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

Quantized critical supercurrent in SrTiO3-based quantum point contacts

Evgeny Mikheev Department of Physics, Stanford University, Stanford, CA, 94305, USA Stanford Institute for Materials and Energy Sciences, SLAC National Accelerator Laboratory, Menlo Park, California 94025, USA    Ilan T. Rosen Department of Applied Physics, Stanford University, Stanford, CA, 94305, USA Stanford Institute for Materials and Energy Sciences, SLAC National Accelerator Laboratory, Menlo Park, California 94025, USA    David Goldhaber-Gordon Department of Physics, Stanford University, Stanford, CA, 94305, USA Stanford Institute for Materials and Energy Sciences, SLAC National Accelerator Laboratory, Menlo Park, California 94025, USA
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 towards establishing SrTiO3 as a single-material platform for mesoscopic superconducting transport experiments, that also intrinsically contains the necessary ingredients to engineer topological superconductivity.

preprint: APS/123-QED

Introduction

Conductance quantization in ballistic quantum point contacts (QPC) is a striking 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\delta G=2e^{2}/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 quantum point contact (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 ICI_{C} with a non-universal step height δIC\delta I_{C}. 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/\delta I_{C}=e\Delta/\hbar. This ideal step height is again geometry-independent and scales only with the superconducting gap Δ\Delta.

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 (for example Nb, Al). Such hybrid systems have been successfully used to demonstrate quantized critical supercurrent, but with quantization step heights far below eΔ/e\Delta/\hbar [4, 5, 6, 7, 8, 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\xi 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 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 2D electron systems (2DESs) at the surface of STO, such as LaAlO3/SrTiO3 (LAO/STO), LaTiO3/SrTiO3 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 such control is well established using back-gating through the STO substrate, top-gating through a dielectric layer, and ionic liquid gating [15, 13, 14, 16, 17, 18].

Refer to caption
Refer to caption
Figure 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. (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 VleadV_{\text{lead}} 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 VG12V_{\text{G12}} is also shown in (G) to illustrate independence of local gate voltage. The top axis shows the mapping from critical field BCB_{C} (red circles) to the coherence length ξ\xi . The estimated ξ\xi is shown in (C) for comparison with device dimensions, along with the mean free path from Hall measurements in the leads (see supplementary section S4). In (D-G), VGIL=V_{\text{GIL}}= 3 V, VBG=V_{\text{BG}}= 50 V.

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\approx 10^{13} cm-2, close to the limit of conventional dielectrics. Spatial inhomogeneity and relatively short mean free paths in such 2DESs are another challenge, leading patterned constrictions to often be dominated by tunneling through accidental quantum dots [19, 20, 21]. A QPC with normal state but not superconducting conductance quantization has recently been demonstrated in underdoped, non-superconducting 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, though it did not show direct effects of quantization. A different technique is to write conductive channels on LAO/STO with voltage-biased 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, 24, 25, 26] but not superconductivity.

In this work, we demonstrate quantized supercurrent in quantum point contacts in a split-gate geometry, based on ionic liquid gated SrTiO3. We demonstrate a quantized critical current, with tuning from zero to three ballistic modes. Step height per mode δIC\delta I_{C} is only 353-5x smaller than the canonical value eΔ/e\Delta/\hbar, 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 lift-off of metal and atomic layer deposited Hafnia (HfO2) with feature size close to 40 nm. This is distinct from the approaches taken in previous works on LAO/STO weak links [27, 28, 29, 20, 21, 22]. 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 μ\mum 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. 1A,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 5 or 6 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.

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 (VGILV_{\text{GIL}}), a back gate (VBGV_{\text{BG}}) and two split gates (VG1V_{\text{G1}} and VG2V_{\text{G2}}, denoted as VG12V_{\text{G12}} for the case VG1=VG2V_{\text{G1}}=V_{\text{G2}}). VGILV_{\text{GIL}} 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). VGILV_{\text{GIL}} 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. VGILV_{\text{GIL}} is the primary control knob for the carrier density in the leads, which can be tuned from 5×1012\approx 5\times 10^{12} to 101410^{14} cm-2 [32, 33]. The superconducting transition temperature as a function of density has a maximum near 3×10133\times 10^{13} cm-2 (see supplementary section S3). The main results presented in this paper will focus on this nearly optimally doped state, obtained by cooling the device under VGIL=V_{\text{GIL}}= +3 V. For additional data on the second constriction on the right side of the Hall bar in Fig. 1B, different devices and cool-downs with carrier density tuned across a larger range, see supplementary sections S2-S6.

Refer to caption
Figure 2: Critical current quantization. (A) DC current dependence of constriction and lead resistances at VG12=V_{\text{G12}}= 3V. (B) Constriction resistance, normalized to normal state resistance at VDC=V_{\text{DC}}= 100 μ\muV. The solid red line indicates the critical current ICI_{C}. The dashed lines indicate 1, 2 and 3 integer multiples of δIC=\delta I_{C}= 2.48 nA. (C) VG12V_{G12} dependence of ICI_{C} normalized to δIC\delta I_{C} and (D) normal state conductance GNG_{N} at VDC=V_{\text{DC}}= 100 μ\muV, with a series resistance of 800 Ω\Omega subtracted from the raw data. The shaded connection between (C) and (D) emphasizes the numerical correspondence in the observed number of ballistic modes nn. The dashed line in (C) is a fit to the saddle potential QPC model (see supplementary section S1). (E) Split- and back-gate voltage dependence of zero-bias conductance above TCT_{C}. GG has been corrected for a variable series resistance gradually increasing from 1.15 to 2.1 kΩ\Omega. Short plateaus can be seen at integer multiples of 2e2/h2e^{2}/h (n=n= 1, 2, and hints at higher multiples). Unintentional coulomb blockade levels can be seen near 0.2e2/he^{2}/h, e2/he^{2}/h and 2.5e2/he^{2}/h.

The voltage VBGV_{\text{BG}} on a back gate contacting the bottom of the SrTiO3 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=V_{\text{BG}}= +50 V to pull the electron density farther away from surface disorder (see [34] and supplementary section S4).

Fig. 1D shows the superconducting transition TCT_{C} measured by sourcing a small AC excitation through a constriction at VG12=V_{\text{G12}}= +3 V and VBG=V_{\text{BG}}= +50 V. In the following, constriction resistance and conductance will be denoted as R=dVQPC/dIACR=dV_{\text{QPC}}/dI_{\text{AC}} and G=1/RG=1/R, and the resistances of the leads as Rlead=dVlead/dIACR_{\text{lead}}=dV_{\text{lead}}/dI_{\text{AC}} (see Fig. 1B and the Methods section for more details). On both sides of the constriction, RleadR_{\text{lead}} shows a sharp transition near 350 mK. This is near the optimal value for 2D SrTiO3 [14, 17]. The measured Hall density of 3.05×10133.05\times 10^{13} cm-2 and the slight increase of TCT_{C} 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 supplementary section S4).

The constriction resistance RR also starts decreasing near the lead TCT_{C}, but its transition to zero resistance (within accuracy of our measurement) is significantly broader than that of the leads. Decreasing VG12V_{\text{G12}} suppresses both the zero resistance state and the normal state conductance, and eventually pinches off the weak link (Fig. 1E,F). At base temperature, superconductivity can also be suppressed by a perpendicular magnetic field (Fig. 1f). Using ξ2=Φ0/(2πBC)\xi^{2}=\Phi_{0}/(2\pi B_{\text{C}}) [35], with Φ0=h/2e\Phi_{0}=h/2e being the flux quantum, the critical field BC=B_{C}= 130-140 mT in the constriction yields an estimated coherence length ξ=\xi= 50 nm (43 nm in the leads). This estimate is consistent with the dirty-limit BCS superconductor picture [36, 37], in which the coherence length is set by the mean free path LMFPL_{\text{MFP}}. From Hall measurements on the leads, we extract a Hall mobility μ=\mu= 600 cm2/Vs and LMFP=L_{\text{MFP}}= 55 nm.

The shortness of these length scales illustrates the challenge of fabricating QPCs and SQPCs in SrTiO3 (see Fig.1C). Observing ballistic transport requires junction length L<LMFPL<L_{\text{MFP}}. Achieving a single-ABS junction with critical current quantization also requires short junction length: L<ξL<\xi. Though 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 VG12V_{\text{G12}} results in a staircase shape of the critical supercurrent IC(VG12)I_{C}(V_{\text{G12}}) (Fig. 2). Adopting a definition of ICI_{C} as the current at which the normal state resistance is halved, plateaus at both positive and negative integer multiples of δIC=\delta I_{C}= 2.48 nA are seen in the VG12IDCV_{\text{G12}}-I_{\text{DC}} map of constriction resistance normalized to its normal state value (Fig. 2B).

In the ballistic SQPC picture, IC/δICI_{C}/\delta I_{C} corresponds to nn, 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 VG12V_{G12} (see supplementary 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 GNG_{N}, where each fully transmitting spin-degenerate mode is expected contribute a conductance δG=2e2/h\delta G=2e^{2}/h. The number of modes inferred by dividing GNG_{N} by this increment matches that extracted from the sequence of steps in supercurrent. We also see hints of plateaus in normal state conductance near n=n= 1 and 2 (Fig. 2D). Features suggestive of normal state conductance quantization are more clearly apparent above TCT_{C} (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, as detailed further in supplementary section S5.

Refer to caption
Figure 3: SN transparency and junction length. (A) The DC current-voltage curve of the constriction at VG12=V_{\text{G12}}= 3 V and the definition of the excess current IexcI_{\text{exc}}. (B) Split gate voltage dependence of the excess and critical currents. (C) eIexcRN/ΔeI_{\text{exc}}R_{N}/\Delta, the input quantity of the SNS model in ref [38, 39], and its mapping onto SN boundary transparency τSN\tau_{\text{SN}}. (D) δIC\delta I_{C} suppression by finite transparency, and finite junction length. Comparison to the ballistic short-limit model [2] (solid black line), full calculation at L/ξ=L/\xi= 0.56 [3] (blue squares,) and with the approximate correction for arbitrary L/ξL/\xi from [40] (dashed lines). The shaded region reflects in (C) and error bars in (D) reflect the uncertainty on the gap, see supplementary section S8.

Ideally, the magnitude of steps in ICI_{C} through a constriction should scale only with the superconducting gap as

δIC=eΔ.\delta I_{C}=\frac{e\Delta}{\hbar}. (1)

This scaling is expected to hold for a short junction (L<<ξL<<\xi) with perfectly transparent SN contacts [2, 3].

For most experimental realizations of SQPC’s in hybrid metal superconductor/semiconductor devices, neither of these requirements is fully satisfied, and δIC\delta I_{C} is generally suppressed by at least an order of magnitude [4, 5, 7, 8, 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 suggests a comparable factor of 3-5. The uncertainty comes from the choice of method to extract Δ\Delta (see supplementary section S8): from TCT_{C} of the constriction [δIC/(eΔ/)=\delta I_{C}/(e\Delta/\hbar)= 2.9], from TCT_{C} of the leads [δIC/(eΔ/)=\delta I_{C}/(e\Delta/\hbar)= 4.1], or from the temperature dependence of the excess current [δIC/(eΔ/)=\delta I_{C}/(e\Delta/\hbar)= 4.8].

Analysis of the excess current IexcI_{\text{exc}} allows separating the role of imperfect SN contact transparency τSN\tau_{\text{SN}} from that of finite junction length. We define IexcI_{\text{exc}} as the zero-bias intercept of the normal-state resistance extrapolated from high VDCV_{\text{DC}} (Fig. 3A). Its evolution with VG12V_{\text{G12}} approximately tracks that of ICI_{C}. The quantity eIexcRN/ΔeI_{\text{exc}}R_{N}/\Delta can be non-linearly mapped onto τSN\tau_{\text{SN}} following the treatment of Andreev reflections in an SNS junction in [38, 39]. Over the gate voltage range with a well defined and quantized supercurrent (1.5<VG12<2.51.5<V_{\text{G12}}<2.5), we thereby extract τSN=0.750.08+0.12\tau_{\text{SN}}=0.75^{+0.12}_{-0.08}.

In the short junction limit LξL\ll\xi [2], we can predict the suppression of δIC\delta I_{C} as a function of τSN\tau_{\text{SN}} (Fig. 3D). The experimentally measured δIC\delta I_{C} is only slightly below the theoretical curve, and its full suppression can be accounted for by multiplying it by an additional factor α=\alpha= 0.7. This additional suppression can be explained by considering the finite length of the junction. An approximate theoretical description obtained in [40] is α=1/(1+L/2ξ)\alpha=1/(1+L/2\xi), which is in good agreement with calculations for the case in [3], where L=0.56ξL=0.56\xi. In this work, assuming α=0.7\alpha=0.7 yields L=0.85ξ=42L=0.85\xi=42 nm, which is close to the 40 nm lithographic width of our QPC.

Discussion

So transparency is likely to be the main driver for the reduction in δIC\delta I_{C} from its ideal value, despite being competitive with the hybrid III-V/superconductor systems, where τSN\tau_{\text{SN}} is typically estimated below 0.85 [7, 8, 41, 9] 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\tau_{\text{SN}} 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, LMFPL_{\text{MFP}} is typically less than 100 nm. However, improvements to μ>104\mu>10^{4} cm2/Vs and LMFP>L_{\text{MFP}}>μ\mum 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 [42], or forming the channel from high quality MBE-grown STO in the 3D case [43]. The fabrication route used in this work is relatively simple – based on commercially available STO crystals, avoiding epitaxial growth steps – so complex patterning or 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 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 pre-growth templates [19, 28, 29, 20], electrostatic depletion by patterned gates [27, 21, 22], or conductive channel writing by voltage biased AFM tips [23, 24, 25, 26, 44]. Our method appears particularly suited for realizations of ballistic superconducting transport, which requires maintaining a high carrier density within nanopatterned constrictions. Naturally-occurring depletion near the edges of an STO-based conducting channel [45, 44] 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 one-dimensional nanostructures [46, 47, 48]. Even an unpatterned SrTiO3 2DES may host intrinsic topological superconductivity in certain conditions due to interplay between its multi-orbital band structure, spin-orbit coupling, and ferroelectricity [49, 50, 51]. 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 [52, 53, 54]. The single-mode ballistic Josephson junction regime demonstrated here is also a requisite ingredient of theoretical proposals for realizing topological Andreev bound state spectra in multi-terminal junctions [55, 56]. Finally, this work is an important step toward realizing controlled negative-UU quantum dots [23, 20] in the classic geometry of an “island” coupled to two QPCs [57].

Materials and Methods

Fabrication is based on commercial (001)-oriented SrTiO3 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 minutes and annealed at 1000 °C for 2 hours in flowing Ar and O2 in a tube furnace.

All subsequent patterning was performed with lift-off processes using e-beam patterned PMMA 950K, 4% in anisole for the first step, 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 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. Lift-off of both HfO2 and Ti/Au layers was then performed by soaking in heated 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 lift-off 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 prior to e-beam evaporation of 10 nm Ti / 80 nm Au, followed by lift-off in acetone. The fourth patterning step is the mesa insulation, deposited by magnetron sputtering 70 nm of SiO2, followed by lift-off 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 minutes 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 VGILV_{\text{GIL}} 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 resistance per successful ohmic contact was 3-10 kΩ\Omega, which includes a 2-3 kΩ\Omega contribution from the measurement lines and built-in RF filters in the probe. Measurements were performed by voltage sourcing nominal AC and DC excitations (VACV_{\text{AC}}^{*} and VDCV_{\text{DC}}^{*}) through an adder circuit and measuring the drained current. VACV_{\text{AC}} and VDCV_{\text{DC}} refer to the measured AC and DC components of VQPCV_{\text{QPC}}, the voltage drop across the weak link.

References

  • Büttiker [1990] M. Büttiker, Physical Review B 41, 7906 (1990).
  • Beenakker and Van Houten [1991] C. W. J. Beenakker and H. Van Houten, Physical Review Letters 66, 3056 (1991).
  • Furusaki et al. [1992] A. Furusaki, H. Takayanagi,  and M. Tsukada, Physical Review B 45, 10563 (1992).
  • Takayanagi et al. [1995] H. Takayanagi, T. Akazaki,  and J. Nitta, Physical Review Letters 75, 3533 (1995).
  • Bauch et al. [2005] T. Bauch, E. Huerfeld, V. M. Krasnov, P. Delsing, H. Takayanagi,  and T. Akazaki, Physical Review B 71, 174502 (2005).
  • Xiang et al. [2006] J. Xiang, A. Vidan, M. Tinkham, R. M. Westervelt,  and C. M. Lieber, Nature Nanotechnology 1, 208 (2006).
  • Abay et al. [2013] S. Abay, D. Persson, H. Nilsson, H. Q. Xu, M. Fogelstrom, V. Shumeiko,  and P. Delsing, Nano Letters 13, 3614 (2013).
  • Irie et al. [2014] H. Irie, Y. Harada, H. Sugiyama,  and T. Akazaki, Physical Review B 89, 165415 (2014).
  • Hendrickx et al. [2019] N. W. Hendrickx, M. L. V. Tagliaferri, M. Kouwenhoven, R. Li, D. P. Franke, A. Sammak, A. Brinkman, G. Scappucci,  and M. Veldhorst, Physical Review B 99, 075435 (2019).
  • Kjærgaard et al. [2017] M. Kjærgaard, H. J. Suominen, M. Nowak, A. Akhmerov, J. Shabani, C. Palmstrøm, F. Nichele,  and C. M. Marcus, Physical Review Applied 7, 034029 (2017).
  • Lin et al. [2013] X. Lin, Z. Zhu, B. Fauqué,  and K. Behnia, Physical Review X 3, 021002 (2013).
  • Gastiasoro et al. [2020] M. N. Gastiasoro, J. Ruhman,  and R. M. Fernandes, Annals of Physics , 168107 (2020).
  • Bell et al. [2009] C. Bell, S. Harashima, Y. Kozuka, M. Kim, B. G. Kim, Y. Hikita,  and H. Y. Hwang, Physical Review Letters 103, 226802 (2009).
  • Joshua et al. [2012] A. Joshua, S. Pecker, J. Ruhman, E. Altman,  and S. Ilani, Nature Communications 3, 1 (2012).
  • Caviglia et al. [2008] A. D. Caviglia, S. Gariglio, N. Reyren, D. Jaccard, T. Schneider, M. Gabay, S. Thiel, G. Hammerl, J. Mannhart,  and J.-M. Triscone, Nature 456, 624 (2008).
  • Gallagher et al. [2015] P. Gallagher, M. Lee, T. A. Petach, S. W. Stanwyck, J. R. Williams, K. Watanabe, T. Taniguchi,  and D. Goldhaber-Gordon, Nature Communications 6, 1 (2015).
  • Chen et al. [2018] Z. Chen, A. G. Swartz, H. Yoon, H. Inoue, T. A. Merz, D. Lu, Y. Xie, H. Yuan, Y. Hikita, S. Raghu,  and H. Y. Hwang, Nature Communications 9, 1 (2018).
  • Christensen et al. [2019] D. V. Christensen, F. Trier, W. Niu, Y. Gan, Y. Zhang, T. S. Jespersen, Y. Chen,  and N. Pryds, Advanced Materials Interfaces 6, 1900772 (2019).
  • Maniv et al. [2016] E. Maniv, A. Ron, M. Goldstein, A. Palevski,  and Y. Dagan, Physical Review B 94, 045120 (2016).
  • Prawiroatmodjo et al. [2017] G. E. D. K. Prawiroatmodjo, M. Leijnse, F. Trier, Y. Chen, D. V. Christensen, M. Von Soosten, N. Pryds,  and T. S. Jespersen, Nature Communications 8, 1 (2017).
  • Thierschmann et al. [2018] H. Thierschmann, E. Mulazimoglu, N. Manca, S. Goswami, T. M. Klapwijk,  and A. D. Caviglia, Nature Communications 9, 1 (2018).
  • Jouan et al. [2020] A. Jouan, G. Singh, E. Lesne, D. C. Vaz, M. Bibes, A. Barthélémy, C. Ulysse, D. Stornaiuolo, M. Salluzzo, S. Hurand, J. Lesueur, C. Feuillet-Palma,  and N. Bergeal, Nature Electronics 3, 201 (2020).
  • Cheng et al. [2015] G. Cheng, M. Tomczyk, S. Lu, J. P. Veazey, M. Huang, P. Irvin, S. Ryu, H. Lee, C.-B. Eom, C. S. Hellberg,  and J. Levy, Nature 521, 196 (2015).
  • Cheng et al. [2016] G. Cheng, M. Tomczyk, A. B. Tacla, H. Lee, S. Lu, J. P. Veazey, M. Huang, P. Irvin, S. Ryu, C.-B. Eom, A. Daley, D. Pekker,  and J. Levy, Physical Review X 6, 041042 (2016).
  • Annadi et al. [2018] A. Annadi, G. Cheng, H. Lee, J.-W. Lee, S. Lu, A. Tylan-Tyler, M. Briggeman, M. Tomczyk, M. Huang, D. Pekker, C.-B. Eom,  and J. Irvin, P.and Levy, Nano Letters 18, 4473 (2018).
  • Briggeman et al. [2020] M. Briggeman, M. Tomczyk, B. Tian, H. Lee, J.-W. Lee, Y. He, A. Tylan-Tyler, M. Huang, C.-B. Eom, D. Pekker, R. S. K. Mong, P. Irvin,  and J. Levy, Science 367, 769 (2020).
  • Goswami et al. [2015] S. Goswami, E. Mulazimoglu, L. M. K. Vandersypen,  and A. D. Caviglia, Nano Letters 15, 2627 (2015).
  • Monteiro et al. [2017] A. M. R. V. L. Monteiro, D. J. Groenendijk, N. Manca, E. Mulazimoglu, S. Goswami, Y. Blanter, L. M. K. Vandersypen,  and A. D. Caviglia, Nano Letters 17, 715 (2017).
  • Stornaiuolo et al. [2017] D. Stornaiuolo, D. Massarotti, R. Di Capua, P. Lucignano, G. P. Pepe, M. Salluzzo,  and F. Tafuri, Physical Review B 95, 140502 (2017).
  • Balakrishnan et al. [2019] P. P. Balakrishnan, M. J. Veit, U. S. Alaan, M. T. Gray,  and Y. Suzuki, APL Materials 7, 011102 (2019).
  • Schneider et al. [2019] C. W. Schneider, M. Döbeli, C. Richter,  and T. Lippert, Physical Review Materials 3, 123401 (2019).
  • Ueno et al. [2008] K. Ueno, S. Nakamura, H. Shimotani, A. Ohtomo, N. Kimura, T. Nojima, H. Aoki, Y. Iwasa,  and M. Kawasaki, Nature Materials 7, 855 (2008).
  • Lee et al. [2011] M. Lee, J. R. Williams, S. Zhang, C. D. Frisbie,  and D. Goldhaber-Gordon, Physical Review Letters 107, 256601 (2011).
  • Chen et al. [2016] Z. Chen, H. Yuan, Y. Xie, D. Lu, H. Inoue, Y. Hikita, C. Bell,  and H. Y. Hwang, Nano Letters 16, 6130 (2016).
  • Kim et al. [2012] M. Kim, Y. Kozuka, C. Bell, Y. Hikita,  and H. Hwang, Physical Review B 86, 085121 (2012).
  • Bert et al. [2012] J. A. Bert, K. C. Nowack, B. Kalisky, H. Noad, J. R. Kirtley, C. Bell, H. K. Sato, M. Hosoda, Y. Hikita, H. Y. Hwang, et al., Physical Review B 86, 060503 (2012).
  • Collignon et al. [2017] C. Collignon, B. Fauqué, A. Cavanna, U. Gennser, D. Mailly,  and K. Behnia, Physical Review B 96, 224506 (2017).
  • Octavio et al. [1983] M. Octavio, M. Tinkham, G. E. Blonder,  and T. M. Klapwijk, Physical Review B 27, 6739 (1983).
  • Flensberg et al. [1988] K. Flensberg, J. B. Hansen,  and M. Octavio, Physical Review B 38, 8707 (1988).
  • Bagwell [1992] P. F. Bagwell, Physical Review B 46, 12573 (1992).
  • Li et al. [2016] S. Li, N. Kang, D. X. Fan, L. B. Wang, Y. Q. Huang, P. Caroff,  and H. Q. Xu, Scientific Reports 6, 24822 (2016).
  • Chen et al. [2015] Y. Z. Chen, F. Trier, T. Wijnands, R. J. Green, N. Gauquelin, R. Egoavil, D. V. Christensen, G. Koster, M. Huijben, N. Bovet, S. Macke, F. He, R. Sutarto, N. H. Andersen, J. A. Sulpizio, M. Honig, G. E. D. K. Prawiroatmodjo, T. S. Jespersen, S. Linderoth, S. Ilani, J. Verbeeck, G. Van Tendeloo, G. Rijnders, G. A. Sawatzky,  and N. Pryds, Nature Materials 14, 801 (2015).
  • Son et al. [2010] J. Son, P. Moetakef, B. Jalan, O. Bierwagen, N. J. Wright, R. Engel-Herbert,  and S. Stemmer, Nature Materials 9, 482 (2010).
  • Boselli et al. [2021] M. Boselli, G. Scheerer, M. Filippone, W. Luo, A. Waelchli, A. B. Kuzmenko, S. Gariglio, T. Giamarchi,  and J.-M. Triscone, Physical Review B 103, 075431 (2021).
  • [45] E. Persky, H. Yoon, Y. Xie, H. Y. Hwang, J. Ruhman,  and B. Kalisky, arXiv:2003.01756 .
  • Fidkowski et al. [2013] L. Fidkowski, H.-C. Jiang, R. M. Lutchyn,  and C. Nayak, Physical Review B 87, 014436 (2013).
  • Mazziotti et al. [2018] M. V. Mazziotti, N. Scopigno, M. Grilli,  and S. Caprara, Condensed Matter 3, 37 (2018).
  • Perroni et al. [2019] C. A. Perroni, V. Cataudella, M. Salluzzo, M. Cuoco,  and R. Citro, Physical Review B 100, 094526 (2019).
  • Scheurer and Schmalian [2015] M. S. Scheurer and J. Schmalian, Nature Communications 6, 1 (2015).
  • Kanasugi and Yanase [2018] S. Kanasugi and Y. Yanase, Physical Review B 98, 024521 (2018).
  • Kanasugi and Yanase [2019] S. Kanasugi and Y. Yanase, Physical Review B 100, 094504 (2019).
  • Wimmer et al. [2011] M. Wimmer, A. Akhmerov, J. Dahlhaus,  and C. Beenakker, New Journal of Physics 13, 053016 (2011).
  • Prada et al. [2012] E. Prada, P. San-Jose,  and R. Aguado, Physical Review B 86, 180503 (2012).
  • Zhang et al. [2019] H. Zhang, D. E. Liu, M. Wimmer,  and L. P. Kouwenhoven, Nature Communications 10, 1 (2019).
  • Riwar et al. [2016] R.-P. Riwar, M. Houzet, J. S. Meyer,  and Y. V. Nazarov, Nature Communications 7, 1 (2016).
  • Xie et al. [2017] H.-Y. Xie, M. G. Vavilov,  and A. Levchenko, Physical Review B 96, 161406 (2017).
  • Hanson et al. [2007] R. Hanson, L. P. Kouwenhoven, J. R. Petta, S. Tarucha,  and L. M. Vandersypen, Reviews of Modern Physics 79, 1217 (2007).
  • Prawiroatmodjo et al. [2016] G. E. D. K. Prawiroatmodjo, F. Trier, D. V. Christensen, Y. Chen, N. Pryds,  and T. S. Jespersen, Physical Review B 93, 184504 (2016).
  • Singh et al. [2018] G. Singh, A. Jouan, L. Benfatto, F. Couëdo, P. Kumar, A. Dogra, R. Budhani, S. Caprara, M. Grilli, E. Lesne, et al., Nature Communications 9, 1 (2018).
  • Golubov et al. [1995] A. A. Golubov, E. P. Houwman, J. Gijsbertsen, V. Krasnov, J. Flokstra, H. Rogalla,  and M. Y. Kupriyanov, Physical Review B 51, 1073 (1995).
  • Aminov et al. [1996] B. Aminov, A. Golubov,  and M. Y. Kupriyanov, Physical Review B 53, 365 (1996).
  • Furusaki [1999] A. Furusaki, Superlattices and microstructures 25, 809 (1999).
  • Wan et al. [2015] Z. Wan, A. Kazakov, M. J. Manfra, L. N. Pfeiffer, K. W. West,  and L. P. Rokhinson, Nature Communications 6, 7426 (2015).
  • Rössler et al. [2011] C. Rössler, S. Baer, E. De Wiljes, P. L. Ardelt, T. Ihn, K. Ensslin, C. Reichl,  and W. Wegscheider, New Journal of Physics 13, 113006 (2011).
  • Mittag et al. [2019] C. Mittag, M. Karalic, Z. Lei, C. Thomas, A. Tuaz, A. T. Hatke, G. C. Gardner, M. J. Manfra, T. Ihn,  and K. Ensslin, Physical Review B 100, 075422 (2019).
  • Mikheev et al. [2014] E. Mikheev, B. D. Hoskins, D. B. Strukov,  and S. Stemmer, Nature Communications 5, 1 (2014).
  • Blonder et al. [1982] G. Blonder, M. Tinkham,  and T. Klapwijk, Physical Review B 25, 4515 (1982).
  • Honig et al. [2013] M. Honig, J. A. Sulpizio, J. Drori, A. Joshua, E. Zeldov,  and S. Ilani, Nature Materials 12, 1112 (2013).
  • Seri et al. [2013] S. Seri, M. Schultz,  and L. Klein, Physical Review B 87, 125110 (2013).
  • Biscaras et al. [2014] J. Biscaras, S. Hurand, C. Feuillet-Palma, A. Rastogi, R. Budhani, N. Reyren, E. Lesne, J. Lesueur,  and N. Bergeal, Scientific Reports 4, 6788 (2014).
  • Hurand et al. [2019] S. Hurand, A. Jouan, E. Lesne, G. Singh, C. Feuillet-Palma, M. Bibes, A. Barthélémy, J. Lesueur,  and N. Bergeal, Physical Review B 99, 104515 (2019).
  • Kouwenhoven et al. [1989] L. Kouwenhoven, B. Van Wees, C. Harmans, J. Williamson, H. Van Houten, C. Beenakker, C. Foxon,  and J. Harris, Physical Review B 39, 8040 (1989).
  • Debray et al. [2009] P. Debray, S. M. S. Rahman, J. Wan, R. S. Newrock, M. Cahay, A. T. Ngo, S. E. Ulloa, S. T. Herbert, M. Muhammad,  and M. Johnson, Nature Nanotechnology 4, 759 (2009).
  • Matsuo et al. [2017] S. Matsuo, H. Kamata, S. Baba, R. S. Deacon, J. Shabani, C. J. Palmstrøm,  and S. Tarucha, Physical Review B 96, 201404 (2017).
  • Crook et al. [2006] R. Crook, J. Prance, K. Thomas, S. Chorley, I. Farrer, D. Ritchie, M. Pepper,  and C. Smith, Science 312, 1359 (2006).
  • Hew et al. [2008] W. Hew, K. Thomas, M. Pepper, I. Farrer, D. Anderson, G. Jones,  and D. Ritchie, Physical Review Letters 101, 036801 (2008).
  • Scheller et al. [2014] C. Scheller, T.-M. Liu, G. Barak, A. Yacoby, L. Pfeiffer, K. West,  and D. Zumbühl, Physical Review Letters 112, 066801 (2014).
  • Biercuk et al. [2005] M. Biercuk, N. Mason, J. Martin, A. Yacoby,  and C. Marcus, Physical Review Letters 94, 026801 (2005).
  • Wan et al. [2009] J. Wan, M. Cahay, P. Debray,  and R. Newrock, Physical Review B 80, 155440 (2009).
  • Kohda et al. [2012] M. Kohda, S. Nakamura, Y. Nishihara, K. Kobayashi, T. Ono, J.-I. Ohe, Y. Tokura, T. Mineno,  and J. Nitta, Nature Communications 3, 1 (2012).
  • Matveev [2004] K. A. Matveev, Physical Review B 70, 245319 (2004).
  • Fiete [2007] G. A. Fiete, Reviews of Modern Physics 79, 801 (2007).
  • Gallagher et al. [2014] P. Gallagher, M. Lee, J. R. Williams,  and D. Goldhaber-Gordon, Nature Physics 10, 748 (2014).
  • Ron and Dagan [2014] A. Ron and Y. Dagan, Physical Review Letters 112, 136801 (2014).
  • Swartz et al. [2018] A. G. Swartz, H. Inoue, T. A. Merz, Y. Hikita, S. Raghu, T. P. Devereaux, S. Johnston,  and H. Y. Hwang, Proceedings of the National Academy of Sciences 115, 1475 (2018).
  • Averin and Bardas [1995] D. Averin and A. Bardas, Physical Review Letters 75, 1831 (1995).
  • Golubov et al. [2004] A. A. Golubov, M. Y. Kupriyanov,  and E. Il’Ichev, Reviews of Modern Physics 76, 411 (2004).
  • Williams et al. [2012] J. Williams, A. Bestwick, P. Gallagher, S. S. Hong, Y. Cui, A. S. Bleich, J. Analytis, I. Fisher,  and D. Goldhaber-Gordon, Physical Review Letters 109, 056803 (2012).
  • Mizuno et al. [2013] N. Mizuno, B. Nielsen,  and X. Du, Nature Communications 4, 1 (2013).
  • Lee et al. [2015] G.-H. Lee, S. Kim, S.-H. Jhi,  and H.-J. Lee, Nature Communications 6, 1 (2015).
  • Ben Shalom et al. [2016] M. Ben Shalom, M. Zhu, V. Fal’Ko, A. Mishchenko, A. Kretinin, K. Novoselov, C. Woods, K. Watanabe, T. Taniguchi, A. Geim, et al., Nature Physics 12, 318 (2016).
  • Borzenets et al. [2016] I. Borzenets, F. Amet, C. Ke, A. Draelos, M. Wei, A. Seredinski, K. Watanabe, T. Taniguchi, Y. Bomze, M. Yamamoto, et al., Physical Review Letters 117, 237002 (2016).
  • Ghatak et al. [2018] S. Ghatak, O. Breunig, F. Yang, Z. Wang, A. A. Taskin,  and Y. Ando, Nano Letters 18, 5124 (2018).
  • Yu and Stroud [1992] W. Yu and D. Stroud, Physical Review B 46, 14005 (1992).
  • Kuerten et al. [2017] L. Kuerten, C. Richter, N. Mohanta, T. Kopp, A. Kampf, J. Mannhart,  and H. Boschker, Physical Review B 96, 014513 (2017).
  • Binnig et al. [1980] G. Binnig, A. Baratoff, H. Hoenig,  and J. Bednorz, Physical Review Letters 45, 1352 (1980).
  • Golubov and Kupriyanov [1996] A. Golubov and M. Y. Kupriyanov, Physica C 259, 27 (1996).
  • Goldhaber-Gordon et al. [1998] D. Goldhaber-Gordon, H. Shtrikman, D. Mahalu, D. Abusch-Magder, U. Meirav,  and M. Kastner, Nature 391, 156 (1998).
  • Falco et al. [1974] C. M. Falco, W. Parker, S. Trullinger,  and P. K. Hansma, Physical Review B 10, 1865 (1974).
  • Dubouchet et al. [2019] T. Dubouchet, B. Sacépé, J. Seidemann, D. Shahar, M. Sanquer,  and C. Chapelier, Nature Physics 15, 233 (2019).

Acknowledgments

We acknowledge Marc Kastner, Malcolm Beasley and Eli Fox for helpful discussions. We acknowledge Richard Tiberio and Michelle Rincon for help with device fabrication.

Funding: This work was supported by the Air Force Office of Scientific Research through grant no. FA9550-16-1-0126. E. M. was supported by the Nano- and Quantum Science and Engineering Postdoctoral Fellowship at Stanford University. I. T. R. 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. Measurement infrastructure was funded in part by the Gordon and Betty Moore Foundation’s EPiQS Initiative through grant GBMF3429 and grant 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.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. Additional data related to this paper may be requested from the authors.

Supplementary material for “Quantized critical supercurrent in SrTiO3-based quantum point contacts”

This PDF file includes:

Supplementary Text

Figs. S1 to S40

S1 Theoretical framework

S1.1 The nature of the SNS weak link

SrTiO3 can be described as a semiconducting superconductor: as a function of carrier density, its ground state evolves from an insulator to a normal metal to a superconductor. A schematic phase diagram is shown in Fig. S1A, roughly summarizing a very large body of work on substitutionally-doped SrTiO3, LaAlO3/SrTiO3, LaTiO3/SrTiO3, and ionic liquid-gated SrTiO3, see e.g. [11, 15, 14, 16, 18].

Fig. S1B illustrates our understanding of how the split gates create a weak link between superconducting reservoirs by locally depleting the carrier density. At higher positive split gate voltage, an SNS junction is formed. At lower gate voltage, the depletion region extends further into the constriction, eventually pinching it off. In this simplistic picture, the underdoped side of the phase diagram is reproduced as a function of distance from the split gate. The resulting weak link is then likely to not have sharp SN boundaries, but instead gradual transitions from near-optimal TcT_{c} to weak superconductivity and then to normal metal (Fig. S1C).

A complete modelling of such a system is a difficult task, in particular due to the non-linear dielectric constant of SrTiO3 [21], and complex interplay between microscopic pairing and macroscopic coherence in the underdoped regime [58, 59, 17].

For simplicity, we choose to model our devices as SNS junctions, with normal region length LL and abrupt SN interfaces with an effective transparency τSN\tau_{\text{SN}} (Fig. S1D). In some cases, a single transparency τ\tau is defined for the entire junction (Fig. S1E); we approximate the relation between the two as τ=τSN2\tau=\tau_{\text{SN}}^{2}, assuming no scattering within the N region.

A potential added complexity that will need to be addressed in follow up work is whether an SS’NS’S description (Fig. S1F) is more appropriate for such junctions than SNS. S’ is either a superconducting region with the order parameter reduced by depletion, or a normal metal with a pairing gap induced by proximity effect [60]. In either case, the S’ pairing scale becomes distinct from the S scale measured in the leads. In the case of a short S’ region, both the S and S’ scales are relevant for Josephson and tunneling transport [60, 61].

Refer to caption
Figure S1: (A) Schematic phase diagram of SrTiO3 as a function of doping. (B) Schematic top view of a normal state constriction with superconducting leads. (C-F): different 1D model representations of the constriction: (C) SNS with diffuse SN boundaries due to gradual change in carrier density, (D) idealized SNS with sharp SN boundaries with transparency τSN\tau_{\text{SN}}, (E) same as (D) but with an alternate definition of junction transparency τ=τSN2\tau=\tau_{\text{SN}}^{2}, (F) SS’NS’S constriction, where S’ is a superconducting region with a reduced order parameter in comparison to the S region.

S1.2 Critical current of a short junction with finite contact transparency

Following [2, 62], the simplest picture of a ballistic SQPC is given by a one-dimensional, “short-limit” SNS model. The pair potential is taken to be a step function: bulk-like in the S region, and zero in the N region. Imperfect junction transparency is modeled by introducing scattering in the N region from a δ\delta-function potential

U(x)=VBδ(x),U(x)=V_{B}\delta(x), (S1)

which is traditionally renormalized into a dimensionless parameter Z=mVB/2kFZ=mV_{B}/\hbar^{2}k_{F}, with kFk_{F} being the Fermi wavevector. The equivalent transmission probability of the N region is τ=1/(1+Z2)\tau=1/(1+Z^{2}). In the short limit, where the length of the junction is much shorter than the superconducting coherence length (LξL\ll\xi), the phase dependence of the Andreev bound state (ABS) spectrum is given by [2]

EB(ϕ)=Δ1τsin2(ϕ/2),E_{B}(\phi)=\Delta\sqrt{1-\tau\sin^{2}(\phi/2)}, (S2)

and the current-phase relationship for a single ballistic mode is

I1(ϕ)=eΔsin(ϕ)2τcos2(ϕ/2)1+τ1tanh(EB2kBT).I_{1}(\phi)=\frac{e\Delta}{\hbar}\cdot\frac{\sin(\phi)}{2}\cdot\sqrt{\frac{\tau}{\cos^{2}(\phi/2)-1+\tau^{-1}}}\cdot\tanh\left(\frac{E_{B}}{2k_{B}T}\right). (S3)

The critical current of one mode is δIc=max(I1(ϕ))\delta I_{c}=\text{max}(I_{1}(\phi)). Fig. S2A-C shows the evolution of the ABS spectrum, the current-phase relationship and δIc\delta I_{c} with transparency.

Refer to caption
Figure S2: short-limit SNS junction model: (A) Andreev bound state spectrum at different SN transparency levels, (B) current-phase relationship, also at different SN transparency levels, (C) critical current carried by a single ballistic mode, shown as a function of τSN\tau_{\text{SN}} (τ=τSN2\tau=\tau_{\text{SN}}^{2}) in the short junction limit, and (D) junction length dependence in perfect transparency limit. Blue symbols in (C) and (D) are numerical results from [3].

S1.3 Critical current suppression with junction length

For a long SNS junction [3, 40], the Josephson current is carried by multiple ABS. With the length LL referring to the junction size along the current direction (Fig. S1D), the number of ABS is approximately L/ξL/\xi. In the absence of scattering, the maximum supercurrent decreases as 1/L1/L. It was shown by Bagwell [40] that the crossover between the short (LξL\ll\xi) and long (LξL\gg\xi) limits can be interpolated as

δIc=eΔ11+L2ξ.\delta I_{c}=\frac{e\Delta}{\hbar}\cdot\frac{1}{1+\frac{L}{2\xi}}. (S4)

To treat the case of finite length and transparency, we adopt the approximation from [63] that the correction factor α\alpha for finite length is a multiplier for the current-phase relationship derived above for an SNS with finite transparency

α=11+L2ξ,\alpha=\frac{1}{1+\frac{L}{2\xi}}, (S5)
δIc=max(I1(ϕ))α.\delta I_{c}=\text{max}(I_{1}(\phi))\cdot\alpha. (S6)

As a cross-check, Fig. S2D illustrates that the equation S4 is in agreement with a different calculation by Furusaki et al. [3]. We also verify that equation S6 closely agrees with the calculation in [3] for the case of finite transparency at L/ξ=0.56L/\xi=0.56 in Fig. S2C. We can see that a junction length smaller than but on the order of coherence length modestly suppresses the supercurrent.

S1.4 Saddle potential constriction model

The conductance plateau structure of a QPC is generally modeled by assuming a saddle potential profile [1]

VQPC(x,y)=VQPC(0,0)mωx2x22+mωy2y22,V_{\text{QPC}}(x,y)=V_{\text{QPC}}(0,0)-\frac{m\omega_{x}^{2}x^{2}}{2}+\frac{m\omega_{y}^{2}y^{2}}{2}, (S7)

where xx and yy axes are parallel and orthogonal to the current flow, VQPC(0,0)V_{\text{QPC}}(0,0) is the potential at the center of the constriction, ωx\omega_{x} and ωy\omega_{y} describe the confining potential curvature. Transverse confinement discretizes the available states

En=VQPC(0,0)+(n+12)ωy,E_{n}=V_{\text{QPC}}(0,0)+\left(n+\frac{1}{2}\right)\hbar\omega_{y}, (S8)

with n=n= 0, 1, 2, 3, … The channels with transverse confinement energy below the Fermi energy are open, and above it are closed. The crossover between open and closed states is described by the transmission coefficient of an individual mode

Tn(E)=11+exp(2π(EEn)ωx).T_{n}(E)=\frac{1}{1+\exp\left(\frac{-2\pi\left(E-E_{n}\right)}{\hbar\omega_{x}}\right)}. (S9)

The split gate voltage tuning of the Fermi energy above the pinch-off voltage VPV_{P} is generally described by a linear lever arm CLAC_{\text{LA}}

Tn(VG12)=11+exp(VPCLAVG122π(n+12)ωyωx).T_{n}(V_{\text{G12}})=\frac{1}{1+\exp\left(V_{P}-C_{\text{LA}}V_{\text{G12}}-2\pi\left(n+\frac{1}{2}\right)\frac{\omega_{y}}{\omega_{x}}\right)}. (S10)

Normal state conductance is then the sum across all available modes, each carrying a conductance quantum of 2e2/h2e^{2}/h at full transmission

GN(VG12)=2e2hnTn(VG12),G_{N}(V_{\text{G12}})=\frac{2e^{2}}{h}\cdot\sum_{n}T_{n}(V_{\text{G12}}), (S11)

where the factor of 2 is from spin degeneracy.

In the extension of the saddle potential model to a ballistic SNS constriction, the critical current follows the same quantization pattern as GNG_{N} [5, 8], but with a non-universal step height δIc\delta I_{c}

Ic(VG12)=δIcnTn(VG12).I_{c}(V_{\text{G12}})=\delta I_{c}\cdot\sum_{n}T_{n}(V_{\text{G12}}). (S12)

For fitting a step structure in Ic/δIcI_{c}/\delta I_{c} or GNG_{N} vs VG12V_{\text{G12}}, the saddle potential model has three adjustable parameters: VPV_{P} and CLAC_{\text{LA}} are used for position and rescaling on the VG12V_{\text{G12}} axis, and the confinement strength ratio ωy/ωx\omega_{y}/\omega_{x} for plateau sharpness. At high ωy/ωx\omega_{y}/\omega_{x} (long and narrow QPC), the discrete channel states are well separated in energy space, resulting in sharp, well defined plateaus [64].

Fig. S3 reproduces Fig.3C and 3D in the main text, but with the saddle potential description of both GNG_{N} and IcI_{c} by equations S11 and S12. The purple dashed line in both figures is a fit to Ic/δIcI_{c}/\delta I_{c} using eq. S12 using values of VG12<V_{\text{G12}}< 2.6 V, giving ωy/ωx=2.07\omega_{y}/\omega_{x}=2.07. The fit is very good for IcI_{c} up to n=3n=3 (Fig. S3A), but applying the same parameters to eq. S11 only approximately describes GNG_{N} (Fig. S3B). The orange line in Fig. S3B is a fit to GNG_{N} using values of VG12<V_{\text{G12}}< 2.4 V, giving ωy/ωx=0.99\omega_{y}/\omega_{x}=0.99 and a good description of GNG_{N} up to n2.5n\approx 2.5.

Refer to caption
Figure S3: Reproduction of Fig. 2C, D in the main text with a Buttiker model fit to GNG_{N}. (A) red circles is IcI_{c} normalized to its step height for device 1A at VGIL=V_{\text{GIL}}= 3 V, VBG=V_{\text{BG}}= 50 V, T=T= 45 mK. (B) solid line is the normal state conductance at VDC=V_{\text{DC}}= 100 μ\muV, with a series resistance of 800 Ω\Omega subtracted from the raw data. The purple dashed lines in (A) is a fit of IcI_{c} to eq. S12 with ωy/ωx=2.07\omega_{y}/\omega_{x}=2.07, and an arbitrary lever arm. The purple dashed line in (B) uses the same parameters to describe GNG_{N} with eq. S11. The orange dashed line in (B) is a fit of GNG_{N} to eq. S11 in the n=02.5n=0-2.5 region with ωy/ωx=0.99\omega_{y}/\omega_{x}=0.99.

We do not have a complete explanation of why the crossover between modes is sharper in IcI_{c} then in GNG_{N}. The description above by a difference in confinement ratios suggests that the shape of the constriction is not the same in the normal and superconducting states (that it is effectively narrower and/or longer in the latter). An alternate explanation could involve partial breakdown of the assumption that the the constriction is adiabatically coupled to the leads [1, 65], and a different extent of this breakdown in the normal and superconducting states.

Another unusual aspect of Fig. S3 is the occurrence of VPV_{P} at positive VG12V_{\text{G12}}. Generically, pinch off is caused by depletion at negative split gate voltage [64]. As documented further in section S5, VPV_{P} in our devices is a strong function of doping in the leads. It quickly shifts from positive at low electron density to negative at high density. A positive VPV_{P} is consistent with a built-in depletion field at VG12=V_{\text{G12}}= 0, for example from trapped charge at the gate metal/oxide interface [66]. A nominally positive VG12V_{\text{G12}} can thus still correspond to depletion around the split gates.

S1.5 OBTK model

In this section, we briefly summarize the Octavio-Blonder-Tinkham-Klapwijk model [38, 39] of an SNS constriction. It assumes a one-dimensional SNS weak link with two scattering barriers at each SN interface, with transparency τSN\tau_{\text{SN}}. Its relationship to the equivalent SN barrier height is τSN=1/(1+ZSN2)\tau_{\text{SN}}=1/(1+Z_{\text{SN}}^{2}).

The current across the junction is calculated by integrating the distributions of right and left moving moving electrons (ff_{\rightarrow} and ff_{\leftarrow}) in the energy space

I=1eRN+𝑑E(f(E)f(E)),I=\frac{1}{eR_{N}}\int_{-\infty}^{+\infty}dE(f_{\rightarrow}(E)-f_{\leftarrow}(E)), (S13)
f(E)=A(E)f(EeV)+B(E)(1f(EeV))+T(E)f0(E),f_{\rightarrow}(E)=A(E)f_{\rightarrow}(E-eV)+B(E)(1-f_{\rightarrow}(-E-eV))+T(E)f_{0}(E), (S14)
f(E)=f(EeV),f_{\rightarrow}(E)=f_{\leftarrow}(-E-eV), (S15)

where f0f_{0} is the standard Fermi function. At the SN interfaces, AA is the Andreev reflection probablity, BB is the ordinary reflection probablity, T=1ABT=1-A-B is the transmission probablity [67]. For E<ΔE<\Delta

A(E)=Δ2E2+(Δ2E2)(1+2ZSN2)2,A(E)=\frac{\Delta^{2}}{E^{2}+(\Delta^{2}-E^{2})(1+2Z_{\text{SN}}^{2})^{2}}, (S16)
B(E)=1A(E).B(E)=1-A(E). (S17)

for E>ΔE>\Delta:

A(E)=u02v02γ2,A(E)=\frac{u^{2}_{0}v^{2}_{0}}{\gamma^{2}}, (S18)
B(E)=(u02v02)Z2(1+ZSN2)γ2.B(E)=\frac{(u^{2}_{0}-v^{2}_{0})Z^{2}(1+Z_{\text{SN}}^{2})}{\gamma^{2}}. (S19)

The superconducting density of states NSN_{S} enters the above equations as

NS=1u02v02,N_{S}=\frac{1}{u_{0}^{2}-v_{0}^{2}}, (S20)
u02=1v02=12(1+(E2Δ2Δ2)1/2),u_{0}^{2}=1-v_{0}^{2}=\frac{1}{2}\left(1+\left(\frac{E^{2}-\Delta^{2}}{\Delta^{2}}\right)^{1/2}\right), (S21)
γ=u02+ZSN2(u02v02).\gamma=u_{0}^{2}+Z_{\text{SN}}^{2}(u_{0}^{2}-v_{0}^{2}). (S22)

The excess current IexcI_{\text{exc}} for a particular value of τSN\tau_{\text{SN}} can be found by calculating the I(V)I(V) curve with eq. S13, and linearly extrapolating from VΔ/eV\gg\Delta/e to V=0V=0. The reverse mapping from dimensionless quantity eIexcRN/ΔeI_{\text{exc}}R_{N}/\Delta to transparency is shown in Fig. S4. This curve was used to estimate τSN\tau_{\text{SN}} from experimental IDC(VDC)I_{\text{DC}}(V_{\text{DC}}) curves (Fig. 3 in the main text and additional data shown in section S6).

Refer to caption
Figure S4: Mapping between excess current and transparency in the SNS model [38, 39]: (A) SN and SNS transparency, (B) equivalent barrier heights.

S2 Additional devices and fabrication notes

Sample 1: fabrication outlined in the methods section of the main text

  • Device 1A: 40 nm gap between local gates. Data for a cooldown at VGIL=V_{\text{GIL}}= +3 V are discussed in the main text and throughout the supplementary material.

  • Device 1B: 60 nm gap between local gates. Data for a cooldown at VGIL=V_{\text{GIL}}= +3 V are discussed in section S6.

Refer to caption
Figure S5: Optical images of (A) sample 2 and (B) sample 3, taken at the end of the fabrication process. Optically visible gaps near the constriction are caused by the progressive narrowing of the gate tips.
Refer to caption
Figure S6: AFM image of atomic terrace steps on the surface of sample 1.

Sample 2: same overall fabrication method as sample 1. Different design with three devices on the Hall bar with minor distinctions. The SiO2 mesa insulator thickness was 100 nm and the final anneal before ionic liquid deposition was 1 minute at 180 °C. Optical image shown in Fig. S5A

  • Device 2A: 40 nm gap between local gates. Data for cooldowns at VGIL=V_{\text{GIL}}= +3, +3.5 and +3.7 V are discussed in section S5.

  • Device 2B: 60 nm gap between local gates. Data for cooldowns at VGIL=V_{\text{GIL}}= +3, +3.5 and +3.7 V are discussed in section S5.

  • Device 2C: 100 nm gap between local gates, one of the gates was electrically shorted to the STO channel.

Sample 3: same overall fabrication method as sample 1, but sputtering of SiO2 as mesa insulator in the last fabrication step is replaced with atomic layer deposition of thick HfO2. The lift-off procedure was similar to the local gate patterning in step 1, but with 200 cycles of atomic layer deposition. The final anneal was 20 minutes at 115 °C. Optical image shown in Fig. S5B. We found this alternative approach to depositing mesa insulators to be viable but detrimental to ohmic contacts, which suffered from poor yield and were only functional at relatively high carrier densities.

  • Device 3A: 40 nm gap between local gates. Data for a cooldown at VGIL=V_{\text{GIL}}= +3 V are discussed in section S6.

  • Device 3B: 60 nm gap between local gates. Data for a cooldown at VGIL=V_{\text{GIL}}= +3 V are discussed in section S6.

Fig. S6 shows an atomic force microscope (AFM) image of SrTiO3 surface. Atomic terrace steps due to surface miscut are clearly visible. The image was taken on sample 1 after surface preparation with a deionized water soak, followed by an anneal at 1000 °C in an Ar/O2 atmosphere. It was taken prior to fabrication of devices 1A and 1B on this sample.

Refer to caption
Figure S7: Optical images of devices 1A and 1B taken at intermediate stages of the fabrication process: (A) gate lift-off, (B) gate contact lift-off, (C) ohmic contact lift-off, (D) mesa insulation lift-off.
Refer to caption
Figure S8: Optical image of sample 1, taken after the measurements in a dilution fridge, with the ionic liquid covering devices 1A, 1B, and the large coplanar gate.

Intermediate stages of the fabrication process are illustrated in Fig. S7. The presented optical images are centered around the Hall bar-style mesa that is defined in the final step by depositing SiO2 insulation. A larger view of the device is shown in Fig. S8, which includes the ionic liquid and the large coplanar gate used to drive an insulator-to-metal transition on the exposed SrTiO3 surface.

S3 Tuning superconductivity in the leads with ionic liquid gating

Fig. S9 illustrates the initial steps of the device measurement. The initial ramping of the coplanar gate (VGILV_{\text{GIL}}) is performed at room temperature and is monitored by a two terminal-like measurement. A DC voltage of 1 mV is sourced to one chosen ohmic contact and all remaining contacts on the Hall bar are grounded. The resulting current I2TI_{\text{2T}} typically becomes measurable near VGIL=V_{\text{GIL}}= 1 - 2 V and quickly increases by several orders of magnitude. The corresponding resistance R2TR_{\text{2T}} typically saturates around 100 kΩ\Omega. It includes contributions from diffusive scattering in the SrTiO3 channel, contact and line resistances. The dominant contributions at room temperature are channel and contact resistances.

If the carrier density induced in the channel is sufficiently large to make it metallic (N>51012N>5\cdot 10^{12} cm-2), the measured resistance quickly decreases upon cooling. R2TR_{\text{2T}} becomes dominated by contact and line resistances at low temperatures. During low temperature measurements, VGILV_{\text{GIL}} is kept at a fixed value chosen at the start of the cooldown. Below 220 K, the ionic liquid is frozen and does not respond to adjustments of VGILV_{\text{GIL}}. To re-adjust the carrier density in the Hall bar by changing VGILV_{\text{GIL}}, the device needs to be thermally cycled above that point, as was done for device 2.

Fig. S10 summarizes the various cooldowns performed on samples 1-3. While the trend of Hall density with VGILV_{\text{GIL}} shows significant scatter between samples, it is monotonic for successive cooldowns on the same sample (sample 2). The superconducting transition points shown in Fig. S10B are defined as the midpoint of the resistance drop measured in the leads or across the gated constriction. For constrictions, transition temperatures were extracted in “open” state at large positive local gate voltage. The results are consistent with a dome-shaped superconducting phase, with an onset of TcT_{c} at carrier densities above 110131\cdot 10^{13} cm-2 and a peak near 310133\cdot 10^{13} cm-2.

This work focuses on the underdoped and near optimal regimes, where the carrier density is low enough that local constrictions are electrostatically tunable by HfO2 dielectric gates.

A likely source of uncertainty in extracted carrier densities is multiple band occupancy resulting in non-linear Hall effect. Strong non-linearity of transverse resistance with magnetic field of is commonly observed at high carrier density in ionic liquid-gated SrTiO3 and LAO/STO [16, 14]. However, only weak deviation from linearity was observed up to 14 T for device 1 and up to 3 T for device 3.

Refer to caption
Figure S9: (A) Initial sweep of the large coplanar gate VGILV_{\text{GIL}} measured for device 3 at room temperature. Channel current I2TI_{\text{2T}} and coplanar gate leakage IGILI_{\text{GIL}} are shown. (B) Same sweep presented as channel resistance R2TR_{\text{2T}}. (C) R2TR_{\text{2T}} measured during cooldown at fixed VGIL=V_{\text{GIL}}= 3 V.
Refer to caption
Figure S10: Summary of different devices and cooldowns. (A) low-field Hall density measured in the leads at base temperature. For sample 1, the effect of applying a back voltage VBG=V_{\text{BG}}= +50 V is indicated. VBG=V_{\text{BG}}= 0 V for other points. (B) Superconducting transition as indicated by the midpoint of resistance drop. Symbol markers indicate measurement in the leads vs across the point contact. Colors correspond to different samples.

S4 Additional characterization of the local and back gates

Fig. S11 illustrates the extraction of the Ginzburg-Landau coherence length estimate ξ\xi from the critical field BcB_{c} of the constriction and the leads with

ξ2=Φ02πBC,\xi^{2}=\frac{\Phi_{0}}{2\pi B_{\text{C}}}, (S23)

with Φ0\Phi_{0} being the flux quantum. Using the criterion Bc=B(R/RN=0.5)B_{c}=B(R/R_{N}=0.5), with RNR_{N} being the normal state resistance at high BB, ξ=\xi= 43 nm in the leads. In the constriction, ξ\xi decreases from 50 to 48 nm with increasing VG12V_{\text{G12}}. If one chooses a lower criterion Bc=B(R/RN=0.25)B_{c}=B(R/R_{N}=0.25), the slightly modified estimates are: 45 nm in the leads and 53-60 nm in the constriction. As a function of VG12V_{\text{G12}}, the BcB_{c} measurement in the leads remains unchanged until the constriction is closed near VG12=V_{\text{G12}}= 0.8 V and there is no sourcing current. In the constriction measurement, the supercurrent becomes intermittent below VG12=V_{\text{G12}}= 1.5 V due to the resonant nature of barrier transmission in this regime. The spikes in BcB_{c} and ξ\xi extracted at low VG12V_{\text{G12}} are thus not necessarily reflective of any actual change in the superconducting order near the constriction.

Refer to caption
Figure S11: Normalized resistance of (A) the constriction and (B) the lead as a function of magnetic field and local gate voltage in device 1A at VBG=V_{\text{BG}}= 50 V, VGIL=V_{\text{GIL}}= 3 V, T=T= 44 mK. Normal state resistance RNR_{N} is taken B=B= 500 mT. Black lines indicate the critical field BcB_{c}, defined as the midpoint of the resistance drop. Critical fields extracted in (A) and (B) are re-plotted in (C). (D) Superconducting coherence length estimate extracted from BcB_{c} using equation S23.
Refer to caption
Figure S12: (A) Normalized resistance of the constriction as a function of temperature and local gate voltage in device 1A at VBG=V_{\text{BG}}= 50 V, VGIL=V_{\text{GIL}}= 3 V. (B) TcT_{c} extracted in (A) and from concurrent measurement of lead resistance. (C) Lead resistance as a function of DC current and local gate voltage at VBG=V_{\text{BG}}= 50 V, VGIL=V_{\text{GIL}}= 3 V, T=T= 44 mK. (D) Comparison between the critical current in the leads and in the constriction.

Fig. S12A shows the extraction of TcT_{c} from the constriction resistance measurement, where similar intermittency is seen at low VG12V_{\text{G12}}. While the TcT_{c} of the constriction is tuned between 200 and 275 mK by the local gate, TcT_{c} in the leads remains constant at 350 mK.

Interestingly, the only measurement in which the leads are sensitive to VG12V_{\text{G12}} is their critical current. Fig. S12C shows the lead resistance up to high DC current at base temperature, where a decreased IcI_{c} is clearly seen at low VG12V_{\text{G12}}. This suggests that the local gate can also have a very long range effect on the leads, with the closest voltage probe being 5 microns away. Such an effect is plausible given the highly non-linear dielectric constant of SrTiO3, and a presumably non-uniform carrier density profile in the Hall bar, with increased depletion at the edges [45]. Nevertheless, Fig. S12D illustrates that IcI_{c} in the leads remains two orders of magnitude above IcI_{c} in the constriction.

Refer to caption
Figure S13: (A) Carrier density dependence on back gate voltage, extracted from Hall effect measured in the low and high field limits for device 1A at VG12=V_{\text{G12}}= +3 V, VGIL=V_{\text{GIL}}= 3 V, T=T= 45 mK. (B) Hall mobility in the leads, calculated from the high-field limit value of NN. (C) Constriction conductance traces with local gate voltage, at fixed VBG=V_{\text{BG}}= 0-50 V, taken in the normal state at T=T= 866 mK. Data is presented without offset shifting and series resistance correction.
Refer to caption
Figure S14: Back gate tuning of TcT_{c} in device 1A at VG12=V_{\text{G12}}= 3 V, VGIL=V_{\text{GIL}}= 3 V . (A) Constriction and (B) lead resistance as a function of temperature and back gate voltage. Solid black lines in (A,B) indicate TcT_{c}. TcT_{c} values extracted in (A) and (B) are re-plotted in (C). (D) Constriction and lead resistance measured above TcT_{c}.
Refer to caption
Figure S15: Back gate tuning of critical current IcI_{c} in device 1A at VG12=V_{\text{G12}}= 3 V, VGIL=V_{\text{GIL}}= 3 V, T=T= 44 mK. (A) Constriction and (B) lead resistance as a function of DC source current and back gate voltage. Black lines indicate TcT_{c}. (C) Critical current values for the resistance drop in (A) and (B). (D) Normal state constriction and lead resistance taken at high DC current.

Back gate voltage VBGV_{\text{BG}} is an additional tuning knob for our device. VBGV_{\text{BG}} is applied between the device and the bottom surface of a 0.5 mm thick SrTiO3 crystal, connected to the bottom of a chip carrier with silver paint. Such gates can have an appreciable capacitance due to the quantum paraelectric nature of SrTiO3, resulting in a dielectric constant of order 104 in the low temperature limit [68, 34, 45].

Unlike VG12V_{\text{G12}}, VBGV_{\text{BG}} affects both the constriction and the leads. Similarly to many previous experiments on LAO/STO 2DESs [13, 34], the main effect of VBGV_{\text{BG}} is to change the carrier mobility μ=1/(eNRlead)\mu=1/(eNR_{\text{lead}}), rather than to change the density NN. The back gate effect in Fig. S13 is 3% on NN and 10% on μ\mu. This is consistent with the understanding that positive VBGV_{\text{BG}} pulls the 2DES away from disorder scattering at the surface, increasing the mobility [34]. This work has thus mainly focused on the VBGV_{\text{BG}} = 50 V state, which offers the highest 2DES mobility. In our devices, the capability to deplete using VBGV_{\text{BG}} is limited due to rapid damage suffered by ohmic contacts upon applying negative VBGV_{\text{BG}}. Such damage is largely reversible upon thermally cycling the device to near room temperatures, suggesting a charge trapping mechanism similar to the one documented in [69, 70]. Similar contact damage can occur at negative VG12V_{\text{G12}}, particularly at low VBGV_{\text{BG}}. This is likely a consequence of capacitive cross-coupling between the gates.

Fig. S14 shows the modulation of the superconducting transition by VBGV_{\text{BG}}. The lead TcT_{c} is decreased from 370 mK to 350 mK by increasing VBGV_{\text{BG}}. This is consistent with a near-optimal, slightly overdoped position on the the superconducting dome.

The back gate effect on the constriction resistance is proportionally much larger than for the leads. This is shown in one dimensional sweeps of VBGV_{\text{BG}} in the normal state at above TcT_{c} in Fig. S14D and at high DC bias in Fig. S15D. A comparison of VG12V_{\text{G12}} sweeps at different VBGV_{\text{BG}} (Fig. S13D) show that the most obvious effect is a horizontal shift in VG12V_{\text{G12}}, suggesting a cross-coupling effect. Beyond the horizontal shift, there is a vertical shift of most non-monotonic features in GG, such as the short plateaus near 2e2/h2e^{2}/h and 4e2/h4e^{2}/h that are most clearly visible at VBG=V_{\text{BG}}= 50 V. The downward trend of such features suggests an increased series resistance at low VBGV_{\text{BG}}, consistent with increased RleadR_{\text{lead}}. For further discussion of plateau features in the normal state and series resistance correction, see section S5.

Fig. S15 shows the superconducting critical current IcI_{c} of an open constriction (VG12=V_{\text{G12}}= 3 V). Back gate voltage has a strong effect on normal state resistance. While the simultaneous lack of a strong trend in IcI_{c} of the lead is in apparent conflict with the trend in TcT_{c}, similar trends (optimal IcI_{c} at higher than optimal doping TcT_{c}) have been observed in LAO/STO Hall bars [58, 71].

S5 Normal state conductance of the constriction, evolution from normal to superconducting leads

S5.1 Sample 2

Sample 2 was studied across three cooldowns at VGIL=V_{\text{GIL}}= 3, 3.5, and 3.7 V. Larger VGILV_{\text{GIL}} increases the carrier density in the Hall bar channel (Fig. S10A), and drives it towards more robust metallicity (Fig. S16A) and superconductivity (Fig. S10B). Fig. S16B and S16C illustrate the evolution of constriction conductance behavior with VGILV_{\text{GIL}} and split gate voltage VG12V_{\text{G12}}, at base temperature (28-35 mK). Increased metallicity in the channel at higher VGILV_{\text{GIL}} translates into a rapid shift of the constriction pinch off point to lower VG12V_{\text{G12}}.

Refer to caption
Figure S16: Sample 2, normal state conductance of devices (A) 2A and (B) 2B at base temperature for three cooldowns at different VGILV_{\text{GIL}}. Raw data without RSR_{S} subtraction is shown.
Refer to caption
Figure S17: Sample 2, temperature dependence of constriction conductance at zero DC bias. The red color emphasizes a trace at an intermediate TT, which reduces the strength of low-TT fluctuations but remains below the onset of strong thermal smearing. (A) device 2A, VGIL=V_{\text{GIL}}= 3 V, (B) device 2A, VGIL=V_{\text{GIL}}= 3.5 V, (C) device 2B, VGIL=V_{\text{GIL}}= 3.5 V. A series resistance RSR_{S} = 3.5 kΩ\Omega was subtracted in (A, B) and 1.4 kΩ\Omega in (C).
Refer to caption
Figure S18: Illustration of series resistance subtraction from conductance-split gate voltage traces for device 2A at (A) VGIL=V_{\text{GIL}}= 3 V, 301 mK, (B) VGIL=V_{\text{GIL}}= 3.5 V, 569 mK. RS=R_{S}= 0 corresponds to the raw data. The red traces correspond to the ultimately chosen value of RS=R_{S}= 3.5 kΩ\Omega.

At VGIL=V_{\text{GIL}}= 3 V, the 60 nm wide constriction (device 2B) remains closed up to VG12=V_{\text{G12}}= 4.5 V. The 40 nm wide constriction (device 2A) becomes open near VG12=V_{\text{G12}}= 3 V. The occurrence of pinch-off at positive VG12V_{\text{G12}} despite metallic conductivity in the leads can be understood in terms of a built-in depletion field around constriction edges. At VGIL=V_{\text{GIL}}= 3.5 and 3.7 V, the Hall bar channel becomes superconducting at base temperature. In both cases, the constriction in device 2B also becomes superconducting in its open state at large VG12V_{\text{G12}}. The constriction in the device 2A does not show a clear supercurrent at VGIL=V_{\text{GIL}}= 3.5 V, but does become superconducting at 3.7 V. In presence of a supercurrent, the traces shown in Fig. S16 are taken in the normal state in magnetic field or at high DC bias.

Fig. S17 shows constriction conductance traces with VG12V_{\text{G12}} at different temperatures. Common trends for both devices 2A and 2B, VGIL=V_{\text{GIL}}= 3 and 3.5 V are: plateau signatures near both even and odd multiples of δG=e2/h\delta G=e^{2}/h, fluctuations of GG, repeatable between VG12V_{\text{G12}} sweeps, with amplitude decreasing with temperature. Quantization is best seen at an intermediate TT, where fluctuations are reduced, but the thermal smearing is not yet fully onset. This highlighted by selected red line traces in Fig. S17.

Additional evidence for conductance quantization is seen in DC bias spectroscopy in Fig. S19, S20 and S21. Line traces with VDCV_{\text{DC}} at different VG12V_{\text{G12}} tend to crowd around multiples of e2/he^{2}/h at VDC=0V_{\text{DC}}=0. This is again most clearly seen at a TT slightly elevated from base, and it is obscured at low TT by fluctuations. At finite VDCV_{\text{DC}}, traces crowd at half values between multiples of e2/he^{2}/h (most clearly seen in Fig. S19A). This is consistent with the classic picture of DC bias adding an extra available ballistic mode for carriers moving in one direction only [72].

Regarding conductance quantization in steps of δG=e2/h\delta G=e^{2}/h (as opposed to δG=2e2/h\delta G=2e^{2}/h), it is difficult to unambiguously disentangle the series contact resistance contribution. In this paper, we adopt the standard simple approach of subtracting a constant value RSR_{S} from the measured constriction resistance RR (Fig. S18). Identification of the lower conductance plateaus, particularly at the first two multiples of e2/he^{2}/h, is only weakly affected by the arbitrary choice of RSR_{S}. However, the absolute value of higher conductance plateaus is quite sensitive to small adjustments in RSR_{S}, and can thus easily be misidentified. Additionally, it is unclear whether in our device geometry RSR_{S} is truly independent of VG12V_{\text{G12}}. One reasonable scenario is a reduction of RSR_{S} at high VG12V_{\text{G12}} by carrier accumulation in the STO regions neighboring the constriction. One hint at challenges with constant RSR_{S} subtraction is the absence of plateau signatures at 4e2/h4e^{2}/h in Fig. S17A and 6e2/h6e^{2}/h in Fig. S17B. Nevertheless, the data in Fig. S17 is clearly more compatible with a pattern of multiples in e2/he^{2}/h rather than 2e2/h2e^{2}/h, particularly given fairly robust features at G=e2/hG=e^{2}/h and 3e2/h3e^{2}/h.

Conductance quantization with δG=e2/h\delta G=e^{2}/h is expected in any QPC when the spin degenaracy is lifted by a magnetic field BB [1]. However, even in the absence of BB, such half-quantization has been reported in many studies of gated constrictions and quantum wires based on InAs [73, 74, 65], GaAs [75, 76, 77], and carbon nanotubes [78]. The precise origins of this effect have arguably not been elucidated [74, 65]. One ingredient suspected to be important is the presence of either intrinsic or field-induced spin-orbit interaction, resulting in spontaneous spin polarization [79] or filtering transmission for opposite spins [80]. Another possible essential ingredient is electron-electron interaction in 1D confinement, creating a spin incoherent or helically ordered Luttinger liquid [81, 82]. Another proposed mechanism is disruption of the adiabaticity of the constriction-lead coupling by the disorder potential [65]. All of these mechanisms are potentially relevant for the case of a narrow constriction in STO. Quantization with δG=e2/h\delta G=e^{2}/h has also been reported for STO in accidental QPC’s in shorted line junctions [83] and LAO/STO wires [84].

Refer to caption
Figure S19: Constriction conductance map and corresponding line traces with VG12V_{\text{G12}} and DC bias. Device 2A, VGIL=V_{\text{GIL}}= 3 V at (A, B) 302 mK and (C, D) 35 mK. A series resistance RSR_{S} = 3.5 kΩ\Omega was subtracted in (A-D).
Refer to caption
Figure S20: Constriction conductance map and corresponding line traces with VG12V_{\text{G12}} and DC bias. Device 2A, VGIL=V_{\text{GIL}}= 3.5 V at (A, B) 751 mK and (C, D) 28 mK. A series resistance RSR_{S} = 3.5 kΩ\Omega was subtracted in (A-D).
Refer to caption
Figure S21: Constriction conductance map and corresponding line traces with VG12V_{\text{G12}} and DC bias. Device 2B, VGIL=V_{\text{GIL}}= 3.5 V at (A,B) 302 mK and (C, D) 28 mK. A series resistance RSR_{S} = 1.4 kΩ\Omega was subtracted in (A, B) and RSR_{S} = 0 in (C, D).

S5.2 Sample 1

Fig. S22 shows VG12V_{\text{G12}} sweeps at different temperatures for device 1A. All data shown in this section are for the same device state as in the main text: VGIL=V_{\text{GIL}}= 3 V. Unless specified otherwise, VBG=V_{\text{BG}}= 50 V.

Refer to caption
Figure S22: (A) Device 1A, temperature dependence of constriction conductance at zero DC bias, VGIL=V_{\text{GIL}}= 3 V, VBG=V_{\text{BG}}= 50 V. (B) Same at B=B= 0.25 T. (C) Peak-to-dip crossover in DC bias characteristics as a function of magnetic field, at 364 mK, VG12=V_{\mathrm{G12}}= 2 V. A series resistance RSR_{S} = 1.15 kΩ\Omega was subtracted in (A-C).

In Fig. S22A , the supercurrent dominates at low temperature. TT needs to be raised above 500 mK to access normal state conductance, with plateau-like features near G=2e2/hG=2e^{2}/h and 4e2/h4e^{2}/h. This fits the classic pattern of a QPC with spin-degenerate ballistic modes. Fig. S22B shows a similar measurement, but in a 0.25 T field, which suppresses the supercurrent at all temperatures. The pattern of fluctuations in GG with a decreasing amplitude at higher TT is more clearly visible in this measurement.

Fig. S22C illustrates a challenge with precisely assigning absolute value of plateau conductance in presence of the supercurrent. In the normal state at high TT, excess conductance persists above TcT_{c}. Upon applying a magnetic field, a negative contribution to conductance appears above BcB_{c}. Above TcT_{c}, this manifests itself as a field-driven crossover between a peak and a dip of GG around VDC=V_{\text{DC}}= 0. The same effect is responsible for traces Fig. S22A consistently lying above the ones in Fig. S22B.

This zero bias peak behavior persists at temperatures significantly above TcT_{c}. The conductance map with VG12V_{\text{G12}} and VDCV_{\text{DC}} at 511 mK in Fig. S23A, B has zero bias peaks across entire the entire VG12V_{\text{G12}} range, with peak heights up to 0.5e2/h\approx 0.5e^{2}/h. Fig. S24A and B show similarly sized dips at B=B= 0.5 T and 45 mK. Fig. S23C, D shows a measurement at 364 mK and 0.1 T, i.e. very close to the peak to dip crossover in Fig. S22C. It shows clear trace crowding near G=2e2/hG=2e^{2}/h and 4e2/h4e^{2}/h (n=n= 1, 2), and much weaker crowding slightly below 6e2/h6e^{2}/h and 8e2/h8e^{2}/h (n=n= 3, 4).

Fig. S24C and D show a measurement at 0.5 T, as a function of a single local gate VG1V_{\text{G1}}, with the other gate fixed at VG2=V_{\text{G2}}= 0.9 V. As discussed in section S7, this gate trajectory bypasses a number of unintentional quantum dot resonances and thus shows a cleaner observation of trace crowding near G=2e2/hG=2e^{2}/h and 4e2/h4e^{2}/h

Fig. S25A and B show the conductance map at base temperature. This measurement taken in the same state as Fig. 2 in the main text, but with smaller resolution and across a larger DC bias range. The supercurrent and the subgap structure makes plateau identification difficult in this state. Trace crowding can nevertheless be seen below VDCV_{\text{DC}}\approx 200 μ\muV, near GNG_{N}\approx 1.5 and 2.5 e2/he^{2}/h (corresponding to n=n= 1 and 2, no RSR_{S} was subtracted in this plot). Above 200 μ\muV, most regions with trace crowding get split, as one expects for ballistic modes of a QPC [72].

While plateau features near even multiples of 2e2/h2e^{2}/h were emphasized in the above discussion of device 1A, other features can also be identified near G=0.2e2/hG=0.2e^{2}/h, e2/he^{2}/h, and 2.5e2/h2.5e^{2}/h (most easily seen in Fig. S22 and Fig. S26). As discussed in section S7, these features coincide with charging resonances of an accidental Coulomb blockade.

The overall situation also bears resemblance to the one in [65], where gate voltage tuning of the disorder potential surrounding the constriction resulted into spurious appearances of features and plateaus at odd integer multiples of e2/he^{2}/h. In our case, the back gate voltage VBGV_{\text{BG}} serves a similar function by tuning the depth of the 2DES. This is seen in the data previously shown in Fig. S13C. Fig. S26 shows the same measurement at a temperature closer to TcT_{c} (511 instead of 866 mK). Fig. S26B reproduces Fig. 2E in the main text. In raw data, all features in GG are gradually shifted downwards in GG as VBGV_{BG} is lowered due to a VBGV_{BG}-dependent series resistance RSR_{S}. Subtraction of RSR_{S} that matches the features near 2e2/h2e^{2}/h and 4e2/h4e^{2}/h naturally aligns most other features in GG. Plateau signatures at n=n= 1 and 2 are present for all VBGV_{BG}. At intermediate VBG=V_{\text{BG}}= 30-40 V, faint features are also visible near G=6e2/hG=6e^{2}/h and 8e2/h8e^{2}/h (n=n= 3, 4). The Coulomb blockade features are near G=0.2e2/hG=0.2e^{2}/h, e2/he^{2}/h, and a set of smaller fluctuations near 2.5e2/h2.5e^{2}/h. The location of the latter feature is particularly sensitive to VBGV_{\text{BG}}, consistent with back gate tuning of the disorder potential.

Refer to caption
Figure S23: Constriction conductance map and corresponding line traces with VG12V_{\text{G12}} and DC bias. Device 1A, VGIL=V_{\text{GIL}}= 3 V, VBG=V_{\text{BG}}= 50 V at (A, B) 511 mK, 0 T and (C, D) 364 mK, 0.1 T. A series resistance RSR_{S} = 1.15 kΩ\Omega was subtracted in (A-D).
Refer to caption
Figure S24: (A, B) Constriction conductance map and corresponding line traces with VG12V_{\text{G12}} and DC bias. Device 1A, VGIL=V_{\text{GIL}}= 3 V, VBG=V_{\text{BG}}= 50 V, 45 mK, 0.5 T. (C, D) Same, but B=B= 0.25 T and the local gate voltage is applied to only one gate (VG1V_{\text{G1}}), the other gate (VG2V_{\text{G2}}) is fixed at 0.9 V. A series resistance RSR_{S} = 1.15 kΩ\Omega was subtracted in (A-D).
Refer to caption
Figure S25: Constriction conductance map and corresponding line traces with VG12V_{\text{G12}} and DC bias. Device 1A, VGIL=V_{\text{GIL}}= 3 V, VBG=V_{\text{BG}}= 50 V, 45 mK. No series resistance was subtracted.
Refer to caption
Figure S26: Constriction conductance traces with split gate voltage, at fixed VBGV_{\text{BG}}= 0-50 V, taken in the normal state at T= 511 mK, VGIL=V_{\text{GIL}}= 3 V, Device 1A. (A) Data presented without series resistance subtraction. Solid light purple lines indicate features corresponding to n=n= 1, 2, 3, 4. (B) Same data, presented with subtraction of a variable series resistance gradually decreasing from to 2.1 to 1.15 kΩ\Omega with VBGV_{BG} (shown in the inset).

S6 DC bias spectroscopy and critical current in the superconducting state

This section presents conductance maps with DC bias and split gate voltage for various devices in the superconducting regime. Fig. S27A-D shows the same data as in the Fig. 2 of the main text. All of the data shown is taken in the same measurement, in which a nominal DC bias is applied to an ohmic contact, a more accurate DC bias VDCV_{\text{DC}} is measured at voltage probes near the constriction, and the DC current IDCI_{\text{DC}} is measured at the grounded ohmic contact. The selected traces of GG with VDCV_{\text{DC}} in Fig. S27 illustrate the split gate-driven crossover from complete pinch-off to tunneling and Josephson junction regimes.

Refer to caption
Figure S27: Back gate tuning of supercurrent in device 1A. (B) Constriction conductance as a function of VG12V_{\text{G12}} and VDCV_{\text{DC}}. Selected cuts in VDCV_{\text{DC}} at VG12V_{\text{G12}} indicated by diamond markers are plotted in (A). Dashed lines in (A,B) indicate ±Δ/e\pm\Delta/e and ±2Δ/e\pm 2\Delta/e estimated from the TcT_{c} in the leads. (C) Constriction resistance normalized to its normal state value, taken at VDC=V_{\text{DC}}= 100 μ\muV. The solid red line indicates the critical current at R=RN/2R=R_{N}/2. The dashed lines indicate the integer multiples of the critical current quantum δIc\delta I_{c}. (D) Normalized critical current and normal state conductance as a function of split gate voltage. Both quantities tracks the number of spin degenerate ballistic modes nn in the constriction. (A-C) is for device 1A, VGIL=V_{\text{GIL}}= 3 V, VBG=V_{\text{BG}}= 50 V, TT = 45 mK. (E-H) Same as top row with VBG=V_{\text{BG}}= 25 V. (I-L) Same as top row with VBG=V_{\text{BG}}= 0 V. Series resistance correction is only applied to GNG_{N} in the rightmost plot column: RS=R_{S}= 0.8 kΩ\Omega (D), 1.1 kΩ\Omega (H), 1.6 kΩ\Omega (L).
Refer to caption
Figure S28: Direct comparison of data at VBG=V_{\text{BG}}= 50, 25, 0 V, also shown in the different rows of Fig. S27. (A) Normal state conductance. (B) Critical current (lines) and excess current (circles). (C) normalized critical current. (D) SN contact transparency extracted from eIexcRN/ΔeI_{\text{exc}}R_{N}/\Delta, plotted as a function of normal state conductance. Correction for RSR_{S} was applied to GNG_{N} in figures (A, D), with same values as in Fig. S27
Refer to caption
Figure S29: Same data type as Fig. S27. (A-D) is for device 1B, VGIL=V_{\text{GIL}}= 3 V, VBG=V_{\text{BG}}= 50 V, TT = 45 mK. (E-H) is for device 2B, VGIL=V_{\text{GIL}}= 3.5 V, TT = 27 mK. (I-L) is for device 2A, VGIL=V_{\text{GIL}}= 3.7 V, TT = 28 mK. GNG_{N} is taken at 40 μ\muV in (G, H) and 100 μ\muV in (C, D, K, L). GNG_{N} was corrected for RS=R_{S}= 1 kΩ\Omega in (D), 0.5 kΩ\Omega in (H), 0 kΩ\Omega in (L).
Refer to caption
Figure S30: (A) Direct comparison of the critical (circles) and excess (lines) current in devices 1A, 1B, 2A, 2B. Data for device 2B is scaled by a factor of 10. (B) Equivalent SN transparency from eIexcRN/ΔeI_{\text{exc}}R_{N}/\Delta, with the gap value taken from TcT_{c} in the leads. It is plotted as a function of normal state resistance from Fig. S27 and S29, without correction for RSR_{S}.
Refer to caption
Figure S31: Devices 3A and 3B at VGIL=V_{\text{GIL}}= 3 V, VBG=V_{\text{BG}}= 0 and 50 V, T=T= 29 mK. (A) Normal state conductance at VDC=V_{\text{DC}}= 250 μ\muV. (B) Critical (connected circles) and excess (lines) current. (C) IcI_{c} normalized by δIc\delta I_{c}, matched to GNh/2e2G_{N}\cdot h/2e^{2}. (D) Equivalent SN transparency from eIexcRN/ΔeI_{\text{exc}}R_{N}/\Delta, with the gap value taken from TcT_{c} in the leads. No series resistance correction was applied.
Refer to caption
Figure S32: (A) IcRNI_{c}R_{N} product as a function of normal state conductance. See legend for device state details. (B) Same as (A), but with IcRNI_{c}R_{N} normalized to the gap value from lead TcT_{c}. (C) δIc\delta I_{c}, the critical current per ballistic mode, plotted as a function of SN transparency (averaged in the GN=G_{N}= 4-6 e2/he^{2}/h region). Blacked dashed line is the short limit SNS model (eq. S3), gray dashed line is the same model with an additional factor α=2/3\alpha=2/3 from finite weak link length L=ξL=\xi (eq. S6).

In the tunneling regime (VG12<V_{\text{G12}}< 1 V), conductivity is suppressed near zero bias. The gap is indicated in Fig. S27A and B at Δ=1.76kBTc\Delta=1.76k_{B}T_{c} [85]. The TcT_{c} value is extracted from the measurement of lead resistance RleadR_{\mathrm{lead}} with temperature. Coherence peaks in conductance are seen near VDC=Δ/eV_{\text{DC}}=\Delta/e. This is consistent with the expectations of tunneling across a superconductor/normal metal (SN) interface [67]. In the SNS geometry of our device, this regime can be understood as tunneling across two SN interfaces in series, as discussed in [83]. The intermediate regime (VG12=V_{\text{G12}}= 1-1.5 V) with intermittent supercurrent is discussed in section S7. In the Josephson regime (VG12>V_{\text{G12}}> 1.5 V), there is a robust supercurrent at zero bias. At VDC=Δ/eV_{\text{DC}}=\Delta/e, coherence peaks at low VG12V_{\text{G12}} evolve into conductance dips at high VG12V_{\mathrm{G12}}. Smaller dips can be seen near VDC=2Δ/eV_{\text{DC}}=2\Delta/e. This inversion is characteristic of SNS junctions with highly transparent SN interfaces [86, 10].

Similarly to Fig. 2 in the main text, the analysis of the critical supercurrent is presented in Fig. S27C using constriction resistance normalized to its normal state value RNR_{N}, extracted in the same measurement as the resistance at VDC=V_{\text{DC}}= 100 μ\muV. A direct comparison between IcI_{c} and GNG_{N} is shown in Fig. S27D by normalizing both qunatities into a number of ballistic modes nn. GNG_{N} is divided by δGN=2e2/h\delta G_{N}=2e^{2}/h under the assumption of spin degenerate ballistic modes. IcI_{c} is divided by δIc=eΔ/\delta I_{c}=e\Delta/\hbar, chosen to match the plateau structure in IcI_{c}. This plot emphasizes the numerical correspondence between these two independently measured quantities.

As shown in Fig. S27E-L, lowering the back gate voltage VBGV_{\text{BG}} from 50 V to 25 and 0 V shifts the constriction pinch-off point from VG12=V_{\text{G12}}= 0.5 to 1 and 1.5 V. The patterns of tunnel to Josephson junction crossover, IcI_{c} quantization, and numerical correspondence of Ic/δIcI_{c}/\delta I_{c} with GNG_{N} are preserved. This is emphasized in the side-by-side comparison of key quantities at different VBGV_{\text{BG}}, shown in Fig. S28. δIc\delta I_{c} is slightly reduced from 2.48 nA (VBG=V_{\text{BG}}= 50 V) to 2.34 (25 V) and 2.19 nA (0 V). Fig. S28B also shows the excess current IexcI_{\text{exc}}, which approximately follows IcI_{c} for all VBGV_{\text{BG}}. The SN interface transparency τSN\tau_{\text{SN}} extracted from the quantity eIexcRN/ΔeI_{\text{exc}}R_{N}/\Delta (see section S1.5) is shown in Fig. S28D. When plotted as a function of constriction conductance, τSN\tau_{\text{SN}} overlaps for all VBGV_{\text{BG}}, including the dips near n=n= 1, 2, 3. These features are a natural consequence of a sharper plateau structure in IcI_{c} in comparison to GNG_{N}.

Fig. S29A-D shows the measurement on device 1B, a slightly wider (60 nm nominal width) constriction on the same Hall bar, at VBG=V_{\text{BG}}= 50 V. A prolonged tunneling regime and an intermittent weak supercurrent is seen at VG12=V_{\text{G12}}= 1-2 V. At VG12>V_{\text{G12}}> 2V, a good correspondence numerical correspondence in nn from Ic/δIcI_{c}/\delta I_{c} and GNG_{N} is achieved with δIc=\delta I_{c}= 2.04 nA, a value slightly reduced but close to device 1A. The VDCV_{\text{DC}} dependence also shows a tunneling regime with coherence peaks near VDC=Δ/eV_{\text{DC}}=\Delta/e at low VG12V_{\text{G12}}, that evolve into conductance dips at higher VG12V_{\text{G12}}.

Fig. S29E-L shows data from devices 2A and 2B from separate cooldowns. Device 2B at VGIL=V_{\text{GIL}}= 3.5 V is in the strongly underdoped regime. The lead TcT_{c} is 167 mK. In comparison to device 1A, The VDCV_{\text{DC}} dependence is re-scaled to a smaller gap, but retains the essential features: coherence peaks in the tunneling regime, conductance dips near VDC=Δ/eV_{\text{DC}}=\Delta/e and 2Δ/e2\Delta/e in the Josephson regime. The critical current in this device is also much smaller in magnitude and exhibits strong fluctuations. Its dependence on VG12V_{\text{G12}} is not smooth and has several short plateaus, which coincide with similar features in GNG_{N}. The plateau assignment is complicated by the presence of half-integer features in GNG_{N}, and sensitivity to the choice of RSR_{S} at high conductance. Keeping the spin-degenerate mode notation (G=n2e2/hG=n\cdot 2e^{2}/h), short IcI_{c} plateaus are seen at at n=n= 3, 4 and 4.5. In the n=n= 1, 2 region, IcI_{c} is intermittent between zero and nδIcn\cdot\delta I_{c}.

IcI_{c} and IexcI_{\text{exc}} of this device is shown in Fig. S30, both multiplied by 10x for comparison with devices 1A, 1B and 2A. IexcI_{\text{exc}} again follows IcI_{c}, but lags behind it. The SN transparency extracted from eIexcRN/ΔeI_{\text{exc}}R_{N}/\Delta is τSN\tau_{\text{SN}} = 0.6, appreciably lower than for devices 1A and 1B. This is consistent with the relatively low δIc\delta I_{c}, which is suppressed by a factor of 7 in comparison to eΔ/e\Delta/\hbar.

Device 2A at VGIL=V_{\text{GIL}}= 3.7 V is in the overdoped regime, with the lead TcT_{c} at 253 mK. The gate tunability with VG12V_{\text{G12}} is strongly reduced in comparison to the VGIL=V_{\text{GIL}}= 3.5 V state. The constriction does not reach the pinch-off within the available range of VG12V_{\text{G12}}. GNG_{N} is only modulated between 55 and 7e2/h7e^{2}/h. IcI_{c} is only slightly modulated around 27 nA. With such a weak modulation, normalization by δIc\delta I_{c} can only be done by numerically mathching Ic/δIcI_{c}/\delta I_{c} to GNG_{N}. It is thus not possible to reliably establish whether the ballistic SQPC picture applies for this device. The VDCV_{\text{DC}} dependence of GG is characterized by a very large dip in conductance near VDC=Δ/eV_{\text{DC}}=\Delta/e. This feature results in the excess current that is strongly decreased in comparison to IcI_{c} (Fig. S30A). The implied transparency is however very high, τSN=\tau_{\text{SN}}= 0.8-0.9.

A very similar situation was observed in devices 3A and 3B (Fig. S31), which were measured in the high carrier density regime, with lead Tc=T_{c}= 256 mK. Both constriction conductance and IcI_{c} are only weakly modulated by VG12V_{\text{G12}}. The excess current is significantly lower than IcI_{c} and the implied transparency is τSN=\tau_{\text{SN}}= 0.8-1. In device 3A at high VG12V_{\text{G12}} and VBG=V_{\text{BG}}= 0 V, eIexcRN/ΔeI_{\text{exc}}R_{N}/\Delta exceeds 2.64, the ideal transparency limit in the SNS model.

Fig. S32 compares the IcRNI_{c}R_{N} product of all devices discussed in this section. IcRNI_{c}R_{N} is presented as a function of GNG_{N} (as a proxy for VG12V_{\text{G12}}) to emphasize the distinction between gate-tunable devices with lower IcRNI_{c}R_{N} and weakly tunable ones with a higher IcRNI_{c}R_{N}. Normalizing by the superconducting gap Δ/e\Delta/e (from lead TcT_{c}) further emphasizes this clustering into two groups. This normalization is rationalized by the general expectation that IcRNI_{c}R_{N} of a Josephson junction scales with the gap [87]. For an ideal ballistic SNS constriction (τSN=\tau_{\text{SN}}= 1, LξL\ll\xi), IcRN=δIch/2e2=πΔ/eI_{c}R_{N}=\delta I_{c}\cdot h/2e^{2}=\pi\Delta/e.

In experiments on SNS junctions, IcRNI_{c}R_{N} is ubiquitously used as a metric for junction quality. In most casses IcRNI_{c}R_{N} is substantially lower than Δ/e\Delta/e, and IcRNI_{c}R_{N} of order Δ/e\Delta/e is often invoked as a signature of a high quality junction [88, 89, 90, 91, 92, 93]. For devices 1A and 1B, IcRNI_{c}R_{N} is approximately at or slightly below Δ/e\Delta/e, which is lower than πΔ/e\pi\Delta/e by factor of 3-5. This statement is equivalent to the discussion in the main text on suppression of δIc\delta I_{c} in comparison to eΔ/e\Delta/\hbar, provided that GNh/2e2G_{N}\cdot h/2e^{2} numerically matches with Ic/δIcI_{c}/\delta I_{c}. The corresponding plot of δIc\delta I_{c} as a function of τSN\tau_{\text{SN}} is shown in Fig. S32C. It illustrates that the data on devices 1A, 1B and 2B is consistent with the ballistic SQPC picture, with the weak link length approximately equal to or shorter than the coherence length.

For the devices in the high carrier density limit (2B at VGIL=V_{\text{GIL}}= 3.7 V, 3A, 3B), IcRNI_{c}R_{N} exceeds πΔ/e\pi\Delta/e by a factor of 1.2 - 2. A natural explanation is a crossover from an SNS junction to an SS’S constriction or wire. Establishing a crisp picture requires further study, but two frameworks can be invoked as useful starting points. On one hand, IcRNI_{c}R_{N} in a superconducting wire (S’) connecting two superconducting reservoirs (S) is expected to increase with length until the onset of decoherence, and can exceed πΔ/e\pi\Delta/e [87]. On the other hand, one can make a comparison to the STO leads themselves, which show “weak superconductivity” with a relatively small critical current. A 20x20 μ\mum square of STO in the leads shows eIcRN/Δ=eI_{c}R_{N}/\Delta= 4.5 near devices 1A and 1B and 3.2 near devicees 3A and 3B. In [58], a 50 μ\mum long and 20 μ\mum wide LAO/STO Hall bar has been documented to show eIcRN/Δ=eI_{c}R_{N}/\Delta= 25-70. This was rationalized in terms of an interconnected Josephson junction array, where IcRNI_{c}R_{N} scales with its size and can easily exceed Δ/e\Delta/e. For the case of a square array of NJJ×NJJN_{\text{JJ}}\times N_{\text{JJ}} junctions, eIcRN/Δ=NJJπ/2eI_{c}R_{N}/\Delta=N_{\text{JJ}}\pi/2 [94].

S7 Tunneling regime and accidental Coulomb blockade near pinch-off

In the split gate geometry of our device, one can asymmetrically set the gate voltages VG1V_{\text{G1}} and VG2V_{\text{G2}}. This has the effect of moving the saddle potential location around the constriction. In this manner, one can map the disorder landscape in the QPC, as shown in Fig. S33. The conductance in both the normal (GNG_{N}) and superconducting states (GSG_{S}) rises with VG1V_{\text{G1}} and VG2V_{\text{G2}} in a largely symmetric way. This confirms that the capacitances of the two split gates of our QPC are similar, as intended.

Both GNG_{N} and GSG_{S} show several sets of line resonances in the VG1VG2V_{\text{G1}}-V_{\text{G2}} space, at which conductance is increased. These resonances remain pronounced when plotting the GS/GNG_{S}/G_{N} ratio. In GSG_{S}, they are particularly pronounced at lower gate voltages and near intersections between different resonances. The intersections of these resonances correlate with the intermittent critical current seen near the first plateau (GN2e2/hG_{N}\approx 2e^{2}/h) in Fig. 2 in the main text. They also coincide with the plateau-like features seen in GNG_{N} (Fig. S26B) near 0.2e2/he^{2}/h, e2/he^{2}/h (i.e. inconsistent with the 2e2/h2e^{2}/h quantization), and the smaller features near 2-2.5e2/he^{2}/h.

We attribute these resonances to charging levels of an accidental Coulomb blockade. Spontaneous quantum dot formation near pinch-off in LAO/STO constrictions has been documented in multiple reports [19, 20, 21]. The situation in our case is qualitatively similar. DC bias spectroscopy reveals conductance diamonds near the first two charging levels (VG12=V_{\text{G12}}= 0.6 and 0.8 V at zero DC bias in Fig. S34B). While only 2 charging levels are clearly distinguishable before the onset of a supercurrent, their height (charging energy) starts near 400 μV\mu V and appears to rapidly decrease with VG12V_{\text{G12}}, following the same trend as in [23, 24, 19, 20, 21]. The charging energy is likely dominated by the electrostatic capacitance of the dot rather than the orbital contribution [21]. Its decrease with VG12V_{\text{G12}} can be understood as an increase in quantum dot size or tunnel barrier capacitance, although quantifying them is difficult due to the strongly electric field dependent permittivity of STO [21].

Fig. S35 shows a map with VG12V_{\text{G12}} and small DC bias. Fig. S36 shows a similar map, but with only VG1V_{\text{G1}} being swept and VG2V_{\text{G2}} fixed at 0.9 V, a trajectory that minimizes the amount of encounters with charging resonances. In both cases, at low VG12V_{\text{G12}} and away from the coulomb blockade charging levels, tunneling conductance is observed: GG is strongly suppressed at zero bias, and coherence peaks are seen near VDC=±Δ/eV_{\text{DC}}=\pm\Delta/e. For comparison with experimental conductance, the gap value extracted from TcT_{c} in the leads is indicated by the dashed line in Fig. S34, S35, and S36. This is consistent with the expectation of tunneling across an SN interface [67]. Applicability to our case can be rationalized by considering the SNS junction as two SN interfaces in series [83]. The gradual increase of subgap conductance with VG12V_{\text{G12}} is consistent with a decrease in tunnel barrier strength [67].

Besides the peaks at ±Δ/e\pm\Delta/e, tunneling conductance shows additional in-gap features: double peaks at VDCV_{\text{DC}} considerably lower than Δ/e\Delta/e (\approxμ\muV), and zero bias peaks close to pinch-off. This is reminiscent of the in-gap states observed in vertical LAO/STO tunnel junctions [95]. Possible explanations involve two-band superconductivity with a small second gap [96, 29], suppression of the superconducting order parameter next to the tunneling barrier due to proximity effect [97, 61], Kondo effect [98], Majorana or Andreev bound states [49, 95, 24]. At the present stage, we do not attempt to discriminate between these possibilities.

Refer to caption
Figure S33: Potential mapping by independently sweeping the two split gate voltages VG1V_{\text{G1}} and VG2V_{\text{G2}}, in device 1A at VDC=V_{\text{DC}}= 0, T=T= 45 mK, VGIL=V_{\text{GIL}}= 3 V, VBG=V_{\text{BG}}= 50 V. (A) In the normal state at B=B= 0.25 T. (B) in the superconducting state at B=B= 0. (C) Same as (B), but on a different color scale, emphasizing features at GG<10 e2/he^{2}/h. Solid lines indicate the gate sweep trajectory in Fig. S36 and S35 (D) Ratio of conductance in the superconducting and normal states.
Refer to caption
Figure S34: Conductance map with DC bias and split gate voltage VG12=VG1=VG2V_{\text{G12}}=V_{\text{G1}}=V_{\text{G2}}. (A) Showing the entire VG12V_{\text{G12}} range. The dashed lines indicate VDC=±Δ/eV_{\text{DC}}=\pm\Delta/e. (B) Same data, but focusing on the Coulomb blockade diamonds seen at low VG12V_{\text{G12}} and maximum VDCV_{\text{DC}} range. Data shown is for device 1A at T=T= 45 mK, VGIL=V_{\text{GIL}}= 3 V, VBG=V_{\text{BG}}= 50 V.
Refer to caption
Figure S35: (A) Conductance map with DC bias and split gate voltage VG12=VG1=VG2V_{\text{G12}}=V_{\text{G1}}=V_{\text{G2}}, measurement range focused on the small VDCV_{\text{DC}} range in the tunneling and intermittent supercurrent regimes. The dashed lines indicate VDC=±Δ/eV_{\text{DC}}=\pm\Delta/e. Circle markers in (A) indicate the gate voltage position of line cuts shown in (B). Data shown is for device 1A at T=T= 45 mK, VGIL=V_{\text{GIL}}= 3 V, VBG=V_{\text{BG}}= 50 V.
Refer to caption
Figure S36: Same as Fig. S35, but only a single split gate voltage VG1V_{\text{G1}} is swept, VG2V_{\text{G2}} is fixed at 0.9 V.

S8 Determination of the superconducting gap

The superconducting gap Δ\Delta in SrTiO3 has been shown to be remarkably close to the BCS estimate [85]. In the zero-temperature limit

Δ0=1.76kBTc.\Delta_{0}=1.76k_{B}T_{c}. (S24)

The temperature dependence is well approximated by

Δ(T)=Δ0tanh(1.74(TcT1)12).\Delta(T)=\Delta_{0}\cdot\tanh\left(1.74\cdot\left(\frac{T_{c}}{T}-1\right)^{\frac{1}{2}}\right). (S25)

A difficulty in our device geometry is to choose the appropriate transition temperature for estimating Δ\Delta. The simplest approach is to convert the TcT_{c} from the midway point of the resistance drop in the leads. For device 1A at VBG=V_{\text{BG}}= 50 V, Tc=T_{c}= 350 mK corresponds to Δ/e=\Delta/e= 42 μ\muV. This is the approach adopted throughout this paper for comparison to δIc\delta I_{c}, IcRNI_{c}R_{N}, and the structure in VDCV_{\text{DC}} dependence of GG.

Refer to caption
Figure S37: Different estimates for the superconducting gap and TcT_{c} in device 1A at VBG=V_{\text{BG}}= 50 V. Solid lines are TcT_{c} from the midpoint of the resistive transition in the leads and constriction, plotted as a function of VG12V_{\text{G12}} (full data is shown in Fig. S12). The circle marker at VG12=V_{\text{G12}}= 3 V is the estimate from fitting the excess current to a BCS gap (Fig. S39). The dashed line is the energy scale corresponding to the IcI_{c} quantization step (see Fig. S27)

This approach is based on a sheet resistance measurement, physically separated by 5 microns from the gated constriction. This sidesteps the intricacies of the electrostatic potential landscape in the immediate vicinity of the split gate. While the primary effect of VG12V_{\text{G12}} is to tune the carrier density in the constriction, it is likely that the electric field lines extend into the leads. Due to SrTiO3 being a semiconducting superconductor, this can locally affect the TcT_{c} and Δ\Delta that govern the Josephson effect of our junction. This situation is in contrast with hybrid semiconductor/superconductor systems, where the superconductor is typically a metal that is only negligibly affected by electrostatic gating.

The temperature dependence of the constriction resistance provides another estimate of TcT_{c} and Δ\Delta. The midpoint of the resistive transition is at 275 mK at maximum VG12V_{\text{G12}}. It decreases to \approx 240 mK near the transition to a closed constriction. This is a low estimate for Δ\Delta.

Refer to caption
Figure S38: (A) Temperature and DC current dependence of the constriction resistance, normalized to its normal state value at high bias. The critical current is indicated by the red solid line. (B) Selected cuts from the same data, plotted without normalization. (C) Comparison of: extracted IcI_{c} normalized to the low TT limit Ic(0)I_{c}(0) taken at 45 mK, SNS model (eq. S3) with different values of SN boundary transparency, thermal broadening criterion calculated using Ic(0)I_{c}(0) at base temperature and temperature-dependent values of IcI_{c}. All data shown are for device 1A at VBG=V_{\text{BG}}= 50 V, VG12=V_{\text{G12}}= 3 V.
Refer to caption
Figure S39: (A) Temperature dependence of the excess current, normalized to base temerature value (markers). The solid line is the temperature dependence of the BCS gap with Tc=T_{c}= 405 mK. (B) Selected cuts in DC current, illustrating the broadened peak in GG that persists above 400 mK. All data shown are for device 1A at VBG=V_{\text{BG}}= 50 V, VG12=V_{\text{G12}}= 3 V.
Refer to caption
Figure S40: (A) Device 1A, VBG=V_{\text{BG}}= 50 V, VG12=V_{\text{G12}}= 3 V (Josephson junction regime). Temperature and DC bias dependence of the constriction resistance, normalized to its normal state value at high temperature. The dashed line is the BCS gap dependence with Tc=T_{c}= 350 mK, corresponding to a peak in RR. (B) Selected cuts in DC bias from the same data, plotted as conductance without normalization. (C) Device 1A, VBG=V_{\text{BG}}= 50 V, VG12=V_{\text{G12}}= 0.9 V (tunneling junction regime). Temperature and DC bias dependence of the constriction conductance, normalized to its normal state value at high temperature. The dashed line is the BCS gap dependence with Tc=T_{c}= 290 mK, corresponding to a peak in GG. (D) Selected cuts in DC bias from the same data, plotted as conductance without normalization.

A likely pitfall of this approach is thermal broadening in the supercurrent. To illustrate this, constriction resistance as a function of temperature and DC current is shown in Fig. S38. Below 200 mK, the critical current only slightly decreases with temperature. This is consistent with the expected dependence for a short SNS (eq. S3) in presence of finite SN transparency. Above 200 mK, that model does not accurately describe IcI_{c}. IcI_{c} very briefly increases near 200 mK and quickly decreases to zero. This coincides with a rapid broadening in the IDCI_{\text{DC}} dependence of GG.

In an overdamped Josephson junction (2eIcRN2C/12eI_{c}R_{N}^{2}C/\hbar\ll 1, with CC being the junction capacitance), thermal broadening is governed by the dimensionless criterion γT=ekBT/Ic\gamma_{T}=ek_{B}T/\hbar I_{c} [99]. Within this model, the supercurrent gets significantly rounded for γT>0.1\gamma_{T}>0.1 and completely suppressed for γT>1\gamma_{T}>1. As shown in Fig. S38C, γT\gamma_{T} reaches 1 near 200 mK, rationalizing the rapid decrease of measured IcI_{c}. The overdamped regime hypothesis is consistent with a symmetric IcI_{c} without any hysteresis in IDCI_{\text{DC}}.

Broadening alone does not explain the apparent increase in IcI_{c} near γT\gamma_{T}, which could be a manifestation of proximity effect. In a SS’NS’S junction, where S’ is a proximitized normal metal, the induced pairing gap Δ\Delta’ is suppressed in comparison to the bulk gap Δ\Delta in the S region. At higher temperature, the two gaps merge, increasing the relative strength of the proximity effect Δ\Delta/Δ/\Delta [61].

Another way to probe the superconducting gap is to look at the temperature dependence of the excess current (Fig. S38), which is expected to scale as IexcΔ/eRNI_{\text{exc}}~{}\Delta/eR_{N} [67, 38]. Below 400 mK, IexcI_{\text{exc}} is well described by a scaled BCS dependence (eq. S25) with Tc=T_{c}= 405 mK. This provides an upper estimate for Δ\Delta.

Surprisingly, IexcI_{\text{exc}} does not completely vanish above the TcT_{c} implied by the BCS dependence. This residual IexcI_{\text{exc}} can also be seen as a small, heavily broadened dip in RR persisting above TcT_{c} (Fig. S39B). We speculate that this might be a signature of pre-formed Cooper pairs without macroscopic coherence [23, 100].

An independent confirmation of the superconducting gap can in principle be extracted from DC bias spectroscopy. Fig. S40A shows GG as a function of VDCV_{\text{DC}} and temperature. It also shows a BCS gap dependence for Tc=T_{c}= 350 mK, which matches the RR peak (GG dip) feature identified in Fig. S27. The temperature dependence of this feature matches the BCS prediction, but thermal broadening sets in prior to the expected decrease of Δ\Delta to zero. Consequently, from this data the transition point can only be estimated to be consistent with the 240-405 mK range discussed above. In the tunneling regime at low VG12V_{\text{G12}}, conductance peaks corresponding to TcT_{c}\approx 290 mK are clearly seen at low temperature (Fig. S40C). However, similarly to the Josephson regime at high VG12V_{\text{G12}}, thermal broadening obscures the transition region.

In summary, the uncertainty on the gap can be summarized as follows: the middle estimate is from the resistance drop with TT in the leads (Tc=T_{c}= 350 mK, Δ/e=\Delta/e= 42 μ\muV), the lower estimate is from the resistance drop in the constriction (TcT_{c}\approx 240 mK, Δ/e=\Delta/e= 29 μ\muV), the high estimate is from the temperature dependence of the excess current (Tc=T_{c}= 405 mK, Δ/e=\Delta/e= 49 μ\muV). Independent estimates from DC bias spectroscopy are consistent with TcT_{c} falling within that range.