High transparency induced superconductivity in field effect two-dimensional electron gases in undoped InAs/AlGaSb surface quantum wells

E. Annelise Bergeron Institute for Quantum Computing, University of Waterloo, Waterloo N2L 3G1, Canada Department of Physics, University of Waterloo, Waterloo N2L 3G1, Canada    F. Sfigakis corresponding author: [email protected] Institute for Quantum Computing, University of Waterloo, Waterloo N2L 3G1, Canada Northern Quantum Lights inc., Waterloo N2B 1N5, Canada Department of Chemistry, University of Waterloo, Waterloo N2L 3G1, Canada    A. Elbaroudy Department of Electrical and Computer Engineering, University of Waterloo, Waterloo N2L 3G1, Canada Department of Physics, University of Waterloo, Waterloo N2L 3G1, Canada    A. W. M. Jordan Institute for Quantum Computing, University of Waterloo, Waterloo N2L 3G1, Canada Department of Chemistry, University of Waterloo, Waterloo N2L 3G1, Canada    F. Thompson Institute for Quantum Computing, University of Waterloo, Waterloo N2L 3G1, Canada   
George Nichols
Institute for Quantum Computing, University of Waterloo, Waterloo N2L 3G1, Canada Department of Physics, University of Waterloo, Waterloo N2L 3G1, Canada
   Y. Shi Department of Electrical and Computer Engineering, University of Waterloo, Waterloo N2L 3G1, Canada Department of Physics, University of Waterloo, Waterloo N2L 3G1, Canada Waterloo Institute for Nanotechnology, University of Waterloo, Waterloo N2L 3G1, Canada    Man Chun Tam Department of Electrical and Computer Engineering, University of Waterloo, Waterloo N2L 3G1, Canada Waterloo Institute for Nanotechnology, University of Waterloo, Waterloo N2L 3G1, Canada    Z. R. Wasilewski Institute for Quantum Computing, University of Waterloo, Waterloo N2L 3G1, Canada Department of Physics, University of Waterloo, Waterloo N2L 3G1, Canada Northern Quantum Lights inc., Waterloo N2B 1N5, Canada Department of Electrical and Computer Engineering, University of Waterloo, Waterloo N2L 3G1, Canada Waterloo Institute for Nanotechnology, University of Waterloo, Waterloo N2L 3G1, Canada    J. Baugh [email protected] Institute for Quantum Computing, University of Waterloo, Waterloo N2L 3G1, Canada Department of Physics, University of Waterloo, Waterloo N2L 3G1, Canada Northern Quantum Lights inc., Waterloo N2B 1N5, Canada Department of Chemistry, University of Waterloo, Waterloo N2L 3G1, Canada Waterloo Institute for Nanotechnology, University of Waterloo, Waterloo N2L 3G1, Canada
Abstract

We report on transport characteristics of field effect two-dimensional electron gases (2DEG) in 24 nm wide indium arsenide surface quantum wells. High quality single-subband magnetotransport with clear quantized integer quantum Hall plateaus are observed to filling factor ν=2𝜈2\nu=2italic_ν = 2 in magnetic fields of up to B=18𝐵18B=18italic_B = 18 T, at electron densities up to 8×1011absentsuperscript1011\times 10^{11}× 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT /cm2. Peak mobility is 11,000 cm2/Vs at 2×1012absentsuperscript1012\times 10^{12}× 10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT /cm2. Large Rashba spin-orbit coefficients up to 124 meV\cdotÅ are obtained through weak anti-localization (WAL) measurements. Proximitized superconductivity is demonstrated in Nb-based superconductor-normal-superconductor (SNS) junctions, yielding 78--99 % interface transparencies from superconducting contacts fabricated ex-situ (post-growth), using two commonly-used experimental techniques for measuring transparencies. These transparencies are on a par with those reported for epitaxially-grown superconductors. These SNS junctions show characteristic voltages IcRnsubscript𝐼𝑐subscript𝑅nI_{c}R_{\textsc{n}}italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT n end_POSTSUBSCRIPT up to 870 μ𝜇\muitalic_μV and critical current densities up to 9.6 μ𝜇\muitalic_μA/μ𝜇\muitalic_μm, among the largest values reported for Nb-InAs SNS devices.

The last decade has seen spectacular progress in InAs/AlGaSb two-dimensional electron gases (2DEGs).[1, 2, 3, 4, 5, 6] This material system has joined the small club where the fractional quantum Hall effect can be routinely observed,[7, 8] in addition to 2DEGs in GaAs/AlGaAs,[9, 10] in AlAs/AlGaAs,[11] in graphene,[12, 13] in Si/SiGe,[14, 15], in Ge/SiGe,[16] in CdTe,[17, 18] and in ZnO/MgZnO.[19, 20] The highest mobility reported in InAs/AlGaSb is 2.4×\times×106 cm2/Vs,[21, 22] only exceeded by GaAs/AlGaAs,[23, 24] Ge/SiGe,[25] and AlAs/AlGaAs[11] material systems. The combination of high mobilities, strong spin orbit interactions (SOI), pinning of the Fermi level in the conduction band, small effective mass, and large Landé g-factor could make InAs/AlGaSb a strong candidate material system for topological quantum computing with Majorana zero modes.[26, 27, 28]

In the last decade, most efforts towards realizing Majorana fermions in a scalable platform have focused on surface quantum wells in the In(Ga)As/In0.8Al0.2As material system, where mobilities have significantly improved from 1×104absentsuperscript104\times 10^{4}× 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT cm2/Vs initially[26] to more than 1×105absentsuperscript105\times 10^{5}× 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT cm2/Vs recently.[29] However, in the context of topological quantum computing, the InAs/Al0.8Ga0.2Sb material system could offer possible advantages over the In(Ga)As/In0.8Al0.2As system, including better strain engineering,111The critical thickness of an InAs quantum well (QW) grown on Al0.8Ga0.2Sb is much larger (>>>24 nm; tensile strain) than on In0.8Al0.2As (7 nm; compressive strain) despite similar differences in lattice constant mismatch (similar-to\sim 8 pm) between InAs and either barrier material. higher electron densities,222Al0.8Ga0.2Sb has a larger conduction band offset (similar-to\sim 1.9 eV) relative to InAs than In0.8Al0.2As does (similar-to\sim 0.3 eV), thus providing a higher barrier next to the quantum well and allowing higher carrier densities to be achieved within a single 2D subband. Higher electron densities in turn can enable higher mobilities and stronger SOI than at lower electron densities. and higher mobilities.333A 2DEG hosted in a binary alloy quantum well instead of a ternary alloy quantum well would not suffer from alloy scattering, which only occurs in ternary alloys and is a significant mechanism limiting mobilities.

Furthermore, most efforts in this field have also centered on semiconductor-superconductor hybrid devices proximitized by “epitaxial aluminum,” grown directly on In(Ga)As/InAlAs heterostructures in the same growth chamber. This approach appeared to be the only way to reliably generate strong, “hard” superconducting gaps in the semiconductor,[33] as opposed to smaller, “soft” gaps, typically produced by superconductor contacts deposited ex-situ, post-growth.

In this Letter, we demonstrate gated 2DEGs in InAs/Al0.8Ga0.2Sb surface quantum wells, without parallel conduction from another conducting layer in magnetic fields up to 18 T and at electron densities up to 3×10123superscript10123\times 10^{12}3 × 10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT cm-2. Single-subband operation is demonstrated at lower 2DEG densities. Using SiO2 as dielectric yielded stable and reproducible gating operations all the way down to pinch-off. Rashba spin-orbit coefficients up to 124 meV\cdotÅ are obtained through weak anti-localization (WAL) measurements. Proximitized superconductivity is demonstrated in Nb-based superconductor-normal-superconductor (SNS) Josephson junctions, yielding deep gaps with up to unity transparencies from superconducting contacts fabricated entirely ex-situ (post-growth).

Two nominally identical heterostructures (G743 and G782) were grown by molecular beam epitaxy (MBE), with the following sequence of layers, starting from a GaSb (001) substrate: a 25 nm GaSb nucleation layer, a 800 nm Al0.80Ga0.20Sb0.93As0.07 quaternary buffer, a 20 nm Al0.8Ga0.2Sb bottom barrier, a 24 nm InAs quantum well, and a 6 nm In0.75Ga0.25As cap layer. There is no intentional do** anywhere in the heterostructure. Section I in the supplementary material provides additional details about MBE growth. Figure S1 from Section II in the supplementary material shows nextnano™ self-consistent simulations[34, 35] of bandstructure profiles, showing the extent of the 2DEG wavefunction within the InAs quantum well, and compares it to an In(Ga)As quantum well profile.

Hall bars were fabricated using standard optical lithography and wet-etching techniques, kee** all processes at or below a temperature of 150C to prevent the deterioration of device characteristics.[36, 37, 38] Ti/Au Ohmic contacts were deposited directly on the InGaAs cap layer, with typical resistances of 400--500 ΩΩ\Omegaroman_Ω at magnetic field B=0𝐵0B=0italic_B = 0, and 10 kΩΩ\Omegaroman_Ω at B=18𝐵18B=18italic_B = 18 T. Finally, 60 nm thick silicon dioxide (SiO2) or hafnium dioxide (HfO2) was deposited by atomic layer deposition (ALD) at 150C, followed by the deposition of a Ti/Au global top-gate. See Section III.A of the supplementary material for more details on Hall bar fabrication. Using standard four-terminal ac lock-in measurement techniques, all transport experiments were performed in either a pumped-4He cryostat or 3He/4He dilution refrigerator, with a base temperature of 1.6 K and 11 mK respectively.

The electron density at top-gate voltage V=g0{}_{g}=0start_FLOATSUBSCRIPT italic_g end_FLOATSUBSCRIPT = 0 in all gated Hall bars was significantly larger than the as-grown electron densities in ungated Hall bars. Figure 1 shows the pinch-off characteristics of Hall bars with different gate dielectrics. With SiO2, the pinch-off curves are stable and reproducible, overlap** perfectly when Vgsubscript𝑉𝑔V_{g}italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT is swept in the same direction, but showing some hysteresis when Vgsubscript𝑉𝑔V_{g}italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT is swept in opposite directions [see Fig. 1(a)]. After pinch-off, the 2DEG does not turn itself back on with time.[22] With HfO2 however, gating characteristics become unstable and non-reproducible near pinch-off [see Fig. 1(b)].

Refer to caption
Figure 1: Typical differential conductance G(Vg)=dI/dVsd𝐺subscript𝑉𝑔𝑑𝐼𝑑subscript𝑉𝑠𝑑G(V_{g})=dI/dV_{sd}italic_G ( italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ) = italic_d italic_I / italic_d italic_V start_POSTSUBSCRIPT italic_s italic_d end_POSTSUBSCRIPT (using 100 μ𝜇\muitalic_μV ac excitation) showing the turn-on voltage of a gated Hall bar with gate dielectric (a) SiO2 and (b) HfO2. In panel (a), eight traces are shown, four while increasing Vgsubscript𝑉𝑔V_{g}italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT (black) and four while decreasing Vgsubscript𝑉𝑔V_{g}italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT (gray). Only four traces are shown in panel (b), two going up and two going down in Vgsubscript𝑉𝑔V_{g}italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT. (c) Typical n2D(Vg)subscript𝑛2Dsubscript𝑉𝑔n_{\text{2D}}(V_{g})italic_n start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT ( italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ) of a Hall bar with SiO2. The 2DEG density increases linearly with Vgsubscript𝑉𝑔V_{g}italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT in all samples, and is reproducible along the linear traces. (d) Electron mobility in the same device as in (c).

At high conductances (G>100𝐺100G>100italic_G > 100 μ𝜇\muitalic_μS), both SiO2 and HfO2 produce stable and reproducible gating behavior. Figure 1(c) shows a typical electron density function n2D(Vg)subscript𝑛2Dsubscript𝑉𝑔n_{\text{2D}}(V_{g})italic_n start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT ( italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ) with SiO2. Section IV in the supplementary material shows representative gating and mobility characteristics from four gated Hall bars with SiO2 and HfO2. Figure 1(d) shows a peak in the transport mobility μ=11×103𝜇11superscript103\mu=11\times 10^{3}italic_μ = 11 × 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT cm2/Vs near 2DEG density n2D=2.0×1012subscript𝑛2D2.0superscript1012n_{\text{2D}}=2.0\times 10^{12}italic_n start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT = 2.0 × 10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT cm-2. At lower densities, ionized impurity scattering is most likely limiting mobility. The much higher 2DEG carrier density at Vg=0subscript𝑉𝑔0V_{g}=0italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = 0 (by a factor of 2--3) in gated Hall bars (i.e., after the dielectric deposition) relative to the as-grown 2DEG densities in ungated Hall bars strongly hints at a high density of charge traps forming at the semiconductor-oxide interface. To eliminate the risk of parallel conduction, often observed in similar structures when electrons are supplied by remote do**, the AlGaSb/InAs interface was engineered to have an AlAs character. Such interfaces are known to have high concentrations of As antisites, which act as double donors, supplying carriers to the 2DEG in the InAs quantum well.[39] The resulting high density of As2+ ions at the interface subjects the 2DEG to strong Coulomb scattering, reducing its mobility. However, with the top quantum well barrier being InGaAs a necessary element of our design such a tradeoff is well justified. Indeed, Lee et al.[5] demonstrated a severe reduction of mobility, from 650×103650E3650\text{\times}{10}^{3}start_ARG 650 end_ARG start_ARG times end_ARG start_ARG power start_ARG 10 end_ARG start_ARG 3 end_ARG end_ARG cm2/Vs down to 24×10324E324\text{\times}{10}^{3}start_ARG 24 end_ARG start_ARG times end_ARG start_ARG power start_ARG 10 end_ARG start_ARG 3 end_ARG end_ARG cm2/Vs, a value comparable with our results, by only replacing the AlGaSb top barrier layer with InGaAs in otherwise identical heterostructures in near-surface InAs quantum wells. Such reduction of mobility is likely due to the strong extent of the 2DEG wavefunction into the InGaAs barrier, with a lower conduction band offset than AlGaSb. At carrier densities above the mobility peak in Fig. 1(d), the observed decline in mobility with increasing 2DEG density can be attributed to interface roughness scattering and/or alloy scattering. This is likely because the electron wavefunction is drawn closer to the surface by the stronger electric field from the top-gate.[40, 41]

Refer to caption
Figure 2: (a) SdH and Hall traces for n2D=8.57×1011subscript𝑛2D8.57superscript1011n_{\text{2D}}=8.57\times 10^{11}italic_n start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT = 8.57 × 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT cm-2 (Vg=8.0subscript𝑉𝑔8.0V_{g}=-8.0italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = - 8.0 V). (b) Landau fan of SdH oscillations with densities ranging from n2D=7.3×1011subscript𝑛2D7.3superscript1011n_{\text{2D}}=7.3\times 10^{11}italic_n start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT = 7.3 × 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT cm-2 to n2D=1.8×1012subscript𝑛2D1.8superscript1012n_{\text{2D}}=1.8\times 10^{12}italic_n start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT = 1.8 × 10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT cm-2. White labels identify corresponding effective filling factors (directly obtained from the Hall resistance with ν=h/e2RH𝜈superscript𝑒2subscript𝑅𝐻\nu=h/e^{2}R_{H}italic_ν = italic_h / italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_R start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT, regardless of subband occupation), and dashed white lines identify populating Landau levels from the second 2D subband. (c) Weak anti-localization peak at n2D=(0.5;0.9;1.4;1.9)×1012subscript𝑛2D0.50.91.41.9superscript1012n_{\text{2D}}=(0.5;0.9;1.4;1.9)\times 10^{12}italic_n start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT = ( 0.5 ; 0.9 ; 1.4 ; 1.9 ) × 10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT cm-2. Colored circles are experimental points, black lines are fits to the ILP model, and n2Dsubscript𝑛2Dn_{\text{2D}}italic_n start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT is indicated in the upper right corner of each plot. (d) Spin orbit coefficient αsosubscript𝛼𝑠𝑜\alpha_{so}italic_α start_POSTSUBSCRIPT italic_s italic_o end_POSTSUBSCRIPT as a function of n2Dsubscript𝑛2Dn_{\text{2D}}italic_n start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT, extracted from fits of the WAL data to the ILP and HLN models. The dashed vertical line indicates where the 2ndnd{}^{\text{nd}}start_FLOATSUPERSCRIPT nd end_FLOATSUPERSCRIPT subband populates.

In a Hall bar with SiO2, Figure 2(a) shows Shubnikov-de-Haas (SdH) oscillations in the longitudinal resistivity ρxxsubscript𝜌𝑥𝑥\rho_{xx}italic_ρ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT and well-defined quantized quantum Hall plateaus in the Hall resistance Rhsubscript𝑅hR_{\textsc{h}}italic_R start_POSTSUBSCRIPT h end_POSTSUBSCRIPT. Quantum Hall plateaus occur at specific resistance values Rh=h/νe2subscript𝑅h𝜈superscript𝑒2R_{\textsc{h}}=h/\nu e^{2}italic_R start_POSTSUBSCRIPT h end_POSTSUBSCRIPT = italic_h / italic_ν italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT at filling factors ν=hn2D/eB𝜈subscript𝑛2𝐷𝑒𝐵\nu=hn_{2D}/eBitalic_ν = italic_h italic_n start_POSTSUBSCRIPT 2 italic_D end_POSTSUBSCRIPT / italic_e italic_B, where hhitalic_h is the Planck constant and e𝑒eitalic_e is the single electron charge. The presence of quantized Hall plateaus at ν=2,4𝜈24\nu=2,4italic_ν = 2 , 4 confirms the formation of a 2DEG. We note SdH oscillations are not, by themselves, proof of the presence of a 2DEG, since they are also observed in 3D conductors,[42] albeit with much smaller amplitudes. Despite a large Landé g𝑔gitalic_g-factor (gsimilar-tosuperscript𝑔absentg^{*}\simitalic_g start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ∼ 15), the spin-split ν=3𝜈3\nu=3italic_ν = 3 Hall plateau is only starting to be resolved at B10𝐵10B\approx 10italic_B ≈ 10 T, because of disorder. The visibility of spin splitting is dictated by (gμbBΓ)>kbTsuperscript𝑔subscript𝜇b𝐵Γsubscript𝑘b𝑇(g^{*}\mu_{\textsc{b}}B-\Gamma)>k_{\textsc{b}}T( italic_g start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_μ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT italic_B - roman_Γ ) > italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT italic_T, where μbsubscript𝜇b\mu_{\textsc{b}}italic_μ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT is the Bohr magneton, ΓΓ\Gammaroman_Γ is disorder associated with Landau level broadening, and kbsubscript𝑘bk_{\textsc{b}}italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT is the Boltzmann constant. Thus the late onset in field of spin splitting implies Γ9similar-toΓ9\Gamma\sim 9roman_Γ ∼ 9 meV in our samples. This same disorder Γ9similar-toΓ9\Gamma\sim 9roman_Γ ∼ 9 meV is also responsible for the very late onset of SdH oscillations (B3𝐵3B\approx 3italic_B ≈ 3 T), whose visibility is determined by (eB/mΓ)>kbTPlanck-constant-over-2-pi𝑒𝐵superscript𝑚Γsubscript𝑘b𝑇(\hbar eB/m^{*}-\Gamma)>k_{\textsc{b}}T( roman_ℏ italic_e italic_B / italic_m start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT - roman_Γ ) > italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT italic_T.

For B>10𝐵10B>10italic_B > 10 T, the SdH oscillation minimum at ν=2𝜈2\nu=2italic_ν = 2 reaches ρxx=0subscript𝜌𝑥𝑥0\rho_{xx}=0italic_ρ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT = 0 in Fig. 2(a), implying no parallel conduction from another conductive layer. This remains true until at least n2D=2.8×1012subscript𝑛2D2.8superscript1012n_{\text{2D}}=2.8\times 10^{12}italic_n start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT = 2.8 × 10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT cm-2 (Vg=0.5subscript𝑉𝑔0.5V_{g}=-0.5italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = - 0.5 V). However, this does not exclude the possibility of a second subband populating the 2DEG. Indeed, at n2D=7.3×1011subscript𝑛2D7.3superscript1011n_{\text{2D}}=7.3\times 10^{11}italic_n start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT = 7.3 × 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT cm-2 (Vg=8.5subscript𝑉𝑔8.5V_{g}=-8.5italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = - 8.5 V), there is a small mismatch (<<<3%) between the classical Hall density ntotal=B/eRhsubscript𝑛𝑡𝑜𝑡𝑎𝑙𝐵𝑒subscript𝑅hn_{total}=B/eR_{\textsc{h}}italic_n start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT = italic_B / italic_e italic_R start_POSTSUBSCRIPT h end_POSTSUBSCRIPT and the 2DEG density determined from the periodicity of Shubnikov-de-Haas oscillations versus inverse field, given by nSdH=2eh(1Bν+11Bν)1subscript𝑛𝑆𝑑𝐻2𝑒superscript1subscript𝐵𝜈11subscript𝐵𝜈1n_{SdH}=\frac{2e}{h}\left(\frac{1}{B_{\nu+1}}-\frac{1}{B_{\nu}}\right)^{-1}italic_n start_POSTSUBSCRIPT italic_S italic_d italic_H end_POSTSUBSCRIPT = divide start_ARG 2 italic_e end_ARG start_ARG italic_h end_ARG ( divide start_ARG 1 end_ARG start_ARG italic_B start_POSTSUBSCRIPT italic_ν + 1 end_POSTSUBSCRIPT end_ARG - divide start_ARG 1 end_ARG start_ARG italic_B start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. This mismatch grows significantly as the 2DEG density increases (until ntotal2nSdHsubscript𝑛𝑡𝑜𝑡𝑎𝑙2subscript𝑛𝑆𝑑𝐻n_{total}\approx 2n_{SdH}italic_n start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT ≈ 2 italic_n start_POSTSUBSCRIPT italic_S italic_d italic_H end_POSTSUBSCRIPT near Vg=0subscript𝑉𝑔0V_{g}=0italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = 0). Energy level crossings in the Landau fan from Figure 2(b) confirm the presence of another 2D subband, most likely corresponding to the two spin-split branches of the lowest Landau level. We estimate the second subband populates near n2D8×1011subscript𝑛2D8superscript1011n_{\text{2D}}\approx 8\times 10^{11}italic_n start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT ≈ 8 × 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT cm-2, which is consistent with similar reports of populating second subbands in 24 nm wide InAs/AlGaSb quantum wells.[21, 22] The “knee” observed in the pinch-off characteristics in Figure 1(a) is consistent with the population of a second subband in the 2DEG. We note that the second subband’s branch separating the regions labeled ν=4𝜈4\nu=4italic_ν = 4 and ν=5𝜈5\nu=5italic_ν = 5 at B10similar-to𝐵10B\sim 10italic_B ∼ 10 T in Fig. 2(b) does not appear to cross the ν=3𝜈3\nu=3italic_ν = 3 branch and does not extend into the ν=2𝜈2\nu=2italic_ν = 2 region. This is reminiscent of similar occurrences in the ring-like structures from Landau fans with Landau level crossings between the first and second subbands of GaAs/AlGaAs 2DEGs[43, 44, 45] and InAs/AlGaSb 2DEGs.[21] Nonetheless, our Landau fan, obtained by swee** the top-gate at magnetic field increments, showcases the reproducibility and stability of gating characteristics with dielectric SiO2.

Refer to caption
Figure 3: (a) Two overlapped I-V traces (solid) in a SNS device, where Isdsubscript𝐼𝑠𝑑I_{sd}italic_I start_POSTSUBSCRIPT italic_s italic_d end_POSTSUBSCRIPT is the source-drain dc current and Vdcsubscript𝑉𝑑𝑐V_{dc}italic_V start_POSTSUBSCRIPT italic_d italic_c end_POSTSUBSCRIPT is the four-terminal dc voltage drop across the SNS junction. The dashed lines are the extrapolated normal resistance Rnsubscript𝑅nR_{\textsc{n}}italic_R start_POSTSUBSCRIPT n end_POSTSUBSCRIPT and their intercepts at Vdc=0subscript𝑉𝑑𝑐0V_{dc}=0italic_V start_POSTSUBSCRIPT italic_d italic_c end_POSTSUBSCRIPT = 0 are the excess currents Iexsubscript𝐼𝑒𝑥I_{ex}italic_I start_POSTSUBSCRIPT italic_e italic_x end_POSTSUBSCRIPT. (upper inset) Zoom-in of the superconducting transition, showing critical current Icsubscript𝐼𝑐I_{c}italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. Two overlap** traces are shown (black, grey), swee** Isdsubscript𝐼𝑠𝑑I_{sd}italic_I start_POSTSUBSCRIPT italic_s italic_d end_POSTSUBSCRIPT in opposite directions. (lower inset) Sharp, deep gap in the four-terminal differential resistance of device SNS-4. (b) Normalized four-terminal differential conductance dI/dV𝑑𝐼𝑑𝑉dI/dVitalic_d italic_I / italic_d italic_V, identifying MAR peaks in a SNS device. (c) Temperature dependence of Icsubscript𝐼𝑐I_{c}italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT of a SNS device, where crosses are experimental data and dashed lines are fits to Eqn. (3) for different values of 𝒯𝒯\mathcal{T}caligraphic_T.

The curvature of the energy level associated with ν=2𝜈2\nu=2italic_ν = 2 near B=17𝐵17B=17italic_B = 17 T at high densities in Fig. 2(b) is consistent with strong spin orbit interactions in the system. To quantify the strength of spin-orbit interactions, the Rashba coefficient αsosubscript𝛼𝑠𝑜\alpha_{so}italic_α start_POSTSUBSCRIPT italic_s italic_o end_POSTSUBSCRIPT was determined from fits to the weak antilocalization (WAL) conductivity peak ΔσxxΔsubscript𝜎𝑥𝑥\Delta\sigma_{xx}roman_Δ italic_σ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT [see Fig. 2(c)] using the Hikami-Larkin-Nagaoka (HLN) and Iordanskii-Lyanda-Geller-Pikus (ILP) models,[46, 47] where ΔσxxΔsubscript𝜎𝑥𝑥\Delta\sigma_{xx}roman_Δ italic_σ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT = σxx(B)σxx(0)subscript𝜎𝑥𝑥𝐵subscript𝜎𝑥𝑥0\sigma_{xx}(B)-\sigma_{xx}(0)italic_σ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ( italic_B ) - italic_σ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ( 0 ), σxx(B)subscript𝜎𝑥𝑥𝐵\sigma_{xx}(B)italic_σ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ( italic_B ) is the field-dependent conductivity, and σxx(0)subscript𝜎𝑥𝑥0\sigma_{xx}(0)italic_σ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT ( 0 ) is a constant background conductivity. Figure 2(d) summarizes the resulting fit values of αsosubscript𝛼𝑠𝑜\alpha_{so}italic_α start_POSTSUBSCRIPT italic_s italic_o end_POSTSUBSCRIPT with both models, ranging from 16 to 124 meV\cdotÅ, roughly linear with n2Dsubscript𝑛2𝐷n_{2D}italic_n start_POSTSUBSCRIPT 2 italic_D end_POSTSUBSCRIPT. These are consistent with published values for 2DEGs in InAs/AlGaSb[2, 4, 22] and In(Ga)As/InAlAs.[48, 29, 49, 50] Some of these literature values (and ours) span both the single subband and multi-subband regimes. Figure S6 from Section IV in the supplementary material presents a detailed comparison of our αso(n2D)subscript𝛼𝑠𝑜subscript𝑛2D\alpha_{so}(n_{\text{2D}})italic_α start_POSTSUBSCRIPT italic_s italic_o end_POSTSUBSCRIPT ( italic_n start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT ) with literature. Both the HLN and ILP models are only applicable at magnetic fields where the mean free path esubscript𝑒\ell_{e}roman_ℓ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is smaller than the magnetic length msubscript𝑚\ell_{m}roman_ℓ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT, leading to the condition B<Btr𝐵subscript𝐵𝑡𝑟B<B_{tr}italic_B < italic_B start_POSTSUBSCRIPT italic_t italic_r end_POSTSUBSCRIPT where Btr=/2ee2subscript𝐵𝑡𝑟Planck-constant-over-2-pi2𝑒superscriptsubscript𝑒2B_{tr}=\hbar/2e\ell_{e}^{2}italic_B start_POSTSUBSCRIPT italic_t italic_r end_POSTSUBSCRIPT = roman_ℏ / 2 italic_e roman_ℓ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the transport field. In our sample, one of the distinctive features of the conductivity used for fitting (the minima on either side of the WAL peak) occurs well beyond Btrsubscript𝐵𝑡𝑟B_{tr}italic_B start_POSTSUBSCRIPT italic_t italic_r end_POSTSUBSCRIPT for the data point at n2D=1.9×1012subscript𝑛2𝐷1.9E12n_{2D}=$1.9\text{\times}{10}^{12}$italic_n start_POSTSUBSCRIPT 2 italic_D end_POSTSUBSCRIPT = start_ARG 1.9 end_ARG start_ARG times end_ARG start_ARG power start_ARG 10 end_ARG start_ARG 12 end_ARG end_ARG cm-2, and the latter should therefore not be considered reliable.

Device width gap Icsubscript𝐼𝑐I_{c}italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT Iexsubscript𝐼𝑒𝑥I_{ex}italic_I start_POSTSUBSCRIPT italic_e italic_x end_POSTSUBSCRIPT Rnsubscript𝑅nR_{\textsc{n}}italic_R start_POSTSUBSCRIPT n end_POSTSUBSCRIPT ΔmarsubscriptΔmar\Delta_{\textsc{mar}}roman_Δ start_POSTSUBSCRIPT mar end_POSTSUBSCRIPT 𝒯𝒯\mathcal{T}caligraphic_T
ID (μ𝜇\muitalic_μm) (nm) (μ𝜇\muitalic_μA) (μ𝜇\muitalic_μA) (ΩΩ\Omegaroman_Ω) (meV) (%)
SNS-1 10 200 82 158 8.8 1.13 78 ±plus-or-minus\pm± 5
SNS-2 10 200 89 183 8.6 1.16 81 ±plus-or-minus\pm± 4
SNS-3 3 120 29 58 30 1.13 85 ±plus-or-minus\pm± 2
SNS-4 3 120 25 54 33 1.05 87 ±plus-or-minus\pm± 3
SNS-5 3 400 5 24 53 0.91 82 ±plus-or-minus\pm± 2
SNS-6 3 400 2 26 55 0.94 84 ±plus-or-minus\pm± 3
Table 1: List of SNS devices.

Three nominally identical pairs of superconductor-normal-superconductor (SNS) devices [see Table 1] were fabricated from wafers G743 and G782, using standard optical lithography, electron beam lithography, and wet-etching techniques, kee** all processes at or below a temperature of 180C to prevent the deterioration of device characteristics.[36, 37, 38] Ti/Nb (2/80 nm) ohmic contacts were deposited directly on the InGaAs cap layer by sputtering, after Ar ion milling the contact areas for 6.5 minutes at 50 Watts. Immediately prior to loading samples in the Nb deposition system, the contact areas were treated with sulfur passivation.[51] The latter is designed to etch away native oxides, prevent further surface oxidation during transfer in air (similar-to\sim 30 s) from the wetbench to the deposition chamber, and possibly dope the surface.[52, 53, 54] Four devices (SNS-3, … , SNS-6) were 3 μ𝜇\muitalic_μm wide, of which two (two) were fabricated with a gap L=120𝐿120L=120italic_L = 120 nm (L=400𝐿400L=400italic_L = 400 nm) between the Ti/Nb contacts. Devices SNS-1 and SNS-2 were fabricated on wafer G743 rather than G782, with a width W=10𝑊10W=10italic_W = 10 μ𝜇\muitalic_μm and a gap of L=200𝐿200L=200italic_L = 200 nm. Note the gap region is not gated in any device. See Section III.B of the supplementary material for more details on sample fabrication.

The sputter-deposited Nb had a critical temperature Tc=8.1subscript𝑇𝑐8.1T_{c}=8.1italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 8.1 K, yielding the superconducting energy gap ΔNb=1.76kbTc=1.23subscriptΔNb1.76subscript𝑘bsubscript𝑇𝑐1.23\Delta_{\text{Nb}}=1.76\,k_{\textsc{b}}T_{c}=1.23roman_Δ start_POSTSUBSCRIPT Nb end_POSTSUBSCRIPT = 1.76 italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 1.23 meV. Using this value, the coherence length in the proximitized semiconductor of our SNS junctions is ξb=vf/2ΔNb=220subscript𝜉𝑏Planck-constant-over-2-pisubscript𝑣f2subscriptΔNb220\xi_{b}=\hbar v_{\textsc{f}}/2\Delta_{\text{Nb}}=220italic_ξ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = roman_ℏ italic_v start_POSTSUBSCRIPT f end_POSTSUBSCRIPT / 2 roman_Δ start_POSTSUBSCRIPT Nb end_POSTSUBSCRIPT = 220 nm (246 nm) in the ballistic regime for wafer G782 (G743). Using the as-grown 2DEG density and mobility, the mean free path is e=μ2πn2D/e=244subscript𝑒Planck-constant-over-2-pi𝜇2𝜋subscript𝑛2D𝑒244\ell_{e}=\hbar\mu\sqrt{2\pi n_{\text{2D}}}/e=244roman_ℓ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = roman_ℏ italic_μ square-root start_ARG 2 italic_π italic_n start_POSTSUBSCRIPT 2D end_POSTSUBSCRIPT end_ARG / italic_e = 244 nm (175 nm) for wafer G782 (G743). Devices SNS-1 and SNS-2 are in the diffusive, short junction regime, since e<L<ξbsubscript𝑒𝐿subscript𝜉𝑏\ell_{e}<L<\xi_{b}roman_ℓ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT < italic_L < italic_ξ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT. Devices SNS-3 and SNS-4 are in the quasi-ballistic, short junction regime, since L<e,ξb𝐿subscript𝑒subscript𝜉𝑏L<\ell_{e},\xi_{b}italic_L < roman_ℓ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT , italic_ξ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT. Devices SNS-5 and SNS-6 are in the diffusive, long junction regime, since L>e,ξb𝐿subscript𝑒subscript𝜉𝑏L>\ell_{e},\xi_{b}italic_L > roman_ℓ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT , italic_ξ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT. All devices are in the dirty regime, since eξbsubscript𝑒subscript𝜉𝑏\ell_{e}\lessapprox\xi_{b}roman_ℓ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ⪅ italic_ξ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT.

All six SNS devices demonstrated a supercurrent. Figure 3(a) shows typical four-terminal dc I-V traces in a quasi-ballistic device. Remarkably, the I-V traces do not display any hysterectic behavior, as emphasized in the upper inset of Fig. 3(a); this is true for all six SNS devices reported here. All devices are well thermalized and do not adversely suffer from local Joule heating effects,[55, 56] due to the dilution refrigerator (Kelvinox TLM from Oxford Instruments) used in the experiments, where samples and wiring are completely immersed in the 3He/4He mixture. The lower inset in Fig. 3(a) demonstrates a deep, sharply-defined superconductivity gap.

Figure 3(b) plots the four-terminal differential conductance G=dI/dV𝐺𝑑𝐼𝑑𝑉G=dI/dVitalic_G = italic_d italic_I / italic_d italic_V normalized by the conductance in the normal regime Gn=1/Rnsubscript𝐺n1subscript𝑅nG_{\textsc{n}}=1/R_{\textsc{n}}italic_G start_POSTSUBSCRIPT n end_POSTSUBSCRIPT = 1 / italic_R start_POSTSUBSCRIPT n end_POSTSUBSCRIPT, with periodic peaks arising from multiple Andreev reflections (MAR). The peak periodicity in Vdcsubscript𝑉𝑑𝑐V_{dc}italic_V start_POSTSUBSCRIPT italic_d italic_c end_POSTSUBSCRIPT is described by eVn=2Δmar/n𝑒subscript𝑉𝑛2subscriptΔmar𝑛eV_{n}=2\Delta_{\textsc{mar}}/nitalic_e italic_V start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 2 roman_Δ start_POSTSUBSCRIPT mar end_POSTSUBSCRIPT / italic_n where n𝑛nitalic_n is an integer. Up to six peaks are observed in Fig. 3(b), with the first three (n=1,,3𝑛13n=1,...,3italic_n = 1 , … , 3) following the MAR sequence with Δmar=1.05subscriptΔmar1.05\Delta_{\textsc{mar}}=1.05roman_Δ start_POSTSUBSCRIPT mar end_POSTSUBSCRIPT = 1.05 meV. This value is only slightly smaller than ΔNb=1.23subscriptΔNb1.23\Delta_{\text{Nb}}=1.23roman_Δ start_POSTSUBSCRIPT Nb end_POSTSUBSCRIPT = 1.23 meV, and is consistent with a high-quality SNS junction. Section V in the supplementary material shows I-V traces and MAR analysis for all six SNS devices reported here (Figs. S8--S10).

The ratio eIcRn/Δmar𝑒subscript𝐼𝑐subscript𝑅nsubscriptΔmareI_{c}R_{\textsc{n}}/\Delta_{\textsc{mar}}italic_e italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT n end_POSTSUBSCRIPT / roman_Δ start_POSTSUBSCRIPT mar end_POSTSUBSCRIPT and critical current density Ic/Wsubscript𝐼𝑐𝑊I_{c}/Witalic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_W are often used as a figure of merit for SNS junctions, with typical ranges 0.02--0.15 and 0.3--1.6 μ𝜇\muitalic_μA/μ𝜇\muitalic_μm respectively for planar In(Ga)As quantum wells proximitized to Nb.[57, 58, 59, 60] For the four devices in the short junction regime (SNS-1, …, SNS-4 in Table 1), the ratio eIcRn/Δmar𝑒subscript𝐼𝑐subscript𝑅nsubscriptΔmareI_{c}R_{\textsc{n}}/\Delta_{\textsc{mar}}italic_e italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT n end_POSTSUBSCRIPT / roman_Δ start_POSTSUBSCRIPT mar end_POSTSUBSCRIPT ranges from 0.64 to 0.79, and the critical current density Ic/Wsubscript𝐼𝑐𝑊I_{c}/Witalic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_W ranges from 8.2 to 9.6 μ𝜇\muitalic_μA/μ𝜇\muitalic_μm. Both set of numbers indicate a very strong proximity effect from the Nb parent superconductor to the proximitized InAs 2DEG in the gap region.

A high-quality SNS junction is characterized by a high interface transparency 𝒯𝒯\mathcal{T}caligraphic_T at the SN interfaces, which corresponds to a high probability of Andreev reflection.[61, 62, 63, 64] It can be analytically calculated with the generalized Octavion-Tinkham–Blonder-Klapwijk (OTBK) model:[65]

eIexRnΔmar𝑒subscript𝐼𝑒𝑥subscript𝑅nsubscriptΔmar\displaystyle\frac{eI_{ex}R_{\textsc{n}}}{\Delta_{\textsc{mar}}}divide start_ARG italic_e italic_I start_POSTSUBSCRIPT italic_e italic_x end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT n end_POSTSUBSCRIPT end_ARG start_ARG roman_Δ start_POSTSUBSCRIPT mar end_POSTSUBSCRIPT end_ARG =\displaystyle== 2(1+2Z2)tanh1(2Z~)Z(1+Z2)(1+6Z2+4Z4)43212superscript𝑍2superscript12~𝑍𝑍1superscript𝑍216superscript𝑍24superscript𝑍443\displaystyle\frac{2(1+2Z^{2})\tanh^{-1}(2\tilde{Z})}{Z\sqrt{(1+Z^{2})(1+6Z^{2% }+4Z^{4})}}-\frac{4}{3}\qquaddivide start_ARG 2 ( 1 + 2 italic_Z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) roman_tanh start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( 2 over~ start_ARG italic_Z end_ARG ) end_ARG start_ARG italic_Z square-root start_ARG ( 1 + italic_Z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( 1 + 6 italic_Z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 italic_Z start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) end_ARG end_ARG - divide start_ARG 4 end_ARG start_ARG 3 end_ARG (1)
and 𝒯and 𝒯\displaystyle\text{and~{}~{}}\mathcal{T}and caligraphic_T =\displaystyle== (1+Z2)1superscript1superscript𝑍21\displaystyle(1+Z^{2})^{-1}( 1 + italic_Z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (2)

where Z~=Z(1+Z2)/(1+6Z2+4Z4)~𝑍𝑍1superscript𝑍216superscript𝑍24superscript𝑍4\tilde{Z}=Z\sqrt{(1+Z^{2})/(1+6Z^{2}+4Z^{4})}over~ start_ARG italic_Z end_ARG = italic_Z square-root start_ARG ( 1 + italic_Z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) / ( 1 + 6 italic_Z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 italic_Z start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) end_ARG, Iexsubscript𝐼𝑒𝑥I_{ex}italic_I start_POSTSUBSCRIPT italic_e italic_x end_POSTSUBSCRIPT is the excess current obtained from an I-V trace [see Fig. 3(a)], ΔmarsubscriptΔmar\Delta_{\textsc{mar}}roman_Δ start_POSTSUBSCRIPT mar end_POSTSUBSCRIPT is the superconducting gap determined from the MAR periodicity [see Fig. 3(b)], and Z𝑍Zitalic_Z is a dimensionless scattering parameter related to the barrier height at the SN interface. The formalism above explicitly assumes that both SN interfaces in the SNS junction are symmetric. To calculate the SN transparency, Z𝑍Zitalic_Z is first used as a free variable in Eq. (1) to fit the experimental data (Iexsubscript𝐼𝑒𝑥I_{ex}italic_I start_POSTSUBSCRIPT italic_e italic_x end_POSTSUBSCRIPT, Rnsubscript𝑅nR_{\textsc{n}}italic_R start_POSTSUBSCRIPT n end_POSTSUBSCRIPT, ΔmarsubscriptΔmar\Delta_{\textsc{mar}}roman_Δ start_POSTSUBSCRIPT mar end_POSTSUBSCRIPT). Once a value for Z𝑍Zitalic_Z is found, it is then used in Eq. (2) to calculate 𝒯𝒯\mathcal{T}caligraphic_T.

Using only the data at base temperature and Eqns. (1)--(2), all six SNS devices in Table 1 show consistently high transparencies 𝒯𝒯\mathcal{T}caligraphic_T ranging in values from 78% to 87%. For context, reported interface transparencies of epitaxial aluminum to In(Ga)As surface quantum wells range from 75% to 97%,[66, 67, 68] when using the same experimental measurement method.

Another method for measuring 𝒯𝒯\mathcal{T}caligraphic_T involves the temperature dependence of Icsubscript𝐼𝑐I_{c}italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT to be fit to the generalized Kulik-Omelyanchuk (KO) model:[69]

Ic(T,ϕ,𝒯)subscript𝐼𝑐𝑇italic-ϕ𝒯\displaystyle I_{c}(T,\phi,\mathcal{T})italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_T , italic_ϕ , caligraphic_T ) =\displaystyle== πΔ(T)sin(ϕ)2eRn1𝒯sin2(ϕ/2)𝜋Δ𝑇italic-ϕ2𝑒subscript𝑅n1𝒯superscript2italic-ϕ2\displaystyle\frac{\pi\Delta(T)\sin(\phi)}{2eR_{\textsc{n}}\sqrt{1-\mathcal{T}% \sin^{2}(\phi/2)}}\qquad\qquad\qquad\qquaddivide start_ARG italic_π roman_Δ ( italic_T ) roman_sin ( italic_ϕ ) end_ARG start_ARG 2 italic_e italic_R start_POSTSUBSCRIPT n end_POSTSUBSCRIPT square-root start_ARG 1 - caligraphic_T roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ϕ / 2 ) end_ARG end_ARG (3)
×tanh(Δ(T)1𝒯sin2(ϕ/2)2kbT)absentΔ𝑇1𝒯superscript2italic-ϕ22subscript𝑘b𝑇\displaystyle\times\tanh\left(\frac{\Delta(T)\sqrt{1-\mathcal{T}\sin^{2}(\phi/% 2)}}{2k_{\textsc{b}}T}\right)× roman_tanh ( divide start_ARG roman_Δ ( italic_T ) square-root start_ARG 1 - caligraphic_T roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ϕ / 2 ) end_ARG end_ARG start_ARG 2 italic_k start_POSTSUBSCRIPT b end_POSTSUBSCRIPT italic_T end_ARG )

where ϕitalic-ϕ\phiitalic_ϕ is the superconducting phase picked up from Andreev reflections at the semiconductor-superconductor interface, and Δ(T)Δ𝑇\Delta(T)roman_Δ ( italic_T ) is the temperature-dependent superconducting gap, which we model with the BCS theory relation Δ(T)=Δmar1(T/Tc)2Δ𝑇subscriptΔmar1superscript𝑇subscript𝑇𝑐2\Delta(T)=\Delta_{\textsc{mar}}\sqrt{1-(T/T_{c})^{2}}roman_Δ ( italic_T ) = roman_Δ start_POSTSUBSCRIPT mar end_POSTSUBSCRIPT square-root start_ARG 1 - ( italic_T / italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG with Tc=6.0subscript𝑇𝑐6.0T_{c}=6.0italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 6.0 K. The conventional fitting procedure involves finding the value of ϕitalic-ϕ\phiitalic_ϕ that maximizes Icsubscript𝐼𝑐I_{c}italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for each temperature.[70, 71, 72, 73]

Figure 3(c) shows the temperature dependence and fit of Icsubscript𝐼𝑐I_{c}italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT on device SNS-1, where the experimental data is normalized by the maximum current for 𝒯=100𝒯100\mathcal{T}=100caligraphic_T = 100% predicted by Eqn. (3). The experiment yielded a transparency of 𝒯=97.8𝒯97.8\mathcal{T}=97.8caligraphic_T = 97.8%. The same experiment was also performed on device SNS-3 (shown in Fig. S11 of the supplementary material), which yielded 𝒯=99.5𝒯99.5\mathcal{T}=99.5caligraphic_T = 99.5%. Both values are significantly larger than the transparencies obtained by the MAR analysis (𝒯=78±5𝒯plus-or-minus785\mathcal{T}=78\pm 5caligraphic_T = 78 ± 5 %, 85±2plus-or-minus85285\pm 285 ± 2%) in the same sample [see Table 1].

Regardless of which experimental method is used for measuring 𝒯𝒯\mathcal{T}caligraphic_T, our main result is that semiconductor-superconductor interface transparencies achieved in samples with a superconductor deposited post-growth can be comparable to those from “epitaxial” superconductors. This could dramatically expand the repertoire of possible superconductors available for realizing semiconductor-superconductor hybrid devices, potentially impacting the fields of topological quantum computing, superconducting qubits (via the gatemon design), and superconducting logic circuits.

In conclusion, we presented Josephson SNS junctions fabricated with ex-situ sputtered Nb contacts to 2DEGs hosted in InAs/AlGaSb surface quantum wells. We observed consistent and highly transparent interfaces with values of 𝒯𝒯\mathcal{T}caligraphic_T ranging 78--99%. Post-growth superconducting contacts to InAs quantum wells can be a viable method on a par with epitaxial aluminum systems, and do not depend on unreasonably stringent fabrication parameters in the InAs material system.

supplementary material

The supplementary material contains additional information on MBE growth, bandstructure profiles, sample fabrication, characterization of Hall bars, and I-V/MAR traces of SNS junctions.

E.A.B., F.S., and A.E. contributed equally to this paper. The authors thank Kaveh Gharavi, Sean Walker, and Christine Nicoll for useful discussions. E.A.B. acknowledges support from a Mike and Ophelia Lazaridis Fellowship. This research was undertaken thanks in part to funding from the Canada First Research Excellence Fund (Transformative Quantum Technologies) and the Natural Sciences and Engineering Research Council (NSERC) of Canada. The University of Waterloo’s QNFCF facility was used for this work. This infrastructure would not be possible without the significant contributions of CFREF-TQT, CFI, ISED, the Ontario Ministry of Research and Innovation, and Mike and Ophelia Lazaridis. Their support is gratefully acknowledged.

References

  • Shojaei et al. [2015] B. Shojaei, A. McFadden, J. Shabani, B. D. Schultz, and C. J. Palmstrøm, “Studies of scattering mechanisms in gate tunable InAs/(Al,Ga)Sb two dimensional electron gases,” Appl. Phys. Lett. 106, 222101 (2015).
  • Shojaei et al. [2016a] B. Shojaei, P. J. J. O’Malley, J. Shabani, P. Roushan, B. D. Schultz, R. M. Lutchyn, C. Nayak, J. M. Martinis, and C. J. Palmstrøm, “Demonstration of gate control of spin splitting in a high-mobility InAs/AlSb two-dimensional electron gas,” Phys. Rev. B 93, 075302 (2016a).
  • Shojaei et al. [2016b] B. Shojaei, A. C. C. Drachmann, M. Pendharkar, D. J. Pennachio, M. P. Echlin, P. G. Callahan, S. Kraemer, T. M. Pollock, C. M. Marcus, and C. J. Palmstrøm, “Limits to mobility in InAs quantum wells with nearly lattice-matched barriers,” Phys. Rev. B 94, 245306 (2016b).
  • Hatke et al. [2017] A. T. Hatke, T. Wang, C. Thomas, G. C. Gardner, and M. J. Manfra, “Mobility in excess of 106 cm2/Vs in InAs quantum wells grown on lattice mismatched InP substrates,” Appl. Phys. Lett. 111, 142106 (2017).
  • Lee et al. [2019] J. S. Lee, B. Shojaei, M. Pendharkar, M. Feldman, K. Mukherjee, and C. J. Palmstrøm, “Contribution of top barrier materials to high mobility in near-surface InAs quantum wells grown on GaSb(001),” Phys. Rev. Mater. 3, 014603 (2019).
  • Mittag et al. [2021] C. Mittag, J. V. Koski, M. Karalic, C. Thomas, A. Tuaz, A. T. Hatke, G. C. Gardner, M. J. Manfra, J. Danon, T. Ihn, and K. Ensslin, “Few-electron single and double quantum dots in an InAs two-dimensional electron gas,” PRX Quantum 2, 010321 (2021).
  • Ma et al. [2017] M. K. Ma, M. S. Hossain, K. A. V. Rosales, H. Deng, T. Tschirky, W. Wegscheider, and M. Shayegan, “Observation of fractional quantum hall effect in an InAs quantum well,” Phys. Rev. B 96, 241301(R) (2017).
  • Komatsu et al. [2022] S. Komatsu, H. Irie, T. Akiho, T. Nojima, T. Akazaki, and K. Muraki, “Gate tuning of fractional quantum Hall states in an InAs two-dimensional electron gas,” Phys. Rev. B 105, 075305 (2022).
  • Pan et al. [2008] W. Pan, J. S. Xia, H. L. Störmer, D. C. Tsui, C. Vicente, E. D. Adams, N. S. Sullivan, L. N. Pfeiffer, K. W. Baldwin, and K. W. West, “Experimental studies of the fractional quantum hall effect in the first excited Landau level,” Phys. Rev. B 77, 075307 (2008).
  • Kleinbaum et al. [2020] E. Kleinbaum, H. Li, N. Deng, G. C. Gardner, M. J. Manfra, and G. A. Csáthy, “Disorder broadening of even-denominator fractional quantum Hall states in the presence of a short-range alloy potential,” Phys. Rev. B 102, 035140 (2020).
  • Chung et al. [2018] Y. J. Chung, K. A. V. Rosales, H. Deng, K. W. Baldwin, K. W. West, M. Shayegan, and L. N. Pfeiffer, “Multivalley two-dimensional electron system in an AlAs quantum well with mobility exceeding 2×\times×106 cm2/Vs,”  2, 071001(R) (2018).
  • Bolotin et al. [2009] K. I. Bolotin, F. Ghahari, M. D. Shulman, H. L. Stormer, and P. Kim, “Observation of the fractional quantum Hall effect in graphene,” Nature 462, 196 (2009).
  • Dean et al. [2011] C. R. Dean, A. F. Young, P. Cadden-Zimansky, L. Wang, H. Ren, K. Watanabe, T. Taniguchi, P. Kim, J. Hone, and K. L. Shepard, “Multicomponent fractional quantum hall effect in graphene,” Nat. Phys. 7, 693 (2011).
  • Lai et al. [2004] K. Lai, W. Pan, D. C. Tsui, S. Lyon, M. Mühlberger, and F. Schäffler, “Two-flux composite fermion series of the fractional quantum Hall states in strained Si,” Phys. Rev. Lett. 93, 156805 (2004).
  • Lu et al. [2012] T. M. Lu, W. Pan, D. C. Tsui, C.-H. Lee, and C. W. Liu, “Fractional quantum Hall effect of two-dimensional electrons in high-mobility Si/SiGe field-effect transistors,” Phys. Rev. B 85, 121307(R) (2012).
  • Mironov et al. [2016] O. A. Mironov, N. d’Ambrumenil, A. Dobbie, D. R. Leadley, A. V. Suslov, and E. Green, “Fractional quantum Hall states in a Ge quantum well,” Phys. Rev. Lett. 116, 176802 (2016).
  • Piot et al. [2010] B. A. Piot, J. Kunc, M. Potemski, D. K. Maude, C. Betthausen, A. Vogl, D. Weiss, G. Karczewski, and T. Wojtowicz, “Fractional quantum Hall effect in CdTe,” Phys. Rev. B 82, 081307(R) (2010).
  • Betthausen et al. [2014] C. Betthausen, P. Giudici, A. Iankilevitch, C. Preis, V. Kolkovsky, M. Wiater, G. Karczewski, B. A. Piot, J. Kunc, M. Potemski, T. Wojtowicz, and D. Weiss, “Fractional quantum Hall effect in a dilute magnetic semiconductor,” Phys. Rev. B 90, 115302 (2014).
  • Tsukazaki1 et al. [2010] A. Tsukazaki1, S. Akasaka, K. Nakahara, Y. Ohno, H. Ohno, D. Maryenko, A. Ohtomo, and M. Kawasaki, “Observation of the fractional quantum Hall effect in an oxide,” Nat. Mater. 9, 889 (2010).
  • Falson and Kawasaki [2018] J. Falson and M. Kawasaki, “A review of the quantum Hall effects in MgZnO/ZnO heterostructures,” Rep. Prog. Phys. 81, 056501 (2018).
  • Tschirky et al. [2017] T. Tschirky, S. Mueller, C. A. Lehner, S. Fält, T. Ihn, K. Ensslin, and W. Wegscheider, “Scattering mechanisms of highest-mobility InAs/AlxGa1-xSb quantum wells,” Phys. Rev. B 95, 115304 (2017).
  • Thomas et al. [2018] C. Thomas, A. T. Hatke, A. Tuaz, R. Kallaher, T. Wu, T. Wang, R. E. Diaz, G. C. Gardner, M. A. Capano, and M. J. Manfra, “High-mobility InAs 2DEGs on GaSb substrates: A platform for mesoscopic quantum transport,” Phys. Rev. Mater. 2, 104602 (2018).
  • Umansky et al. [2009] V. Umansky, M. Heiblum, Y. Levinson, J. Smet, J. Nübler, and M. Dolev, “MBE growth of ultra-low disorder 2DEG with mobility exceeding 35×10635superscript10635\times 10^{6}35 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPTcm2/Vs,” J. Cryst. Growth 311, 1658 (2009).
  • Chung et al. [2021] Y. J. Chung, K. A. V. Rosales, K. W. Baldwin, P. T. Madathil, K. W. West, M. Shayegan, and L. N. Pfeiffer, “Ultra-high-quality two-dimensional electron systems,” Nat. Mater. 20, 632 (2021).
  • Myronov et al. [2023] M. Myronov, J. Kycia, P. Waldron, W. Jiang, P. Barrios, A. Bogan, P. Coleridge, and S. Studenikin, “Holes outperform electrons in group IV semiconductor materials,” Small Sci. 3, 2200094 (2023).
  • Shabani et al. [2016] J. Shabani, M. Kjærgaard, H. J. Suominen, Y. Kim, F. Nichele, K. Pakrouski, T. Stankevic, R. M. Lutchyn, P. Krogstrup, R. Feidenhans, S. Kraemer, C. Nayak, M. Troyer, C. M. Marcus, and C. J. Palmstrøm, “Two-dimensional epitaxial superconductor-semiconductor heterostructures: A platform for topological superconducting networks,” Phys. Rev. B 93, 155402 (2016).
  • Karzig et al. [2017] T. Karzig, C. Knapp, R. M. Lutchyn, P. Bonderson, M. B. Hastings, C. Nayak, J. Alicea, K. Flensberg, S. Plugge, Y. Oreg, C. M. Marcus, and M. H. Freedman, “Scalable designs for quasiparticle-poisoning-protected topological quantum computation with Majorana zero modes,” Phys. Rev. B 95, 235305 (2017).
  • Ke et al. [2019] C. T. Ke, C. M. Moehle, F. K. de Vries, C. Thomas, S. Metti, C. R. Guinn, R. Kallaher, M. Lodari, G. Scappucci, T. Wang, R. E. Diaz, G. C. Gardner, M. J. Manfra, and S. Goswami, “Ballistic superconductivity and tunable π𝜋\piitalic_π–junctions in InSb quantum wells,” Nat. Commun. 10, 3764 (2019).
  • Zhang et al. [2023] T. Zhang, T. Lindemann, G. C. Gardner, S. Gronin, T. Wu, and M. J. Manfra, “Mobility exceeding 100,000 cm2/Vs in modulation-doped shallow InAs quantum wells coupled to epitaxial aluminium,” Phys. Rev. Mater. 7, 056201 (2023).
  • Note [1] The critical thickness of an InAs quantum well (QW) grown on Al0.8Ga0.2Sb is much larger (>>>24 nm; tensile strain) than on In0.8Al0.2As (7 nm; compressive strain) despite similar differences in lattice constant mismatch (similar-to\sim 8 pm) between InAs and either barrier material.
  • Note [2] Al0.8Ga0.2Sb has a larger conduction band offset (similar-to\sim 1.9 eV) relative to InAs than In0.8Al0.2As does (similar-to\sim 0.3 eV), thus providing a higher barrier next to the quantum well and allowing higher carrier densities to be achieved within a single 2D subband. Higher electron densities in turn can enable higher mobilities and stronger SOI than at lower electron densities.
  • Note [3] A 2DEG hosted in a binary alloy quantum well instead of a ternary alloy quantum well would not suffer from alloy scattering, which only occurs in ternary alloys and is a significant mechanism limiting mobilities.
  • Chang et al. [2015] W. Chang, S. M. Albrecht, T. S. Jespersen, F. Kuemmeth, P. Krogstrup, J. Nygård, and C. M. Marcus, “Hard gap in epitaxial semiconductor–superconductor nanowires,” Nat. Nanotechnol. 10, 232 (2015).
  • Birner et al. [2007] S. Birner, T. Zibold, T. Kubis, M. Sabathil, A. Trellakis, and P. Vogl, “nextnano: General Purpose 3-D Simulations,” IEEE Trans. Electron Dev. 54, 2137 (2007).
  • Trellakis et al. [2006] A. Trellakis, T. Zibold, T. Andlauer, S. Birner, R. K. Smith, R. Morschl, and P. Vogl, “The 3D nanometer device project nextnano: Concepts, methods, results,” J. Comput. Electron. 5, 285 (2006).
  • Uddin et al. [2013] M. M. Uddin, H. W. Liu, K. F. Yang, K. Nagase, K. Sekine, C. K. Gaspe, T. D. Mishima, M. B. Santos, and Y. Hirayama, “Gate depletion of an InSb two-dimensional electron gas,” Appl. Phys. Lett. 103, 123502 (2013).
  • Yi et al. [2015] W. Yi, A. A. Kiselev, J. Thorp, R. Noah, B.-M. Nguyen, S. Bui, R. D. Rajavel, T. Hussain, M. F. Gyure, P. Kratz, Q. Qian, M. J. Manfra, V. S. Pribiag, L. P. Kouwenhoven, C. M. Marcus, and M. Sokolich, “Gate-tunable high mobility remote-doped InSb/In1-xAlxSb quantum well heterostructures,” Appl. Phys. Lett. 106, 142103 (2015).
  • Kulesh et al. [2020] I. Kulesh, C. K. Ke, C. Thomas, S. Karwal, M. C. Moehle, S. Metti, R. Kallaher, C. G. Gardner, M. J. Manfra, and S. Goswami, “Quantum dots in an InSb two-dimensional electron gas,” Phys. Rev. Appl. 13, 041003 (2020).
  • Tuttle, Kroemer, and English [1990] G. Tuttle, H. Kroemer, and J. H. English, “Effects of interface layer sequencing on the transport properties of InAs/AlSb quantum wells: Evidence for antisite donors at the InAs/AlSb interface,” J. Appl. Phys. 67, 3032 (1990).
  • Ando, Fowler, and Stern [1982] T. Ando, A. B. Fowler, and F. Stern, “Electronic properties of two-dimensional systems,” Rev. Mod. Phys. 54, 437 (1982).
  • Shetty et al. [2022] A. Shetty, F. Sfigakis, W. Y. Mak, K. D. Gupta, B. Buonacorsi, M. C. Tam, H. S. Kim, I. Farrer, A. F. Croxall, H. E. Beere, A. R. Hamilton, M. Pepper, D. G. Austing, S. A. Studenikin, A. Sachrajda, M. E. Reimer, Z. R. Wasilewski, D. A. Ritchie, and J. Baugh, “Effects of biased and unbiased illuminations on two-dimensional electron gases in dopant-free GaAs/AlGaAs,” Phys. Rev. B 105, 075302 (2022).
  • Schubnikow and de Haas [1930] L. Schubnikow and W. J. de Haas, “A new phenomenon in the change of resistance in a magnetic field of single crystals of bismuth,” Nature 126, 500 (1930).
  • Muraki, Saku, and Hirayama [2001] K. Muraki, T. Saku, and Y. Hirayama, “Charge excitations in easy-axis and easy-plane quantum hall ferromagnets,” Phys. Rev. Lett. 87, 196801 (2001).
  • Zhang, Faulhaber, and Jiang [2005] X. C. Zhang, D. R. Faulhaber, and H. W. Jiang, “Multiple phases with the same quantized Hall conductance in a two-subband system,” Phys. Rev. Lett. 95, 216801 (2005).
  • Ellenberger et al. [2006] C. Ellenberger, B. Simovic̆, R. Leturcq, T. Ihn, S. E. Ulloa, K. Ensslin, D. C. Driscoll, and A. C. Gossard, “Two-subband quantum hall effect in parabolic quantum wells,” Phys. Rev. B 74, 195313 (2006).
  • Hikami, Larkin, and Nagaoka [1980] S. Hikami, A. I. Larkin, and Y. Nagaoka, “Spin-orbit interaction and magnetoresistance in the two dimensional random system,” Prog. Theoret. Phys. 63, 707 (1980).
  • Iordanskii et al. [1994] S. Iordanskii, Y. B. Lyanda-Geller, , and G. Pikus, “Weak localization in quantum wells with spin-orbit interaction,” JETP Letters 60, 206 (1994).
  • Wickramasinghe et al. [2018] K. S. Wickramasinghe, W. Mayer, J. Yuan, T. Nguyen, L. Jiao, V. Manucharyan, and J. Shabani, “Transport properties of near surface InAs two-dimensional heterostructures,” Appl. Phys. Lett. 113, 262104 (2018).
  • Witt et al. [2023] J. D. S. Witt, S. J. Pauka, G. C. Gardner, S. Gronin, T. Wang, C. Thomas, M. J. Manfra, D. J. Reilly, and M. C. Cassidy, “Spin-relaxation mechanisms in InAs quantum well heterostructures,” Appl. Phys. Lett. 122, 083101 (2023).
  • Farzaneh et al. [2024] S. M. Farzaneh, M. Hatefipour, W. F. Schiela, N. Lotfizadeh, P. Yu, B. H. Elfeky, W. M. Strickland, A. Matos-Abiague, and J. Shabani, “Observing magnetoanisotropic weak antilocalization in near-surface quantum wells,” Phys. Rev. Res. 6, 013039 (2024).
  • Bergeron et al. [2023] E. A. Bergeron, F. Sfigakis, Y. Shi, G. Nichols, P. C. Klipstein, A. Elbaroudy, S. M. Walker, Z. R. Wasilewski, and J. Baugh, “Field effect two-dimensional electron gases in modulation-doped InSb surface quantum wells,” Appl. Phys. Lett. 122, 012103 (2023).
  • Tajik, Haapamaki, and LaPierre [2012] N. Tajik, C. M. Haapamaki, and R. R. LaPierre, “Photoluminescence model of sulfur passivated p-InP nanowires,” Nanotechnology 23, 315703 (2012).
  • Lebedev [2020] M. V. Lebedev, “Modification of the atomic and electronic structure of III-V semiconductor surfaces at interfaces with electrolyte solutions,” Semiconductors 54, 699 (2020).
  • Bessolov and Lebedev [1998] V. N. Bessolov and M. V. Lebedev, “Chalcogenide passivation of III-V semiconductor surfaces,” Semiconductors 32, 1141 (1998).
  • Hazra et al. [2010] D. Hazra, L. M. A. Pascal, H. Courtois, and A. K. Gupta, “Hysteresis in superconducting short weak links and μ𝜇\muitalic_μ-SQUIDs,” Phys. Rev. B 82, 184530 (2010).
  • Vodolazov and Peeters [2011] D. Y. Vodolazov and F. M. Peeters, “Origin of the hysteresis of the current voltage characteristics of superconducting microbridges near the critical temperature,” Phys. Rev. B 84, 094511 (2011).
  • Nitta et al. [1992] J. Nitta, T. Akazaki, H. Takayanagi, and K. Arai, “Transport properties in an InAs-inserted-channel In0.52Al0.48As/In0.53Ga0.47As heterostructure coupled superconducting junction,” Phys. Rev. B 46, 14286(R) (1992).
  • Takayanagi and Akazaki [1995] H. Takayanagi and T. Akazaki, “Temperature dependence of the critical current in a clean-limit superconductor-2DEG-superconductor junction,” Solid State Commun. 96, 815 (1995).
  • Heida et al. [1998] J. P. Heida, B. J. van Wees, T. M. Klapwijk, and G. Borghs, “Nonlocal supercurrent in mesoscopic Josephson junctions,” Phys. Rev. B 57, R5618(R) (1998).
  • Giazotto et al. [2004] F. Giazotto, K. Grove-Rasmussen, R. Fazio, F. Beltram, E. H. Linfield, and D. A. Ritchie, “Josephson current in Nb/InAs/Nb highly transmissive ballistic junctions,” J. Supercond. 17, 317 (2004).
  • Blonder, Tinkham, and Klapwijk [1982] G. E. Blonder, M. Tinkham, and T. M. Klapwijk, “Transition from metallic to tunneling regimes in superconducting microconstrictions: Excess current, charge imbalance, and supercurrent conversion,” Phys. Rev. B 25, 4515 (1982).
  • Octavio et al. [1983] M. Octavio, M. Tinkham, G. E. Blonder, and T. M. Klapwijk, “Subharmonic energy-gap structure in superconducting constrictions,” Phys. Rev. B 27, 6739 (1983).
  • Flensberg, Hansen, and Octavio [1988] K. Flensberg, J. H. B. Hansen, and M. Octavio, “Subharmonic energy-gap structure in superconducting weak links,” Phys. Rev. B 38, 8707 (1988).
  • Cuevas, Martín-Rodero, and Yeyati [1996] J. C. Cuevas, A. Martín-Rodero, and A. L. Yeyati, “Hamiltonian approach to the transport properties of superconducting quantum point contacts,” Phys. Rev. B 54, 7366 (1996).
  • Niebler, Cuniberti, and Novotnỳ [2009] G. Niebler, G. Cuniberti, and T. Novotnỳ, “Analytical calculation of the excess current in the Octavio-Tinkham-Blonder-Klapwijk theory,” Supercond. Sci. Technol. 22, 085016 (2009).
  • Kjaergaard et al. [2017] M. Kjaergaard, H. Suominen, M. Nowak, A. Akhmerov, J. Shabani, C. Palmstrom, F. Nichele, and C. Marcus, “Transparent semiconductor-superconductor interface and induced gap in an epitaxial heterostructure Josephson junction,” Phys. Rev. Appl. 7, 034029 (2017).
  • Mayer et al. [2020] W. Mayer, W. F. Schiela, J. Yuan, M. Hatefipour, W. L. Sarney, S. P. Svensson, A. C. Leff, T. Campos, K. S. Wickramasinghe, M. C. Dartiailh, I. Zutic, and J. Shabani, “Superconducting proximity effect in InAsSb surface quantum wells with in-situ Al contact,” ACS Appl. Electron. Mat. 2, 2351 (2020).
  • Hertel et al. [2021] A. Hertel, L. O. Andersen, D. M. T. van Zanten, M. Eichinger, P. Scarlino, S. Yadav, J. Karthik, S. Gronin, G. C. Gardner, M. J. Manfra, C. M. Marcus, and K. D. Petersson, “Electrical properties of selective-area-grown superconductor-semiconductor hybrid structures on Silicon,” Phys. Rev. Appl. 16, 044015 (2021).
  • Haberkorn, Knauer, and Richter [1978] W. Haberkorn, H. Knauer, and J. Richter, “A theoretical study of the current-phase relation in Josephson contacts,” Phys. Stat. Sol. (a) 47, K161 (1978).
  • Mayer et al. [2019] W. Mayer, J. Yuan, K. S. Wickramasinghe, T. Nguyen, M. C. Dartiailh, and J. Shabani, “Superconducting proximity effect in epitaxial Al-InAs heterostructures,” Appl. Phys. Lett. 114, 103104 (2019).
  • Li et al. [2018] T. Li, J. Gallop, L. Hao, and E. Romans, “Ballistic josephson junctions based on cvd graphene,” Supercond. Sci. Technol. 31, 045004 (2018).
  • Lee et al. [2015] G.-H. Lee, S. Kim, S.-H. Jhi, and H.-J. Lee, “Ultimately short ballistic vertical graphene Josephson junctions,” Nat. Commun. 6, 6181 (2015).
  • Borzenets et al. [2016] I. Borzenets, F. Amet, C. Ke, A. Draelos, M. Wei, A. Seredinski, K. Watanabe, T. Taniguchi, Y. Bomze, M. Yamamoto, S. Tarucha, and G. Finkelstein, “Ballistic graphene Josephson junctions from the short to the long junction regimes,” Phys. Rev. Lett. 117, 237002 (2016).