppression observed in the bipolar JJG was mainly caused by the ripple-induced dephasing effect of the Andreev-reflected quasiparticle pairs. The dephasing effect was highly enhanced at the CNP, as expected in the ripple scattering model, whereas it was negligible in the unipolar junction region (Fig. 1c). The dephasing effect will also be significant in the bipolar junction region. In the bipolar junction region, a localized CNP region forms spatially at the p–n interface or over one-half of the JJG, as the Fermi level aligns with either side of the bipolar potential profile to make the p–0 (or n–0) state.

Figure 4: Dephasing effect by ripples in graphene.
figure 4

(a) The junction critical current Ic (solid lines) is calculated with varying Vbg with respect to VCNP (that is, ΔVbg=VbgVCNP) at fixed values of Vtg near the CNP, taking into account the spatially random pseudomagnetic field induced by ripples. The calculation result is in qualitative agreement with the experimental observation from JA (symbols). (b) Comparison of the scattering rate τ*n−1 by ripples (filled circles) as a function of ΔVbg (obtained from a weak-(anti)localization fit for JA at Vtg=0.5 V; see Supplementary Fig. S9) and the theoretical prediction of τ*n−1~vF/(kFd2) (solid line).

WAL analysis

To find the evidence of ripples in our JJG, we measured the magnetoconductance of the normal state of the JJG and analysed the WAL effects. As the pseudomagnetic field of the ripples breaks the effective time reversal symmetry within a single valley37, the WAL effect38 in graphene can be suppressed by the intravalley scattering due to ripples. It is known that other intravalley scattering sources, such as puddles, may not be detected by WAL analysis, because they do not cause WAL suppression; in addition, the trigonal warping effect38 is typically very small (scattering rate ≤0.03 ps−1) near the CNP. Therefore, the suppression of WAL at the CNP can be interpreted as an evidence of ripples.

The measured magnetoconductance of the normal state of our JJG exhibited the suppression of WAL near the CNP. Magnetoconductance data for a top-gate voltage of Vtg=0.5 V and temperature of 4.8 K are plotted in Supplementary Fig. S9. By analysing the data based on the WAL fit38 for different Vbg, we obtained scattering rates for the normal state, which are summarized in Supplementary Table S1; the details of the analysis are given in the Methods section. The intravalley scattering rate in the n(p) region, , plotted in Fig. 4b, is well described by

Here we use the empirical relation1 for the n region and for the p region. This behaviour agrees with the theoretical expectation35 of a scattering rate of ~ vF/(|kF(F′)|d2) by ripples. The fitting result of the lateral size of a ripple (d=35 nm) is within experimentally reasonable values33,34 (on the order of d=10 nm). These results provide an evidence of the presence of ripples in our JJG. We emphasize that we used the same form of in our mechanism of Ic suppression (see equations (2) and (3)), which shows the consistency of our analysis of WAL suppression and Ic suppression.

In addition, the fitting shows that the intervalley scattering rate (by defects and edges) is two orders smaller than the intravalley scattering rate near the CNP. Moreover, it shows that the intervalley scattering rate does not increase as the system approaches the CNP, which is not consistent with the fact that Ic is strongly suppressed near the CNP. Therefore, the WAL data exclude other possible Ic suppression mechanisms by intervalley scattering.

Discussion

We emphasize that the Ambegaokar–Baratoff relation holds in graphene, when both the actual time reversal symmetry and the effective time reversal symmetry are preserved. Our findings show that, contrary to disorders in conventional JJs, ripples in graphene can give rise to Ic suppression, even in the presence of actual time reversal symmetry, because they break the effective time reversal symmetry (for the role of the symmetries, see Supplementary Note 7). Recent theoretical work on JJGs39 has shown that the pseudomagnetic field, induced by strain, strongly affects Andreev-bound states; in this case, however, the critical current was not fully suppressed, because the pseudomagnetic field was not spatially random.

In contrast, the excess resistivity due to ripple scattering in the normal state is given by δρ=(h/4e2)(z4/R2a2), where z (R) is the characteristic height (radius) of the ripple and a=0.14 nm is the interatomic spacing in graphene40. Using typical values of R~5 nm and z~0.2 nm33,41, we obtained δρ~8.3 × 10−4h/e2, which is much smaller than the maximum sheet resistance of Rsheet~0.13 h/e2 at the CNP in this experiment. Thus, the influence of ripple scattering on Rn is minimal.

The origin of the ripples in our JJG is not certain, but presumed to be due to the difference in thermal expansion between the graphene layer and the poly (methyl methacrylate) (PMMA) layer42, overlaid for the top-gate insulation, as inferred from the control experiments. We compared two conventional JJGs, with and without a PMMA layer covering the graphene (see Supplementary Fig. S10). Figure 5 displays the Vbg dependence of the IcRn product, which is normalized by the maximum value in the overdoped region. The JJG without a PMMA layer clearly exhibited the IcRn product at the CNP, at almost one-half the maximum value in the overdoped region; this tendency has been observed and described in previous reports4,10,11. In contrast, the IcRn product was severely suppressed at the CNP for the JJG with a PMMA layer. This result, consistent with our experimental and theoretical results for the p–n junction, can be explained by ripple-induced Ic suppression, with increased ripple density in graphene under the PMMA overlayer. As ripples suppress the supercurrent, but do not significantly affect Rn, the conventional version of the Ambegaokar–Baratoff relation does not hold near the CNP in JJGs with ripples, as shown in Fig. 5.

Figure 5: Control experiments with homogeneous JJGs.
figure 5

Back-gate voltage dependence of normalized IcRn product (ΔVbg=VbgVCNP). The control experimental results from the homogeneous JJG without a PMMA overlayer (blue-filled squares) were very similar to earlier reports from ref. 4 (circles) and ref. 10 (triangles), whereas the homogeneous JJG with a PMMA overlayer (Supplementary Fig. S10b, red-filled squares) exhibited substantial suppression of IcRn at the CNP.

In this study, a bipolar JJG realized the complete supercurrent ON/OFF states for the first time as a genuine graphene supercurrent device. Complete control over the Ic is attributed to the pseudomagnetic field-induced dephasing of the electron–hole phase-coherent Andreev processes in rippled graphene; this provides a unique route for turning off the supercurrent in JJGs, without sacrificing the junction quality in the supercurrent ON state, nor demanding infinite Rn like other nanohybrid supercurrent transistors19,20,21. Completely tuneable hybrid superconductor graphene devices can be used for developing novel superconducting quantum information devices23, such as the Cooper-pair beam splitter24,43 for quantum entanglement.

Methods

Device fabrication

The JJGs3,4 used in this study were fabricated using simple mechanical exfoliation1,2 from graphite flakes (NGS Naturgraphit GmbH) in combination with electron-beam nanofabrication techniques9,11,44. Graphite flakes were exfoliated onto a degenerately doped Si substrate capped with a 300-nm-thick oxidation layer. Visual inspection under an optical microscope allows the identification of single-layer flakes of graphene by the characteristic colouration that arises from optical thin-film interference45. Superconducting Al electrodes were placed on the graphene layer by conventional electron-beam lithography and subsequent electron-beam evaporation of a Ti/Al/Au trilayer (10/90/5 nm) under a base chamber pressure of 5 × 10−8 Torr. The single-layer character of the graphene was confirmed by subsequent quantized Hall conductance measurements in a high magnetic field up to 15 T. The Ti layer was deposited to improve the contact of Al to the graphene layer, whereas the Au capping layer was added to protect the Al from oxidation. To electrically insulate the top gate from the graphene layer, a thin layer of PMMA (molecular mass 950 K, 2% in anisole) was spin-coated onto the graphene layer at 4,000 r.p.m. for 30 s, followed by baking at 180 °C for 120 s on a hot plate. The PMMA was then cross-linked46,47 by exposing it to a 20 keV electron beam at a dose48 of 15,000 μC cm−2. The unexposed PMMA was removed by soaking the entire device in acetone for 15 min (see the yellow-dotted-line boundary of the insulating layer in Fig. 6b). Additional fabrication processes using electron-beam lithography, metal deposition and lift-off were required to define a local top gate (with thickness of 5/30 nm (Ti/Au) and line width of ~50 nm), which covered part of the graphene region between the two superconducting electrodes of the device (see Fig. 6a,b).

Figure 6: Configuration of JJG with a p–n barrier.
figure 6

(a) Schematic diagram of the side view of the device configuration used in this study. (b) Scanning electron microscopy (SEM) picture of the JJG (devices JA and JB) with a graphene p–n barrier (scale bars, 2 μm (left) 0.5 μm (right)). The bright white electrodes applied the top-gate voltage to the graphene layer. White and yellow-dotted lines represent the boundaries of the graphene and the insulating layer for the top gate, respectively. This measurement configuration was used to obtain the data shown in Fig. 1a.

Low-noise measurements

The radio-frequency noise was reduced by adopting a multi-stage filtering scheme: two-stage RC filters (cut-off frequency ~30 kHz) connected in series with leads of a device and silver-powder filters in a cryogenic environment (T<1 K) in conjunction with π-filters arranged at room temperature. In addition, a small magnetic field was applied to cancel out any residual magnetic fields in the cryostat, allowing zero-field measurements.

Figure 6b shows an scanning electron microscopy image of a representative JJG p–n junction device. For the junctions JA and JB, the spacing L between the superconducting electrodes was ~300 nm, while the width W was ~2.2 μm for JA and ~1 μm for JB. A back-gate voltage Vbg was applied between the highly electron-doped Si substrate and a graphene layer. This globally modulated the carrier density in graphene to n=α|Vbg|, where α ~7.3 × 1010 cm−2 V−1 for a 300-nm-thick SiO2 layer49 on the surface of the substrate. A top-gate voltage Vtg was also applied across the PMMA insulating layer. This locally modulated the carrier density in part of the graphene to n′|Vtg| beneath the local surface of the insulating layer, where α′~7.3 × 1011 cm–2 V–1 for the 30-nm-thick cross-linked PMMA layer47,48. The dielectric constant of the insulating layer47 was about εPMMA=4.5, which is similar to the dielectric constant of SiO2. Measurements were carried out at a base temperature of T=50–60 mK, adopting a two-terminal configuration and the conventional ac lock-in technique. In all of our devices, the contact resistance between the graphene and electrode was <1 Ω, which was sufficiently smaller than the resistance of the graphene layer (on the order of a few hundred Ω).

Fitting to critical-current-switching models

MQT behaviour in a current-biased JJ has been used to develop the superconducting phase qubit50. The MQT behaviour in a plain unipolar JJG has been demonstrated9 using the stochastic distribution of Ic, in which the crossover temperature from the classical TA to the MQT regime can be controlled by the application of Vbg, thus allowing modulation of the Josephson coupling energy EJ. To measure the Ic distribution P(Ic) in our JJG, a saw-tooth-like bias current was applied to the sample through a standard resistor R=10 kΩ with a ramping rate of dI/dt=69 μA s−1. At the same time, a high-precision 18-bit data acquisition board (NI-6281) recorded 8,000–12,000 switching events with a current resolution of 0.7 nA at a fixed temperature T, a back-gate voltage Vbg and a top-gate voltage Vtg. A threshold voltage of Vth=3 μV was used to determine Ic.

P(Ic) measured with a linearly increasing bias current has a well-known relationship with the escape rate Γ of a phase particle trapped in a washboard potential51 as P(Ic)=[Γ(Ic)/(dI/dt)]{1— 0 I c P(I′)dI′}, where φ is the phase difference across the JJ, and Ic0 and EJ0(=ħIc0/2e) are the fluctuation-free switching current and Josephson coupling energy, respectively. Γ has different expressions depending on the escape regimes; MQT, TA and PD regimes. The thermal escape rate has an Arrhenius-like temperature dependence as , where at=(1+1/4Q2)1/2−1/2Q is a damping-dependent factor, ωp=ωp0(1−γ2)1/4 is a Josephson plasma frequency and ΔU=2EJ0[(1−γ2)1/2γ cos−1γ] is the potential barrier height, with quality factor Q=4Ic/πIr, ωp0=(2eIc0C)1/2 and normalized current γ=I/Ic0. Here C is the junction capacitance, a key parameter related to the hysteresis of the I–V curve. In contrast to a superconductor–insulator–superconductor JJ, a proximity JJ of normal metal52, carbon nanotubes21, semiconducting nanowires19, or graphene4 has negligibly small geometrical capacitance to allow hysteresis. In this analysis, we adopted the effective capacitance Ceff originating from the diffusive nature of carriers in graphene, replacing the relaxation time RNC by the diffusion time of Andreev pairs52, ħ/EThwhere EThD/L2 is the Thouless energy (D=vFl/2 is the diffusion constant in graphene, vF is the Fermi velocity, l is the mean free path and L is the junction length)52. ΓTA increases exponentially with temperature, which results in broadening of P(Ic) with increasing temperature. In contrast to ΓTA, the MQT rate does not depend on the temperature explicitly as7 . Therefore, below the MQT crossover temperature, , ΓMQT overtakes ΓTA and MQT dominates the escape mechanism, resulting in temperature-independent P(Ic). However, above the PD crossover temperature , a thermally escaped phase particle can be repeatedly retrapped in the following potential well due to strong dissipation. The escape rate ΓPD in the PD regime is modified from ΓTA by retrapping probability53 PRT as . PRT is obtained from integration of the retrapping rate54,55 , with the retrapping potential barrier ΔURT(I)=(EJ0Q02/2)[(IIr0)/Ic0]2, where Ir0 is the fluctuation-free retrapping current and Q0=4Ic0/πIr0. P(Ic) in the TA regime has a left-tailed asymmetric shape. In the PD regime, however, the low-Ic-tailed feature of P(Ic) is reduced by the retrapping of an escaped phase particle, resulting in a sharper and more symmetric P(Ic).

Figure 7a displays the s.d. of the Ic distribution P(Ic), normalized by ‹Ic› as a function of temperature, while passing through the CNP. In this case, Vbg was varied while Vtg remained fixed at 0.5 V. Except for the point corresponding to Vbg=50 V, all other operation points were in the p–p′ regime (see Fig. 3e). The s.d. of the Ic distribution was almost constant with respect to temperature in the MQT regime, while increasing (decreasing) with temperature in the TA (PD) regime, similar to the previous observations for a plain unipolar JJG. Here it should be emphasized that the crossover temperatures of (upward arrows) between the MQT and TA regimes, and (downward arrows) between the TA and PD regimes, changed drastically with Vbg. When Vbg closely approached the CNP, the crossover temperatures tended to shift to much lower temperatures, resulting in the vanishing of down to the base temperature. This feature forms a striking contrast with the MQT behaviour in a plain unipolar JJG, which remains robust even at the CNP. The temperature dependence of the normalized s.d. at the other gate voltages of Vbg and Vtg in the p–p′ regime is shown in Fig. 7b,c,d (see circular symbols in Fig. 3e). The highly sensitive tunability of and with Vbg, which is indebted to the fully controllable Ic in the JJG, was also observed for different Vtg pertaining to the p–n state (not shown).

Figure 7: T dependence of s.d. of P(Ic).
figure 7

T dependence of the s.d. of the switching-current distribution P(Ic) in the junction JA, normalized by the junction critical current Ic (a) at fixed Vtg=0.5 V, (b) at fixed Vbg=−30 V, (c) at fixed Vbg=−21 V and (d) at fixed Vtg=−8 V. Solid lines are guides for the eyes. Upward and downward arrows represent the crossover temperatures between the MQT and TA regimes (), and those between the TA and PD regimes (), respectively.

Figure 3e is overlaid with coloured filled-circle symbols that represent the critical current-switching regimes at the base temperature determined by the temperature dependence of normalized s.d. in Fig. 7. Circular symbols, corresponding to the operation points of the data in Fig. 7, of a colour belong to a particular switching regime. The switching current regimes, determined by varying temperature at a fixed Vbg and Vtg in Fig. 7, are consistent with the boundaries of the different regimes determined by varying Vbg at the base temperature in Fig. 3e.

WAL analysis

We analysed the magnetoconductance measured in our JJG at T=4.8 K, which is above Tc, and at Vtg=0.5 V. The magnetoconductance of Δσ=σ(B)−σ(0) can be described by the series conductance of the two regions, n and p, as

where Δσn(p)=σn(p)(B)−σn(p)(0) is the magnetoconductance of n(p). It is known38 that the WAL in Δσn,p is suppressed, depending on the intravalley scattering rate (), intervalley scattering rate () and the dephasing rate () of each region:

Here, , ψ is the digamma function, , Dn(p) is the diffusion constant determined by the gate voltage applied to n(p) and B is the external magnetic field perpendicular to graphene. In equation (5), we introduce a parameter α to compensate for the aspect ratio (L/W) associated with the geometry, due to the modification of L by screening of the metallic electrodes and other factors. We chose σn(0)/(σn(0)+σp(0))≈0.5 as a reasonable value. The scattering rates were obtained by fitting the magnetoconductance data with equation (5) using seven fitting parameters (one for α, three rates for region n and another three for region p, because scattering occurs at different energies for the n and p regions). Note that (i) the number of fitting parameters is effectively four, the minimum number used in most WAL studies, because the rates for the n region are nearly the same as those for the p region in the case of a small Vtg; and (ii) because the digamma function has a sufficiently non-trivial dependence on its argument, the overall factor α is not important in determining the rates in the fitting, because the dependence of equation (5) on B is not governed by α, but by the rates. In the fitting, we used two physically reasonable constraints: (i) that the intervalley scattering rate is smaller than the intravalley scattering rate56, and (ii) that all scattering rates are smaller than the momentum relaxation rate33,38,

Additional information

How to cite this article: Choi, J. -H. et al. Complete gate control of supercurrent in graphene p–n junctions. Nat. Commun. 4:2525 doi: 10.1038/ncomms3525 (2013).