A superconducting quantum constriction is realized by locally screening ionic liquid gating in SrTiO3.
Abstract
Superconductivity in SrTiO3 occurs at remarkably low carrier densities and therefore, unlike conventional superconductors, can be controlled by electrostatic gates. Here, we demonstrate nanoscale weak links connecting superconducting leads, all within a single material, SrTiO3. Ionic liquid gating accumulates carriers in the leads, and local electrostatic gates are tuned to open the weak link. These devices behave as superconducting quantum point contacts with a quantized critical supercurrent. This is a milestone toward establishing SrTiO3 as a single-material platform for mesoscopic superconducting transport experiments that also intrinsically contains the necessary ingredients to engineer topological superconductivity.
INTRODUCTION
Conductance quantization in ballistic quantum point contacts (QPCs) is a notable example of departure from the classical Drude picture of electrical conductivity set by the rate of charge carrier scattering (1). When a constriction between two electron reservoirs is sufficiently narrow and disorder free, its conductance becomes quantized according to the number of occupied modes: discrete transverse momenta allowed within the constriction’s confinement potential. Each mode contributes a conductance quantum δG = 2e2/h (spin-degenerate case), a value that does not depend on the exact geometry of the device.
A related phenomenon is expected to arise in a constriction between two superconducting reservoirs (2, 3), i.e., a superconducting QPC (SQPC). Again, the transverse momentum spectrum becomes discretized under the constriction confinement potential. The supercurrent carried by each mode is determined by the Andreev bound state (ABS) spectrum, which is typically a function of constriction geometry. SQPCs are thus characterized by quantized critical supercurrent Ic with a nonuniversal step height δIc. However, in the limit of a short junction length, only one ABS per ballistic mode remains, and the current carried by each mode can reach a maximum value δIc = eΔ/ℏ. This ideal step height is again geometry independent and scales only with the superconducting gap Δ.
The widespread route for fabricating gate-tunable superconducting weak links has been to combine two optimal components in a hybrid system: a clean semiconductor (typically a III-V semiconductor or Ge) and metallic superconducting leads (e.g., Nb and Al). These hybrid systems have been successfully used to demonstrate quantized critical supercurrent, but with quantization step heights far below eΔ/ℏ (4–9). The two major challenges for reaching the universal limit for quantized supercurrent are the geometric requirement that the distance between superconducting leads be much less than the superconducting coherence length ξ and the need for near-perfect semiconductor/superconductor contact transparency (3). Achieving the latter in hybrid semiconductor-superconductor systems has been a major materials science challenge that has required deployment of in situ heteroepitaxial growth techniques (10).
An alternate route taken in this work is to form both leads and constriction in a single electrostatically tunable superconducting material, such as SrTiO3 (STO). Working within a single-material platform is attractive for fabricating SQPCs, as the superconductor/normal metal (SN) boundary can be purely electronic (no structural discontinuity) and thus potentially highly transparent.
One of STO’s remarkable aspects is superconductivity in the extremely dilute charge carrier density limit (11, 12). In two-dimensional (2D) electron systems (2DESs) at the surface of STO, such as LaAlO3 (LAO)/STO, LaTiO3/STO, and ionic liquid–gated STO, superconductivity occurs in the range of 0.01 electrons per unit cell (13, 14). Consequently, one can electrostatically control the transition between superconductor, normal metal, and insulator in this material. On the macroscopic scale, this control is well established using back gating through the STO substrate, top gating through a dielectric layer, and ionic liquid gating (13–18).
More recently, several approaches have emerged for nanoscale patterning of conduction in LAO/STO, leading to demonstration of quantization effects in normal state, but, so far, not in superconducting transport. Realization of a conventional split-gate QPC geometry in LAO/STO is challenging as it involves depleting and/or accumulating charge densities of at least ≈1013 cm–2, close to the limit of conventional dielectrics. Spatial inhomogeneity and relatively short mean free paths in these 2DESs present another challenge, leading patterned constrictions to often be dominated by tunneling through accidental quantum dots (19–21). A QPC with normal state but not superconducting conductance quantization has recently been demonstrated in underdoped, nonsuperconducting LAO/STO (22). In (21), a constriction defined by split gates with normal state conductance about half of a single spin–degenerate ballistic mode was estimated to have corresponding partially transmitting single-mode supercurrent, although it did not show direct effects of quantization. A different technique is to write conductive channels on LAO/STO with voltage-biased atomic force microscope (AFM) tips. This method enabled demonstration of quantum wires and dots coupled by tunnel barriers to superconducting leads, with quantized normal-state transport and indirect signatures of electron pairing (23–26) but not superconductivity.
In this work, we demonstrate quantized supercurrent in QPCs in a split-gate geometry based on ionic liquid–gated STO. We observe a discretized step structure in the critical current, with tuning from zero to three ballistic modes. Step height per mode δIc is only three to five times smaller than the canonical value eΔ/ℏ, as close to ideal as achieved in any hybrid system (6). The fabrication process of our devices is enabled by the fine patterning of local electrostatic gates using liftoff of metal and atomic layer–deposited Hafnia (HfO2) with a feature size close to 40 nm. This is distinct from the approaches taken in previous works on LAO/STO weak links (20–22, 27–29). Notably, we avoid an epitaxial growth step at high temperature, which complicates the workflow for patterning and potentially introduces disorder [see, e.g., (30, 31)]. We thus consider this fabrication technique an attractive alternative for further development of STO as a platform for mesoscale superconducting devices.
RESULTS
Our devices are 20-μm-wide Hall bars covered by ionic liquid, which is polarized to accumulate a 2D carrier density at any exposed STO surface. The coarse contours of the Hall bar are defined by patterning an insulating SiO2 layer, which separates the surface of undoped STO from the ionic liquid (Fig. 1, A and B); underneath the SiO2, the STO surface remains insulating, while the carrier density in the Hall bar region is tuned into the superconducting regime. Split gates with thin, self-aligned HfO2 dielectrics define 40-nm-wide constrictions (Fig. 1C) between neighboring superconducting reservoirs. The design includes five or six ohmic contacts on each side of the split gates (Fig. 1B) to enable four-terminal measurements of both the constriction and the adjacent superconducting leads.
Fig. 1. Electrostatically defined constriction in superconducting SrTiO3.
(A) Schematic cross section of the device and illustration of the gate voltage definitions. (B) Confocal laser microscope image of the Hall bar region of the device and illustration of the measurement scheme. The dashed arrow indicates the location of the cross section in (A). (C) Scanning electron microscope image of the constriction region on a reference device. (D) Superconducting transition in the constriction and lead resistance. “Right” and “Left” refer to measurement of Vlead on both sides of the constriction. (E) Constriction conductance map with temperature and split gate voltage. (F) Constriction conductance map with magnetic field and local gate voltage. Symbols in (F) indicate the selected gate voltage values for which line cuts in field are shown in (G). Lead resistance at extremes of VG12 is also shown in (G) to illustrate the independence of local gate voltage. The top axis shows the mapping from critical field Bc (red circles) to the coherence length. The estimated ξ is shown in (C) for comparison with device dimensions, along with the mean free path from Hall measurements in the leads (see section S4). In (D) to (G), VGIL = 3 V and VBG = 50 V.
The carrier density profile is electrostatically defined by voltages on four gates, as illustrated in Fig. 1A: a large coplanar gate that controls the polarization of the ionic liquid (VGIL), a back gate (VBG) and two split gates (VG1 and VG2, denoted as VG12 for the case VG1 = VG2). VGIL and VBG are “global” gates that tune the carrier density and vertical confinement in both the 2DES leads and the constriction. VG12 is a “local” gate that only tunes carrier density and lateral confinement in the constriction. VGIL is set when the device is near room temperature and maintained as the sample is cooled below the freezing temperature of the ionic liquid (220 K). VGIL is used to polarize a drop of ionic liquid that covers both the coplanar gate electrode and the device. At lower temperatures, the polarization of the ionic liquid is frozen in. VGIL is the primary control knob for the carrier density in the leads, which can be tuned from ≈5 × 1012 to 1014 cm−2 (32, 33). The superconducting transition temperature as a function of density has a maximum near 3 × 1013 cm−2 (see section S3). The main results presented here will focus on this nearly optimally doped state obtained by cooling the device under VGIL = +3 V. For additional data on the second constriction on the right side of the Hall bar in Fig. 1B, different devices, and cooldowns with carrier density tuned across a larger range, see sections S2 to S6.
The voltage VBG on a back gate contacting the bottom of the STO crystal provides additional global tuning of the 2DES at base temperature, primarily by modulating the depth of the 2DES. For most experiments on this device, we set VBG = +50 V to pull the electron density farther away from surface disorder [see (34) and section S4].
Figure 1D shows the superconducting transition Tc measured by sourcing a small AC excitation through a constriction at VG12 = +3 V and VBG = +50 V. In the following, constriction resistance and conductance will be denoted as R = dVQPC/dIAC and G = 1/R and the resistances of the leads as Rlead = dVlead/dIAC (see Fig. 1B and Materials and Methods for more details). On both sides of the constriction, Rlead shows a sharp transition near 350 mK. This is near the optimal Tc value for 2D STO (14, 17). The measured Hall density of 3.05 × 1013 cm−2 and the slight increase of Tc by 20 mK upon removing the back-gate voltage suggest that this device state is slightly on the overdoped side of the superconducting dome (see section S4).
The constriction resistance R also starts decreasing near the lead Tc, but its transition to zero resistance (within accuracy of our measurement) is significantly broader than that of the leads. Decreasing VG12 suppresses both the zero resistance state and the normal state conductance and eventually pinches off the weak link (Fig. 1, E and F). At base temperature, superconductivity can also be suppressed by a perpendicular magnetic field (Fig. 1F). Using ξ2 = Φ0/(2πBC) (35), with Φ0 = h/2e being the flux quantum, the critical field Bc = 130 to 140 mT in the constriction yields an estimated coherence length of ξ = 50 nm (43 nm in the leads). This estimate is consistent with the dirty-limit Bardeen-Cooper-Schrieffer superconductor picture (36, 37) in which the coherence length is set by the mean free path LMFP. From Hall measurements on the leads, we extract a Hall mobility μ = 600 cm2/Vs and LMFP = 55 nm.
The shortness of these length scales illustrates the challenge of fabricating QPCs and SQPCs in STO (see Fig. 1C). Observing ballistic transport requires junction length L < LMFP. Achieving a single-ABS junction with critical current quantization also requires short junction length: L < ξ. Although the junction length is not well defined in a split-gate geometry, we fabricated the gates with very narrow lateral spacing (40 nm) and sharp tips to strive for the ballistic (or quasi-ballistic) regime.
The ballistic nature of the SQPC is most apparent in differential resistance at finite DC current. Filling of states in the constriction with VG12 results in a staircase shape of the critical supercurrent Ic(VG12) (Fig. 2). Adopting a definition of Ic as the current at which the resistance R is halved with respect to the high-VDC normal state, plateaus at both positive and negative integer multiples of δIc = 2.48 nA are seen in the VG12 − IDC map of constriction resistance normalized to its normal state value (Fig. 2B).
Fig. 2. Critical current quantization.
(A) DC current dependence of constriction and lead resistances at VG12 = 3V. (B) Constriction resistance, normalized to normal state resistance at VDC = 100 μV. The solid red line indicates the critical current Ic. The dashed lines indicate 1, 2, and 3 integer multiples of δIc = 2.48 nA. (C) VG12 dependence of Ic normalized to δIc and (D) normal state conductance GN at VDC = 100 μV, with a series resistance of 800 Ω subtracted from the raw data. The shaded connection between (C) and (D) emphasizes the numerical correspondence in the observed number of ballistic modes n. The dashed line in (C) is a fit to the saddle potential QPC model (see section S1). (E) Split- and back-gate voltage dependence of zero-bias conductance above Tc. G has been corrected for a variable series resistance gradually increasing from 1.15 to 2.1 kilohm. Short plateaus can be seen at integer multiples of 2e2/h (n = 1, 2, and hints at higher multiples). Unintentional Coulomb blockade levels can be seen near 0.2e2/h, e2/h, and 2.5e2/h.
In the ballistic SQPC picture, Ic/δIc corresponds to n, which is the number of ballistic modes below the Fermi energy in the constriction (Fig. 2C). The first mode plateau is intermittent as a function of gate voltage due to resonant transmission through the weak link, correlated with the charging levels of an accidental Coulomb blockade observed near pinch-off at low VG12 (see section S7), whereas the second and third plateaus are more stable. An alternative way to estimate the number of modes is from normal state conductance GN, where each fully transmitting spin-degenerate mode is expected to contribute a conductance δG = 2e2/h. The number of modes inferred by dividing GN by this increment matches that extracted from the sequence of steps in supercurrent. We also see hints of plateaus in normal state conductance at fixed VDC = 100 μV near n = 1 and 2 (Fig. 2D). Features suggestive of normal state conductance quantization are more clearly apparent above Tc (Fig. 2E), where one does not need to apply a DC bias to suppress the supercurrent, and disorder-induced fluctuations are reduced. The plateau structure persists as a function of back gate voltage (detailed further in section S5).
Ideally, the magnitude of steps in Ic through a constriction should scale only with the superconducting gap as
| (1) |
This scaling is expected to hold for a short junction (L ≪ ξ) with perfectly transparent SN contacts (2, 3).
For most experimental realizations of SQPCs in hybrid metal superconductor/semiconductor devices, neither of these requirements is fully satisfied, and δIc is generally suppressed by at least an order of magnitude (4, 5, 7–9). One work on Si/Ge nanowires with Nb contacts reported suppression by only a factor of 2.9 (6). In our case, data in Fig. 2 suggest a comparable factor of 3 to 5. The uncertainty comes from the choice of method to extract Δ (see section S8): from Tc of the constriction [δIc/(eΔ/ℏ) = 2.9], from Tc of the leads [δIc/(eΔ/ℏ) = 4.1], or from the temperature dependence of the excess current [δIc/(eΔ/ℏ) = 4.8].
Analysis of the excess current Iexc allows separating the role of imperfect SN contact transparency τSN from that of finite junction length. We define Iexc as the zero-bias intercept of the normal-state resistance extrapolated from high VDC (Fig. 3A). Both Ic and Iexc are expected to scale with GN (see section S1) and, thus, approximately track each other with VG12. However, these two quantities encode different physics: the shape of the ABS spectrum for Ic (2) and the balance between ordinary and Andreev reflections for Iexc (38). The quantity eIexcRN/Δ can be nonlinearly mapped onto τSN following the treatment of Andreev reflections in an superconductor/normal metal/superconductor (SNS) junction in (39, 40). Over the gate voltage range with a well-defined and quantized supercurrent (1.5 < VG12 < 2.5), we thereby extract .
Fig. 3. SN transparency and junction length.
(A) The DC current–voltage curve of the constriction at VG12 = 3 V and the definition of the excess current Iexc. (B) Split-gate voltage dependence of the excess and critical currents. (C) eIexcRN/Δ, the input quantity of the SNS model in (39, 40), and its mapping onto SN boundary transparency τSN. (D) δIc suppression by finite transparency and finite junction length. Comparison to the ballistic short-limit model (2) (solid black line), full calculation at L/ξ = 0.56 (3) (blue squares), and the approximate correction for arbitrary L/ξ from (41) (dashed lines). The shaded region reflects in (C) and error bars in (D) reflect the uncertainty on the gap (see section S8).
In the short junction limit L ≪ ξ (2), we can predict the suppression of δIc as a function of τSN (Fig. 3D). The experimentally measured δIc is only slightly below the theoretical curve, and its full suppression can be accounted for by multiplying it by an additional factor α= 0.7. This additional suppression can be explained by considering the finite length of the junction. An approximate theoretical description obtained in (41) is α = 1/(1 + L/2ξ), which is in good agreement with calculations for the case in (3), where L = 0.56ξ. In this work, assuming that α = 0.7 yields L = 0.85ξ = 42 nm, which is close to the 40-nm lithographic width of our QPC.
DISCUSSION
Transparency is likely to be the main driver for the reduction in δIc from its ideal value despite being competitive with the hybrid III-V/superconductor systems, where τSN is typically estimated below 0.85 (7–9, 42) except for pristine epitaxial interfaces (10). An advantage of our single-material system is that the SN contact interface is electrostatically defined and presumably does not have a structural discontinuity. In our present realization, transparency is likely limited by the smooth gate-induced density variation, which, in turn, entails a gradually varying order parameter. We anticipate that τSN can be further improved by manipulating the SN boundary with additional local gates near the weak link.
Furthermore, we anticipate improvements by increasing the mean free path. In ionic liquid–gated STO and LAO/STO, LMFP is typically less than 100 nm. However, improvements to μ > 104 cm2/Vs and LMFP> 1 μm have been demonstrated by separating the ionic liquid from the channel by an ultrathin spacer layer (16), band engineering with spacer layers in LAO/STO (43), or forming the channel from high-quality molecular beam epitaxy–grown STO in the 3D case (44). These demonstrations have so far only been accomplished in unpatterned or coarsely patterned films, but we anticipate that these approaches can be compatible with the nanopatterning technique developed here. The fabrication route used in this work is relatively simple—based on commercially available STO crystals, avoiding epitaxial growth steps—so complex patterning, device, or heterostructure design refinements could be added without rendering it unwieldy.
Using ionic liquid–gated STO as a platform, we have realized SQPCs with quantized critical supercurrent, tunable between zero and three ballistic modes by split gates. This is a first realization of a quantized gate-tunable SQPC in a single material system, enabling highly transparent SN contacts without structural discontinuity at the boundary. This work establishes spatially patterned screening of ionic liquid from an STO surface as a promising alternative to existing methods for nanoscale patterning of conduction and superconductivity in STO: patterning LAO/STO with pregrowth templates (19, 20, 28, 29), electrostatic depletion by patterned gates (21, 22, 27), or conductive channel writing by voltage-biased AFM tips (23–26, 45). Our method appears particularly suited for realizations of ballistic superconducting transport, which require maintaining high carrier densities within nanopatterned constrictions. Naturally occurring depletion near the edges of an STO-based conducting channel (45, 46) can be counteracted with local gates as we have shown.
Our approach may also be especially attractive for exploring topological superconductivity in several contexts. Combining ballistic transport with superconductivity, strong spin-orbit coupling, and tunable dimensionality offers hope for engineering extrinsic topological superconductivity in 1D nanostructures (47–49). Even an unpatterned STO 2DES may host intrinsic topological superconductivity in certain conditions because of interplay between its multiorbital band structure, spin-orbit coupling, and ferroelectricity (50–52). A ballistic point contact similar to the SQPC demonstrated here could serve as the tunnel probe central to many detection schemes for the resulting Majorana bound states (53–55). The single-mode ballistic Josephson junction regime demonstrated here is also a requisite ingredient of theoretical proposals for realizing topological ABS spectra in multiterminal junctions (56, 57). Last, this work is an important step toward realizing controlled negative U quantum dots (20, 23) in the classic geometry of an “island” coupled to two QPCs (58).
MATERIALS AND METHODS
Fabrication is based on commercially (001) oriented STO single-crystal substrates, purchased from MTI. To obtain a Ti-terminated surface with terrace-step morphology, these substrates were soaked in heated deionized water for 20 min and annealed at 1000°C for 2 hours in flowing Ar and O2 in a tube furnace.
All subsequent patterning was performed with liftoff processes using e-beam–patterned poly(methyl methacrylate) (PMMA) 950K, 4% in anisole for the first step and 8% for all subsequent steps. The first step is the local split-gate pattern, written on a 100-kV e-beam write system. Atomic layer deposition was used to deposit 15 nm of HfO2 (100 cycles of Hf precursor and water.) The deposition stage temperature was 85°C. We note the importance of loading the sample and starting the deposition quickly to avoid PMMA pattern reflow. The 5-nm Ti/50-nm Au gate contact was then deposited by e-beam evaporation. Liftoff of both HfO2 and Ti/Au layers was then performed by soaking in heated N-Methyl-2-pyrrolidone (NMP), followed by ultrasonication in acetone.
The remaining patterning was performed with a 30-kV e-beam write system. The second step is the gate contact using liftoff of 40-nm Ti/100-nm Au in acetone. The third step is the ohmic contact deposition. It requires exposing the pattern to Ar+ ion milling before e-beam evaporation of 10-nm Ti/80-nm Au followed by liftoff in acetone. The fourth patterning step is the mesa insulation, deposited by magnetron sputtering 70 nm of SiO2, followed by liftoff in acetone. The measured devices were imaged with a conventional optical microscope and with a Keyence VK-X confocal laser microscope. Scanning electron microscope imaging was performed on reference patterns written on the same chips.
Finished devices were annealed for 20 min at 150°C in air. The back gate contact to a gold pad on an alumina ceramic chip carrier was made with silver paste. Immediately after depositing a drop of ionic liquid diethylmethyl(2-methoxyethyl)ammonium bis(trifluoromethylsulfonyl)imide (DEME-TFSI) to cover both the device and the surrounding side gate, the samples were loaded into the dilution refrigerator system then vacuum-pumped overnight to minimize contamination of the ionic liquid by water from exposure to air.
The ionic liquid gate voltage VGIL was slowly ramped up to desired value at room temperature, followed by several minutes of stabilization and then rapid-cooling the measurement probe below the freezing point of DEME-TFSI (220 K).Typical measured 2-terminal resistance per successful ohmic contact was 3 to 10 kilohm, which includes a 2- to 3-kilohm contribution from the measurement lines and built-in radio frequency filters in the probe. All measurements presented in the main text and in the Supplementary Materials were performed in a four-probe configuration (with the exception of fig. S9). Measurements were performed by voltage-sourcing nominal AC and DC excitations ( and ) through an adder circuit and measuring the drained current (IAC and IDC). VQPC and VDC refer to the AC and DC components of the voltage drop across the weak link measured at the adjacent voltage probes (as shown in Fig. 1B). The constriction resistance is R = 1/G = dVAC/dIAC. Vlead is the AC voltage drop between the next adjacent pair of voltage probes. The resistance of the unpatterned 2DES is Rlead = Vlead/IAC.
Measured R contains two contributions: the resistance of the constriction itself (tuned by VG12) and a series resistance RS. Unless specified, R and G are shown as measured. In selected plots of G in the normal state, a specified VG12-independent value of RS was subtracted from the measured R. As discussed in more detail in section S5, RS was chosen to match the plateau structure to integer multiples of e2/h, with RS/Rlead ≈ 2 to 3. On the basis of the device geometry, RS is expected to be approximately equal or larger than Rlead.
The Supplementary Materials to this report present extensive additional discussion and characterization of a total of six devices fabricated on three different STO chips. The contents of each supplementary section are briefly summarized below to facilitate their navigation. Section S1: theoretical framework for the ballistic SNS constriction model, IC dependence on τNS, L, and VG12, IEXC dependence on τNS. Section S2: description of additional devices and their fabrication. Section S3: tuning of carrier density and superconductivity with VGIL. Section S4: VG12 and VBG effect on R, Rlead, Bc, Tc, Ic, and Hall density. Additional characterization of transport through the constriction in the normal state (section S5), in the superconducting state (section S6), and in the tunneling and accidental Coulomb blockade regimes (section S7). Section S8: discussion of the different methods to extract the superconducting gap Δ and the associated uncertainty.
Acknowledgments
We acknowledge M. Kastner, M. Beasley, and E. Fox for helpful discussions. We acknowledge R. Tiberio and M. Rincon for help with device fabrication. Funding: Experimental work (fabrication and measurement) by E.M. was primarily supported by the Air Force Office of Scientific Research through grant no. FA9550-16-1-0126. E.M. was also supported by the Nano- and Quantum Science and Engineering Postdoctoral Fellowship at Stanford University. For the final stages of the work (primarily extensive data analysis and manuscript writing), E.M. and D.G.-G.’s engagement was supported by the U.S. Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division under contract DE-AC02-76SF00515 (in E.M.’s case, in conjunction with internal Stanford University funds). Measurement contribution by I.T.R. was also supported by the U.S. Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division, under contract DE-AC02-76SF00515. I.T.R. acknowledges support from the ARCS foundation during the final stages of the work. Measurement infrastructure was funded in part by the Gordon and Betty Moore Foundation’s EPiQS Initiative through grants GBMF3429 and GBMF9460. Part of this work was performed at the Stanford Nano Shared Facilities (SNSF)/Stanford Nanofabrication Facility (SNF), supported by the National Science Foundation under award ECCS-1542152. Author contributions: E.M. and D.G.-G. designed the experiment. E.M. fabricated the devices. E.M. and I.T.R. performed the measurements. E.M. carried out data analysis. All authors discussed the results and wrote the manuscript. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials. Data for all graphs presented here are available at (59).
Supplementary Materials
This PDF file includes:
Supplementary Text
Sections S1 to S8
Figs. S1 to S40
References
REFERENCES AND NOTES
- 1.Büttiker M., Quantized transmission of a saddle-point constriction. Phys. Rev. B 41, 7906–7909 (1990). [DOI] [PubMed] [Google Scholar]
- 2.Beenakker C. W. J., Van Houten H., Josephson current through a superconducting quantum point contact shorter than the coherence length. Phys. Rev. Lett. 66, 3056–3059 (1991). [DOI] [PubMed] [Google Scholar]
- 3.Furusaki A., Takayanagi H., Tsukada M., Josephson effect of the superconducting quantum point contact. Phys. Rev. B 45, 10563–10575 (1992). [DOI] [PubMed] [Google Scholar]
- 4.Takayanagi H., Akazaki T., Nitta J., Observation of maximum supercurrent quantization in a superconducting quantum point contact. Phys. Rev. Lett. 75, 3533–3536 (1995). [DOI] [PubMed] [Google Scholar]
- 5.Bauch T., Huerfeld E., Krasnov V. M., Delsing P., Takayanagi H., Akazaki T., Correlated quantization of supercurrent and conductance in a superconducting quantum point contact. Phys. Rev. B 71, 174502 (2005). [Google Scholar]
- 6.Xiang J., Vidan A., Tinkham M., Westervelt R. M., Lieber C. M., Ge/Si nanowire mesoscopic Josephson junctions. Nat. Nanotechnol. 1, 208–213 (2006). [DOI] [PubMed] [Google Scholar]
- 7.Abay S., Persson D., Nilsson H., Xu H. Q., Fogelstrom M., Shumeiko V., Delsing P., Quantized conductance and its correlation to the supercurrent in a nanowire connected to superconductors. Nano Lett. 13, 3614–3617 (2013). [DOI] [PubMed] [Google Scholar]
- 8.Irie H., Harada Y., Sugiyama H., Akazaki T., Josephson coupling through one-dimensional ballistic channel in semiconductor-superconductor hybrid quantum point contacts. Phys. Rev. B 89, 165415 (2014). [Google Scholar]
- 9.Hendrickx N. W., Tagliaferri M. L. V., Kouwenhoven M., Li R., Franke D. P., Sammak A., Brinkman A., Scappucci G., Veldhorst M., Ballistic supercurrent discretization and micrometer-long Josephson coupling in germanium. Phys. Rev. B 99, 075435 (2019). [Google Scholar]
- 10.Kjærgaard M., Suominen H. J., Nowak M. P., Akhmerov A. R., Shabani J., Palmstrøm C. J., Nichele F., Marcus C. M., Transparent semiconductor-superconductor interface and induced gap in an epitaxial heterostructure Josephson junction. Phys. Rev. Appl. 7, 034029 (2017). [Google Scholar]
- 11.Lin X., Zhu Z., Fauqué B., Behnia K., Fermi surface of the most dilute superconductor. Phys. Rev. X 3, 021002 (2013). [Google Scholar]
- 12.Gastiasoro M. N., Ruhman J., Fernandes R. M., Superconductivity in dilute SrTiO3: A review. Ann. Phys. 417, 168107 (2020). [Google Scholar]
- 13.Bell C., Harashima S., Kozuka Y., Kim M., Kim B. G., Hikita Y., Hwang H. Y., Dominant mobility modulation by the electric field effect at the LaAlO3/SrTiO3 interface. Phys. Rev. Lett. 103, 226802 (2009). [DOI] [PubMed] [Google Scholar]
- 14.Joshua A., Pecker S., Ruhman J., Altman E., Ilani S., A universal critical density underlying the physics of electrons at the LaAlO3/SrTiO3 interface. Nat. Commun. 3, 1129 (2012). [DOI] [PubMed] [Google Scholar]
- 15.Caviglia A. D., Gariglio S., Reyren N., Jaccard D., Schneider T., Gabay M., Thiel S., Hammerl G., Mannhart J., Triscone J.-M., Electric field control of the LaAlO3/SrTiO3 interface ground state. Nature 456, 624–627 (2008). [DOI] [PubMed] [Google Scholar]
- 16.Gallagher P., Lee M., Petach T. A., Stanwyck S. W., Williams J. R., Watanabe K., Taniguchi T., Goldhaber-Gordon D., A high-mobility electronic system at an electrolyte-gated oxide surface. Nat. Commun. 6, 6437 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Chen Z., Swartz A. G., Yoon H., Inoue H., Merz T. A., Lu D., Xie Y., Yuan H., Hikita Y., Raghu S., Hwang H. Y., Carrier density and disorder tuned superconductor-metal transition in a two-dimensional electron system. Nat. Commun. 9, 4008 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Christensen D. V., Trier F., Niu W., Gan Y., Zhang Y., Jespersen T. S., Chen Y., Pryds N., Stimulating Oxide Heterostructures: A review on controlling SrTiO3-Based heterointerfaces with external stimuli. Adv. Mater. Interfaces 6, 1900772 (2019). [Google Scholar]
- 19.Maniv E., Ron A., Goldstein M., Palevski A., Dagan Y., Tunneling into a quantum confinement created by a single-step nanolithography of conducting oxide interfaces. Phys. Rev. B 94, 045120 (2016). [Google Scholar]
- 20.Prawiroatmodjo G. E. D. K., Leijnse M., Trier F., Chen Y., Christensen D. V., Von Soosten M., Pryds N., Jespersen T. S., Transport and excitations in a negative-U quantum dot at the LaAlO3/SrTiO3 interface. Nat. Commun. 8, 395 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Thierschmann H., Mulazimoglu E., Manca N., Goswami S., Klapwijk T. M., Caviglia A. D., Transport regimes of a split gate superconducting quantum point contact in the two-dimensional LaAlO3/SrTiO3 superfluid. Nat. Commun. 9, 2276 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Jouan A., Singh G., Lesne E., Vaz D. C., Bibes M., Barthélémy A., Ulysse C., Stornaiuolo D., Salluzzo M., Hurand S., Lesueur J., Feuillet-Palma C., Bergeal N., Quantized conductance in a one-dimensional ballistic oxide nanodevice. Nat. Electron. 3, 201–206 (2020). [Google Scholar]
- 23.Cheng G., Tomczyk M., Lu S., Veazey J. P., Huang M., Irvin P., Ryu S., Lee H., Eom C.-B., Hellberg C. S., Levy J., Electron pairing without superconductivity. Nature 521, 196–199 (2015). [DOI] [PubMed] [Google Scholar]
- 24.Cheng G., Tomczyk M., Tacla A. B., Lee H., Lu S., Veazey J. P., Huang M., Irvin P., Ryu S., Eom C.-B., Daley A., Pekker D., Levy J., Tunable electron-electron interactions in LaAlO3/SrTiO3 nanostructures. Phys. Rev, X 6, 041042 (2016). [Google Scholar]
- 25.Annadi A., Cheng G., Lee H., Lee J.-W., Lu S., Tylan-Tyler A., Briggeman M., Tomczyk M., Huang M., Pekker D., Eom C.-B., Irvin P., Levy J., Quantized ballistic transport of electrons and electron pairs in LaAlO3/SrTiO3 nanowires. Nano Lett. 18, 4473–4481 (2018). [DOI] [PubMed] [Google Scholar]
- 26.Briggeman M., Tomczyk M., Tian B., Lee H., Lee J.-W., He Y., Tylan-Tyler A., Huang M., Eom C.-B., Pekker D., Mong R. S. K., Irvin P., Levy J., Pascal conductance series in ballistic one-dimensional LaAlO3/SrTiO3 channels. Science 367, 769–772 (2020). [DOI] [PubMed] [Google Scholar]
- 27.Goswami S., Mulazimoglu E., Vandersypen L. M. K., Caviglia A. D., Nanoscale electrostatic control of oxide interfaces. Nano Lett. 15, 2627–2632 (2015). [DOI] [PubMed] [Google Scholar]
- 28.Monteiro A. M. R. V. L., Groenendijk D. J., Manca N., Mulazimoglu E., Goswami S., Blanter Y., Vandersypen L. M. K., Caviglia A. D., Side gate tunable Josephson junctions at the LaAlO3/SrTiO3 interface. Nano Lett. 17, 715–720 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Stornaiuolo D., Massarotti D., Di Capua R., Lucignano P., Pepe G. P., Salluzzo M., Tafuri F., Signatures of unconventional superconductivity in the LaAlO3/SrTiO3 two-dimensional system. Phys. Rev. B 95, 140502 (2017). [Google Scholar]
- 30.Balakrishnan P. P., Veit M. J., Alaan U. S., Gray M. T., Suzuki Y., Metallicity in SrTiO3 substrates induced by pulsed laser deposition. APL Mater. 7, 011102 (2019). [Google Scholar]
- 31.Schneider C. W., Döbeli M., Richter C., Lippert T., Oxygen diffusion in oxide thin films grown on SrTiO3. Phys. Rev. Mater. 3, 123401 (2019). [Google Scholar]
- 32.Ueno K., Nakamura S., Shimotani H., Ohtomo A., Kimura N., Nojima T., Aoki H., Iwasa Y., Kawasaki M., Electric-field-induced superconductivity in an insulator. Nat. Mater. 7, 855–858 (2008). [DOI] [PubMed] [Google Scholar]
- 33.Lee M., Williams J. R., Zhang S., Frisbie C. D., Goldhaber-Gordon D., Electrolyte gate-controlled Kondo effect in SrTiO3. Phys. Rev. Lett. 107, 256601 (2011). [DOI] [PubMed] [Google Scholar]
- 34.Chen Z., Yuan H., Xie Y., Lu D., Inoue H., Hikita Y., Bell C., Hwang H. Y., Dual-gate modulation of carrier density and disorder in an oxide two-dimensional electron system. Nano Lett. 16, 6130–6136 (2016). [DOI] [PubMed] [Google Scholar]
- 35.Kim M., Kozuka Y., Bell C., Hikita Y., Hwang H. Y., Intrinsic spin-orbit coupling in superconducting δ-doped SrTiO3 heterostructures. Phys. Rev. B 86, 085121 (2012). [Google Scholar]
- 36.Bert J. A., Nowack K. C., Kalisky B., Noad H., Kirtley J. R., Bell C., Sato H. K., Hosoda M., Hikita Y., Hwang H. Y., Moler K. A., Gate-tuned superfluid density at the superconducting LaAlO3/SrTiO3 interface. Phys. Rev. B 86, 060503 (2012). [Google Scholar]
- 37.Collignon C., Fauqué B., Cavanna A., Gennser U., Mailly D., Behnia K., Superfluid density and carrier concentration across a superconducting dome: The case of strontium titanate. Phys. Rev. B 96, 224506 (2017). [Google Scholar]
- 38.Blonder G. E., Tinkham M., Klapwijk T. M., Transition from metallic to tunneling regimes in superconducting microconstrictions: Excess current, charge imbalance, and supercurrent conversion. Phys. Rev. B 25, 4515–4532 (1982). [Google Scholar]
- 39.Octavio M., Tinkham M., Blonder G. E., Klapwijk T. M., Subharmonic energy-gap structure in superconducting constrictions. Phys. Rev. B 27, 6739–6746 (1983). [Google Scholar]
- 40.Flensberg K., Hansen J. B., Octavio M., Subharmonic energy-gap structure in superconducting weak links. Phys. Rev. B 38, 8707–8711 (1988). [DOI] [PubMed] [Google Scholar]
- 41.Bagwell P. F., Suppression of the Josephson current through a narrow, mesoscopic, semiconductor channel by a single impurity. Phys. Rev. B 46, 12573–12586 (1992). [DOI] [PubMed] [Google Scholar]
- 42.Li S., Kang N., Fan D. X., Wang L. B., Huang Y. Q., Caroff P., Xu H. Q., Coherent charge transport in ballistic InSb nanowire Josephson junctions. Sci. Rep. 6, 24822 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Chen Y. Z., Trier F., Wijnands T., Green R. J., Gauquelin N., Egoavil R., Christensen D. V., Koster G., Huijben M., Bovet N., Macke S., He F., Sutarto R., Andersen N. H., Sulpizio J. A., Honig M., Prawiroatmodjo G. E. D. K., Jespersen T. S., Linderoth S., Ilani S., Verbeeck J., Van Tendeloo G., Rijnders G., Sawatzky G. A., Pryds N., Extreme mobility enhancement of two-dimensional electron gases at oxide interfaces by charge-transfer-induced modulation doping. Nat. Mater. 14, 801–806 (2015). [DOI] [PubMed] [Google Scholar]
- 44.Son J., Moetakef P., Jalan B., Bierwagen O., Wright N. J., Engel-Herbert R., Stemmer S., Epitaxial SrTiO3 films with electron mobilities exceeding 30,000 cm2 V–1 s–1. Nat. Mater. 9, 482–484 (2010). [DOI] [PubMed] [Google Scholar]
- 45.Boselli M., Scheerer G., Filippone M., Luo W., Waelchli A., Kuzmenko A. B., Gariglio S., Giamarchi T., Triscone J.-M., Electronic transport in submicrometric channels at the LaAlO3/SrTiO3 interface. Phys. Rev. B 103, 075431 (2021). [Google Scholar]
- 46.E. Persky, H. Yoon, Y. Xie, H. Y. Hwang, J. Ruhman, B. Kalisky, Electrostatic modulation of the lateral carrier density profile in field effect devices with non-linear dielectrics (2020); arXiv:2003.01756.
- 47.Fidkowski L., Jiang H.-C., Lutchyn R. M., Nayak C., Magnetic and superconducting ordering in one-dimensional nanostructures at the LaAlO3/SrTiO3 interface. Phys. Rev. B 87, 014436 (2013). [Google Scholar]
- 48.Mazziotti M. V., Scopigno N., Grilli M., Caprara S., Majorana fermions in one-dimensional structures at LaAlO3/SrTiO3 oxide interfaces. Condens. Matter 3, 37 (2018). [Google Scholar]
- 49.Perroni C. A., Cataudella V., Salluzzo M., Cuoco M., Citro R., Evolution of topological superconductivity by orbital-selective confinement in oxide nanowires. Phys. Rev. B 100, 094526 (2019). [Google Scholar]
- 50.Scheurer M. S., Schmalian J., Topological superconductivity and unconventional pairing in oxide interfaces. Nat. Commun. 6, 6005 (2015). [DOI] [PubMed] [Google Scholar]
- 51.Kanasugi S., Yanase Y., Spin-orbit-coupled ferroelectric superconductivity. Phys. Rev. B 98, 024521 (2018). [Google Scholar]
- 52.Kanasugi S., Yanase Y., Multiorbital ferroelectric superconductivity in doped SrTiO3. Phys. Rev. B 100, 094504 (2019). [Google Scholar]
- 53.Wimmer M., Akhmerov A. R., Dahlhaus J. P., Beenakker C. W. J., Quantum point contact as a probe of a topological superconductor. New J. Phys. 13, 053016 (2011). [Google Scholar]
- 54.Prada E., San-Jose P., Aguado R., Transport spectroscopy of NS nanowire junctions with Majorana fermions. Phys. Rev. B 86, 180503 (2012). [DOI] [PubMed] [Google Scholar]
- 55.Zhang H., Liu D. E., Wimmer M., Kouwenhoven L. P., Next steps of quantum transport in Majorana nanowire devices. Nat. Commun. 10, 5128 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Riwar R.-P., Houzet M., Meyer J. S., Nazarov Y. V., Multi-terminal Josephson junctions as topological matter. Nat. Commun. 7, 11167 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Xie H.-Y., Vavilov M. G., Levchenko A., Topological Andreev bands in three-terminal Josephson junctions. Phys. Rev. B 96, 161406 (2017). [Google Scholar]
- 58.Hanson R., Kouwenhoven L. P., Petta J. R., Tarucha S., Vandersypen L. M. K., Spins in few-electron quantum dots. Rev. Mod. Phys. 79, 1217–1265 (2007). [Google Scholar]
- 59.E. Mikheev, I. T. Rosen, D. Goldhaber-Gordon, Data for “Quantized critical supercurrent in SrTiO3-based quantum point contacts” (Zenodo, 2021); 10.5281/zenodo.5143210. [DOI] [PMC free article] [PubMed]
- 60.Prawiroatmodjo G. E. D. K., Trier F., Christensen D. V., Chen Y., Pryds N., Jespersen T. S., Evidence of weak superconductivity at the room-temperature grown LaAlO3/SrTiO3 interface. Phys. Rev. B 93, 184504 (2016). [Google Scholar]
- 61.Singh G., Jouan A., Benfatto L., Couëdo F., Kumar P., Dogra A., Budhani R., Caprara S., Grilli M., Lesne E., Barthélémy A., Bibes M., Feuillet-Palma C., Lesueur J., Bergeal N., Competition between electron pairing and phase coherence in superconducting interfaces. Nat. Commun. 9, 407 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Golubov A. A., Houwman E. P., Gijsbertsen J. G., Krasnov V. M., Flokstra J., Rogalla H., Kupriyanov M. Y., Proximity effect in superconductor-insulator-superconductor Josephson tunnel junctions: Theory and experiment. Phys. Rev. B 51, 1073–1089 (1995). [DOI] [PubMed] [Google Scholar]
- 63.Aminov B. A., Golubov A. A., Kupriyanov M. Y., Quasiparticle current in ballistic constrictions with finite transparencies of interfaces. Phys. Rev. B 53, 365–373 (1996). [DOI] [PubMed] [Google Scholar]
- 64.Furusaki A., Josephson current carried by Andreev levels in superconducting quantum point contacts. Superlattices Microstruct. 25, 809–818 (1999). [Google Scholar]
- 65.Wan Z., Kazakov A., Manfra M. J., Pfeiffer L. N., West K. W., Rokhinson L. P., Induced superconductivity in high-mobility two-dimensional electron gas in gallium arsenide heterostructures. Nat. Commun. 6, 7426 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Rössler C., Baer S., De Wiljes E., Ardelt P. L., Ihn T., Ensslin K., Reichl C., Wegscheider W., Transport properties of clean quantum point contacts. New J. Phys. 13, 113006 (2011). [Google Scholar]
- 67.Mittag C., Karalic M., Lei Z., Thomas C., Tuaz A., Hatke A. T., Gardner G. C., Manfra M. J., Ihn T., Ensslin K., Gate-defined quantum point contact in an InAs two-dimensional electron gas. Phys. Rev. B 100, 075422 (2019). [Google Scholar]
- 68.Mikheev E., Hoskins B. D., Strukov D. B., Stemmer S., Resistive switching and its suppression in Pt/Nb:SrTiO3 junctions. Nat. Commun. 5, 3990 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Honig M., Sulpizio J. A., Drori J., Joshua A., Zeldov E., Ilani S., Local electrostatic imaging of striped domain order in LaAlO3/SrTiO3. Nat. Mater. 12, 1112–1118 (2013). [DOI] [PubMed] [Google Scholar]
- 70.Seri S., Schultz M., Klein L., Thermally activated recovery of electrical conductivity in LaAlO3/SrTiO3. Phys. Rev.B 87, 125110 (2013). [Google Scholar]
- 71.Biscaras J., Hurand S., Feuillet-Palma C., Rastogi A., Budhani R. C., Reyren N., Lesne E., Lesueur J., Bergeal N., Limit of the electrostatic doping in two-dimensional electron gases of LaXO3(X = Al, Ti)/SrTiO3. Sci. Rep. 4, 6788 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Hurand S., Jouan A., Lesne E., Singh G., Feuillet-Palma C., Bibes M., Barthélémy A., Lesueur J., Bergeal N., Josephson-like dynamics of the superconducting LaAlO3/SrTiO3 interface. Phys. Rev. B 99, 104515 (2019). [Google Scholar]
- 73.Kouwenhoven L. P., Van Wees B. J., Harmans C. J. P. M., Williamson J. G., Van Houten H., Beenakker C. W. J., Foxon C. T., Harris J. J., Nonlinear conductance of quantum point contacts. Phys. Rev. B 39, 8040–8043 (1989). [DOI] [PubMed] [Google Scholar]
- 74.Debray P., Rahman S. M. S., Wan J., Newrock R. S., Cahay M., Ngo A. T., Ulloa S. E., Herbert S. T., Muhammad M., Johnson M., All-electric quantum point contact spin-polarizer. Nat. Nanotechnol. 4, 759–764 (2009). [DOI] [PubMed] [Google Scholar]
- 75.Matsuo S., Kamata H., Baba S., Deacon R. S., Shabani J., Palmstrøm C. J., Tarucha S., Magnetic field inducing Zeeman splitting and anomalous conductance reduction of half-integer quantized plateaus in InAs quantum wires. Phys. Rev. B 96, 201404 (2017). [Google Scholar]
- 76.Crook R., Prance J., Thomas K., Chorley S. J., Farrer I., Ritchie D. A., Pepper M., Smith C. G., Conductance quantization at a half-integer plateau in a symmetric GaAs quantum wire. Science 312, 1359–1362 (2006). [DOI] [PubMed] [Google Scholar]
- 77.Hew W. K., Thomas K. J., Pepper M., Farrer I., Anderson D., Jones G. A. C., Ritchie D. A., Spin-incoherent transport in quantum wires. Phys. Rev. Lett. 101, 036801 (2008). [DOI] [PubMed] [Google Scholar]
- 78.Scheller C. P., Liu T.-M., Barak G., Yacoby A., Pfeiffer L. N., West K. W., Zumbühl D. M., Possible evidence for helical nuclear spin order in gaas quantum wires. Phys. Rev. Lett. 112, 066801 (2014). [DOI] [PubMed] [Google Scholar]
- 79.Biercuk M. J., Mason N., Martin J., Yacoby A., Marcus C. M., Anomalous conductance quantization in carbon nanotubes. Phys. Rev. Lett. 94, 026801 (2005). [DOI] [PubMed] [Google Scholar]
- 80.Wan J., Cahay M., Debray P., Newrock R., Possible origin of the 0.5 plateau in the ballistic conductance of quantum point contacts. Phys. Rev. B 80, 155440 (2009). [Google Scholar]
- 81.Kohda M., Nakamura S., Nishihara Y., Kobayashi K., Ono T., Ohe J.-I., Tokura Y., Mineno T., Nitta J., Spin–orbit induced electronic spin separation in semiconductor nanostructures. Nat. Commun. 3, 1082 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Matveev K. A., Conductance of a quantum wire at low electron density. Phys. Rev. B 70, 245319 (2004). [DOI] [PubMed] [Google Scholar]
- 83.Fiete G. A., Colloquium: The spin-incoherent Luttinger liquid. Rev. Mod. Phys. 79, 801–820 (2007). [Google Scholar]
- 84.Gallagher P., Lee M., Williams J. R., Goldhaber-Gordon D., Gate-tunable superconducting weak link and quantum point contact spectroscopy on a strontium titanate surface. Nat. Phys. 10, 748–752 (2014). [Google Scholar]
- 85.Ron A., Dagan Y., One-dimensional quantum wire formed at the boundary between two insulating LaAlO3/SrTiO3 interfaces. Phys. Rev. Lett. 112, 136801 (2014). [DOI] [PubMed] [Google Scholar]
- 86.Swartz A. G., Inoue H., Merz T. A., Hikita Y., Raghu S., Devereaux T. P., Johnston S., Hwang H. Y., Polaronic behavior in a weak-coupling superconductor. Proc. Natl. Acad. Sci. 115, 1475–1480 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Averin D., Bardas A., AC Josephson effect in a single quantum channel. Phys. Rev. Lett. 75, 1831–1834 (1995). [DOI] [PubMed] [Google Scholar]
- 88.Golubov A. A., Kupriyanov M. Y., Il’Ichev E., The current-phase relation in Josephson junctions. Rev. Mod. Phys. 76, 411–469 (2004). [Google Scholar]
- 89.Williams J. R., Bestwick A. J., Gallagher P., Hong S. S., Cui Y., Bleich A. S., Analytis J. G., Fisher I. R., Goldhaber-Gordon D., Unconventional Josephson effect in hybrid superconductor-topological insulator devices. Phys. Rev. Lett. 109, 056803 (2012). [DOI] [PubMed] [Google Scholar]
- 90.Mizuno N., Nielsen B., Du X., Ballistic-like supercurrent in suspended graphene Josephson weak links. Nat. Commun. 4, 2716 (2013). [DOI] [PubMed] [Google Scholar]
- 91.Lee G.-H., Kim S., Jhi S.-H., Lee H.-J., Ultimately short ballistic vertical graphene Josephson junctions. Nat. Commun. 6, 6181 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Ben Shalom M., Zhu M. J., Fal’Ko V. I., Mishchenko A., Kretinin A. V., Novoselov K. S., Woods C. R., Watanabe K., Taniguchi T., Geim A. K., Prance J. R., Quantum oscillations of the critical current and high-field superconducting proximity in ballistic graphene. Nat. Phys. 12, 318–322 (2016). [Google Scholar]
- 93.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. 117, 237002 (2016). [DOI] [PubMed] [Google Scholar]
- 94.Ghatak S., Breunig O., Yang F., Wang Z., Taskin A. A., Ando Y., Anomalous Fraunhofer patterns in gated Josephson junctions based on the bulk-insulating topological insulator BiSbTeSe2. Nano Lett. 18, 5124–5131 (2018). [DOI] [PubMed] [Google Scholar]
- 95.Yu W., Stroud D., Resistance steps in underdamped Josephson-junction arrays. Phys. Rev. B 46, 14005–14009 (1992). [DOI] [PubMed] [Google Scholar]
- 96.Kuerten L., Richter C., Mohanta N., Kopp T., Kampf A., Mannhart J., Boschker H., In-gap states in superconducting LaAlO3–SrTiO3 interfaces observed by tunneling spectroscopy. Phys. Rev. B 96, 014513 (2017). [Google Scholar]
- 97.Binnig G., Baratoff A., Hoenig H., Bednorz J. G., Two-band superconductivity in Nb-doped SrTiO3. Phys. Rev. Lett. 45, 1352–1355 (1980). [Google Scholar]
- 98.Golubov A. A., Kupriyanov M. Y., Quasiparticle current in ballistic NcN′S junctions. Physica C 259, 27–35 (1996). [DOI] [PubMed] [Google Scholar]
- 99.Goldhaber-Gordon D., Shtrikman H., Mahalu D., Abusch-Magder D., Meirav U., Kastner M. A., Kondo effect in a single-electron transistor. Nature 391, 156–159 (1998). [Google Scholar]
- 100.Falco C. M., Parker W. H., Trullinger S. E., Hansma P. K., Effect of thermal noise on current-voltage characteristics of Josephson junctions. Phys. Rev. B 10, 1865–1873 (1974). [Google Scholar]
- 101.Dubouchet T., Sacépé B., Seidemann J., Shahar D., Sanquer M., Chapelier C., Collective energy gap of preformed Cooper pairs in disordered superconductors. Nat. Phys. 15, 233–236 (2019). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Supplementary Text
Sections S1 to S8
Figs. S1 to S40
References



