A proximitized quantum dot in germanium

Lazar Lakic Center for Quantum Devices, Niels Bohr Institute, University of Copenhagen, 2100 Copenhagen, Denmark    William Iain L. Lawrie Center for Quantum Devices, Niels Bohr Institute, University of Copenhagen, 2100 Copenhagen, Denmark    David van Driel QuTech and Kavli Institute of Nanoscience, Delft University of Technology, Delft, The Netherlands    Lucas E. A. Stehouwer QuTech and Kavli Institute of Nanoscience, Delft University of Technology, Delft, The Netherlands    Menno Veldhorst QuTech and Kavli Institute of Nanoscience, Delft University of Technology, Delft, The Netherlands    Giordano Scappucci QuTech and Kavli Institute of Nanoscience, Delft University of Technology, Delft, The Netherlands    Ferdinand Kuemmeth Center for Quantum Devices, Niels Bohr Institute, University of Copenhagen, 2100 Copenhagen, Denmark    Anasua Chatterjee [email protected] Center for Quantum Devices, Niels Bohr Institute, University of Copenhagen, 2100 Copenhagen, Denmark QuTech and Kavli Institute of Nanoscience, Delft University of Technology, Delft, The Netherlands
(May 9, 2024)
Abstract

Planar germanium quantum wells have recently been shown to host a hard-gapped superconductor-semiconductor interface. Additionally, quantum dot spin qubits in germanium are well-suited for quantum information processing, with isotopic purification to a nuclear spin-free material expected to yield long coherence times. Therefore, as one of the few group IV materials with the potential to host superconductor-semiconductor hybrid devices, proximitized quantum dots in germanium are a crucial ingredient towards topological superconductivity and novel qubit modalities. Here we demonstrate a quantum dot (QD) in a Ge/SiGe heterostructure proximitized by a platinum germanosilicide (PtGeSi) superconducting lead (SC), forming a SC-QD-SC junction. We show tunability of the QD-SC coupling strength, as well as gate control of the ratio of charging energy and the induced gap. We further exploit this tunability by exhibiting control of the ground state of the system between even and odd parity. Furthermore, we characterize the critical magnetic field strengths, finding a robust critical out-of-plane field of 0.91±0.05plus-or-minus0.910.050.91~{}\pm~{}0.050.91 ± 0.05 T. Finally we explore sub-gap spin splitting in the device, observing rich physics in the resulting spectra, that we model using a zero-bandwidth model in the Yu-Shiba-Rusinov limit. The demonstration of controllable proximitization at the nanoscale of a germanium quantum dot opens up the physics of novel spin and superconducting qubits, and Josephson junction arrays in a group IV material.

I Introduction

Refer to caption
Figure 1: (a) False coloured scanning electron micrograph (SEM) of a nominally identical device. The device comprises three lithographically defined metallic layers, separated by a dielectric of Al2O3 grown using atomic layer deposition (ALD). A plunger gate PG (orange) controls the electrochemical potential of the quantum dot. The coupling of the quantum dot to the superconducting PtSiGe leads S and D is controlled by two barrier gates LB and RB (yellow). A cut-off gate CO prevents accumulation beneath the gate fan-out of PG, and a helper gate HG provides further control of the quantum dot confinement. (b) Heterostructure and gate stack schematic corresponding to the cross section indicated by the black dashed line in (a). (c) Energy schematic depicting the physical system in (a). Here, ΔΔ\Deltaroman_Δ is the SC gap energy, U𝑈Uitalic_U the charging energy of the QD, ΓSsubscriptΓS\Gamma_{\text{S}}roman_Γ start_POSTSUBSCRIPT S end_POSTSUBSCRIPT the hybridization energy of the SC and QD and ϵ0subscriptitalic-ϵ0\epsilon_{0}italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT the electrochemical potential of the QD with respect to the SC Fermi energy. (d) Source-drain current ISDSD{}_{\text{SD}}start_FLOATSUBSCRIPT SD end_FLOATSUBSCRIPT as a function of barrier gates VLBsubscript𝑉LBV_{\text{LB}}italic_V start_POSTSUBSCRIPT LB end_POSTSUBSCRIPT and VRBsubscript𝑉RBV_{\text{RB}}italic_V start_POSTSUBSCRIPT RB end_POSTSUBSCRIPT at bias voltage VSDsubscript𝑉SDV_{\text{SD}}italic_V start_POSTSUBSCRIPT SD end_POSTSUBSCRIPT = 500 μμ\upmuroman_μV. The square, circle and triangle correspond to the indicated gate voltage setting in (e), (f) and (g). (e-g) Bias spectroscopy for the three gate voltages indicated in (d), with high, moderate and low coupling of the quantum dot to the superconducting leads respectively, showing a transition between strongly coupled lead (e) and weakly coupled lead (g). Negative differential conductance observed may indicate Couloumb diamonds of odd occupancy. [1].

New and exotic physical phenomena can emerge in superconducting-semiconducting hybrids, enabling engineered quantum materials [2], circuit quantum electrodynamics (cQED) with novel superconducting qubits [3, 4], and topologically protected phases [5, 6, 7, 8]. In particular, proximitized quantum dots constitute key building blocks for devices such as Cooper pair splitters, Kitaev chains [7, 8], and protected qubits [9, 10]. However, to date the majority of these experiments have been performed in group III-V materials where nuclear spins are unavoidable, critically hampering spin coherence, and where 2D heterostructures exhibit piezoelectricity, deleterious for cQED circuits. Conversely, silicon and germanium are established group IV material platforms to integrate spin qubits hosted in gate defined quantum dots [11, 12], with isotopic purification having proved a gamechanger for ultra-long spin qubit coherence [13]. Contrary to silicon [14], germanium forms low resistance Ohmic contacts due to instrinsic Fermi level pinning close to the valence band [15, 16]. This has motivated a strong effort to induce superconductivity [17, 18, 19, 20, 21, 22] and very recently, hard-gap superconductivity has been demonstrated in mesoscopic devices implemented in a Ge/SiGe heterostructure [21, 22], in Ge/Si core shell nanowires [23] and in a cQED circuit [24].

Here, we present a superconducting-semiconducting hybrid quantum dot, which is hosted in Ge, a group IV material uniquely allowing for both isotopic purification  [25] and a superconducting hard gap [21, 22]. Our demonstration in a two-dimensional heterostructure establishes a novel platform that exhibits enhanced scalability compared to nanowires, is compatible with radiofrequency-reflectometry readout [26] and a highly successful spin qubit platform [27, 28, 29]. It may therefore be useful for long-range qubit interactions mediated via crossed Andreev reflection [30, 31] as well as heterogeneous quantum processors [9, 32] incorporating spin  [11, 12] and superconducting circuits [33]. Isotopically purified, proximitized germanium may be a crucial enabler for coherent Andreev spin qubits, protected superconducting qubits, and quantum dot-based Kitaev chains, and our first demonstration of a quantum dot with gate-tunable proximitization in a group IV heterostructure is a key ingredient.

In this work, superconducting polycrystalline Platinum-Germanium-silicide (PtSiGe) leads are formed by a controlled thermally-activated solid phase reaction between deposited platinum (Pt) and the heterostructure (Ge/SiGe) [21]. Importantly, the PtGeSi leads alleviate the need to etch into the heterostructure to deposit or pattern the superconductor, a potential source of damage exposing the quantum well and interface to air and processing. The leads act as charge reservoirs and as proximitization for the quantum dot (QD). We first demonstrate Coulomb blockade physics of a QD coupled to two superconducting leads (SC) forming a SC-QD-SC junction. We identify a superconducting gap energy window of 4Δ4Δ4\Delta4 roman_Δ inside which transport is suppressed, and outside which standard Coulomb diamonds are recovered. We observe sub-gap states in transport upon increasing the QD-SC coupling ΓSsubscriptΓS\Gamma_{\text{S}}roman_Γ start_POSTSUBSCRIPT S end_POSTSUBSCRIPT, which is consistent with the formation of Yu-Shiba-Rusinov (YSR) states in the system [34, 35]. We demonstrate gate control of ΓSsubscriptΓS\Gamma_{\text{S}}roman_Γ start_POSTSUBSCRIPT S end_POSTSUBSCRIPT by tuning the ground state at half-filling from a singlet state to a doublet state [36]. We then study the critical magnetic field of the hybrid device, finding a robust out-of-plane critical magnetic field Bc=0.91±0.05superscriptsubscript𝐵perpendicular-to𝑐plus-or-minus0.910.05B_{\perp}^{c}=0.91~{}\pm~{}0.05italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT = 0.91 ± 0.05 T. Finally, we investigate spin-splitting in the SC-QD-SC system, finding a g𝑔gitalic_g-factor of 1.5±0.2plus-or-minus0.2~{}\pm~{}0.2± 0.2 for an out-of-plane magnetic field, and also characterize the g𝑔gitalic_g-tensor anisotropy. To explain the energy splitting observed in the SC-QD-SC system we use a zero bandwidth (ZBW) Anderson Impurity model [37, 38] with the possibility of Zeeman splitting on the SC. Our observation of controllable subgap states and subgap spin splitting, magnetic field resilience, and the high tunability of the quantum dot-superconducting coupling establishes Ge/SiGe and PtSiGe as an attractive platform for hybrid quantum information processing.

Refer to caption
Figure 2: (a) Charge stability diagram of VRBsubscript𝑉RBV_{\mathrm{RB}}italic_V start_POSTSUBSCRIPT roman_RB end_POSTSUBSCRIPT vs VPGsubscript𝑉PGV_{\mathrm{PG}}italic_V start_POSTSUBSCRIPT roman_PG end_POSTSUBSCRIPT at VSD=80subscript𝑉SD80V_{\mathrm{SD}}=80italic_V start_POSTSUBSCRIPT roman_SD end_POSTSUBSCRIPT = 80 μμ\upmuroman_μV. (b-d) Bottom panels show bias spectroscopy at decreasing values of VRBsubscript𝑉RBV_{\mathrm{RB}}italic_V start_POSTSUBSCRIPT roman_RB end_POSTSUBSCRIPT corresponding to the square, circle, and triangle icons in panel a. Upper panels portray qualitative phase diagrams following ref. [39] of the expected ground state character of the hybrid system. Here, ϵ0subscriptitalic-ϵ0\epsilon_{0}italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the electrochemical potential of the quantum dot with respect to the grounded superconducting leads, and U𝑈Uitalic_U is the charging energy of the quantum dot. As the SC-QD coupling ΓSsubscriptΓS\Gamma_{\mathrm{S}}roman_Γ start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT is increased, the doublet state becomes energetically unfavorable, as seen by the merging of charge transitions (purple dashed lines). ΓSsubscriptΓS\Gamma_{\mathrm{S}}roman_Γ start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT is roughly estimated by modelling the sub-gap spectrum in the bottom panels (see Supplementary Information), to be 70 μμ\upmuroman_μeV, 110μμ~{}\upmuroman_μeV, and 150 μμ\upmuroman_μeV from b-d respectively, as indicated by the black dashed lines.

II Results

We utilize established fabrication protocols for quantum dot fabrication [40] and superconducting contacts [21] in Ge/SiGe quantum wells [41], to create a quantum dot coupled to two superconducting leads formed by rapid thermal annealing of Pt at 400superscript400400^{\circ}400 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPTC in Ar atmosphere for 15 minutes. Figure 1a shows a false-colored scanning electron micrograph of the device, consisting of one lithographically-defined layer for the superconducting leads (cyan) and two layers of electrostatic gates (yellow and orange, see Methods section for further details). Figure 1b shows a schematic of the cross-section of the device heterostructure and gate stack. Layers are electrically isolated from one another by 7 nm of Al2O3 deposited by atomic layer deposition at 150superscript150150^{\circ}150 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPTC. Barrier gates (LB, RB) control ΓSsubscriptΓS\Gamma_{\text{S}}roman_Γ start_POSTSUBSCRIPT S end_POSTSUBSCRIPT, while the plunger gate (PG) controls the relative electrochemical potential of the quantum dot levels with respect to the superconducting leads (ϵ0subscriptitalic-ϵ0\epsilon_{0}italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) as seen in Figure 1c. Two gates (HG and CO) are also used to confine the quantum dot and prevent unwanted accumulation. We utilize standard DC transport and low frequency lock-in techniques to measure source-drain current ISDsubscript𝐼SDI_{\mathrm{SD}}italic_I start_POSTSUBSCRIPT roman_SD end_POSTSUBSCRIPT and differential conductance G𝐺Gitalic_G across the quantum dot. Notably, our device is also connected to a radiofrequency (RF) reflectometry circuit via the source superconducting lead, and additional datasets in the low-coupling regime measured using this RF probe are presented in the Supplementary Information.

All data are taken at a lock-in frequency of 119 Hz, and amplitude of 2.5 μμ\upmuroman_μV. Figure 1d shows ISDsubscript𝐼SDI_{\mathrm{SD}}italic_I start_POSTSUBSCRIPT roman_SD end_POSTSUBSCRIPT as a function of the tunnel barrier VLBsubscript𝑉LBV_{\mathrm{LB}}italic_V start_POSTSUBSCRIPT roman_LB end_POSTSUBSCRIPT and VRBsubscript𝑉RBV_{\mathrm{RB}}italic_V start_POSTSUBSCRIPT roman_RB end_POSTSUBSCRIPT. Here, the source drain bias energy is set to eVSD𝑒subscript𝑉SDeV_{\mathrm{SD}}italic_e italic_V start_POSTSUBSCRIPT roman_SD end_POSTSUBSCRIPT = 300 μμ\upmuroman_μeV such that it exceeds the expected zero-field superconducting gap energy of 70μeVsimilar-toabsent70μeV\sim 70\,\,\upmu\mathrm{eV}∼ 70 roman_μ roman_eV [21]. We set VLBsubscript𝑉LBV_{\mathrm{LB}}italic_V start_POSTSUBSCRIPT roman_LB end_POSTSUBSCRIPT close to its pinch-off value such that it acts as a tunnel probe, and vary VRBsubscript𝑉RBV_{\mathrm{RB}}italic_V start_POSTSUBSCRIPT roman_RB end_POSTSUBSCRIPT to tune the coupling between superconductor and quantum dot ΓSsubscriptΓS\Gamma_{\mathrm{S}}roman_Γ start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT (Figure 1c). Figures 1e-g show bias spectroscopy at different values of VRBsubscript𝑉RBV_{\mathrm{RB}}italic_V start_POSTSUBSCRIPT roman_RB end_POSTSUBSCRIPT. In the strong coupling regime (Figure 1e), we observe a range in bias energy of 4Δ04subscriptΔ04\Delta_{0}4 roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT where transport is suppressed, from which we extract a superconducting pairing amplitude of Δ0subscriptΔ0\Delta_{0}roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 72 ±plus-or-minus\pm± 6  μμ\upmuroman_μeV, in its fully open state. Figure 1f shows that as ΓSsubscriptΓS\Gamma_{\mathrm{S}}roman_Γ start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT is decreased (positive change on VRBsubscript𝑉RBV_{\mathrm{RB}}italic_V start_POSTSUBSCRIPT roman_RB end_POSTSUBSCRIPT), we observe tunnel-broadened Coloumb oscillations and sub-gap transport features, indicating a hybridized QD. At low coupling between the QD and SC (Fig. 1g) we observe sharp Coulomb diamonds outside a bias window of ±plus-or-minus\pm±2Δ0subscriptΔ0\Delta_{0}roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. We conclude that we have versatile electrostatic control of the degree of hybridization of a quantum dot with a superconductor, consistent with experiments performed in InAs-Al nanowires [42] and InSbAs-Al 2DEGs [7].

II.1 Singlet-doublet quantum phase transition

A quantum dot coupled to a superconducting lead at half-occupancy of charges can have two different ground states, depending on the degree of superconductor-quantum dot coupling. At low coupling strengths and zero magnetic field strength, the ground state at half-filling, ie. ϵ0/U=0.5subscriptitalic-ϵ0𝑈0.5\epsilon_{0}/U=0.5italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_U = 0.5, will be a spin-degenerate doublet state |D={|,|}ket𝐷ketket\ket{D}=\{\ket{\downarrow},\ket{\uparrow}\}| start_ARG italic_D end_ARG ⟩ = { | start_ARG ↓ end_ARG ⟩ , | start_ARG ↑ end_ARG ⟩ }. Here, ϵ0subscriptitalic-ϵ0\epsilon_{0}italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the electrochemical potential of the QD with respect to the SC lead and U𝑈Uitalic_U is the charging energy of the QD. At high coupling, a preference for superconducting pairing will dominate, leading to a singlet ground state |S=u|0v|2ket𝑆𝑢ket0𝑣ket2\ket{S}=u\ket{0}-v\ket{2}| start_ARG italic_S end_ARG ⟩ = italic_u | start_ARG 0 end_ARG ⟩ - italic_v | start_ARG 2 end_ARG ⟩. By utilizing the control of ΓSsubscriptΓS\Gamma_{\text{S}}roman_Γ start_POSTSUBSCRIPT S end_POSTSUBSCRIPT demonstrated above, we show that we can tune between these ground states. We operate in a regime whereby VLBsubscript𝑉LBV_{\mathrm{LB}}italic_V start_POSTSUBSCRIPT roman_LB end_POSTSUBSCRIPT is very close to its pinch-off value, such that it acts as a tunneling probe. We then vary VRBsubscript𝑉RBV_{\mathrm{RB}}italic_V start_POSTSUBSCRIPT roman_RB end_POSTSUBSCRIPT to tune the coupling ΓSsubscriptΓS\Gamma_{\mathrm{S}}roman_Γ start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT. Figure 2a shows a charge stability diagram of the system with the QD plunger gate VPGsubscript𝑉PGV_{\mathrm{PG}}italic_V start_POSTSUBSCRIPT roman_PG end_POSTSUBSCRIPT on the horizontal axis, and the QD-SC barrier gate VRBsubscript𝑉RBV_{\mathrm{RB}}italic_V start_POSTSUBSCRIPT roman_RB end_POSTSUBSCRIPT on the vertical axis, at a bias energy of eVSD=80𝑒subscript𝑉SD80eV_{\mathrm{SD}}=80italic_e italic_V start_POSTSUBSCRIPT roman_SD end_POSTSUBSCRIPT = 80 μμ\upmuroman_μeV, slightly above ΔΔ\Deltaroman_Δ. The vertical lines measured are Coulomb resonances indicating transitions between the N𝑁Nitalic_N, N+1𝑁1N+1italic_N + 1 and N+2𝑁2N+2italic_N + 2 occupations of the quantum dot (from right to left). As we increase ΓSsubscriptΓS\Gamma_{\mathrm{S}}roman_Γ start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT by making RB more negative we observe the merging of two levels at VRBsubscript𝑉RBV_{\mathrm{RB}}italic_V start_POSTSUBSCRIPT roman_RB end_POSTSUBSCRIPT = -1.395 V.

We further investigate these transitions with bias spectroscopy as a function of plunger gate voltage at different values of VRBsubscript𝑉RBV_{\mathrm{RB}}italic_V start_POSTSUBSCRIPT roman_RB end_POSTSUBSCRIPT. In the bottom panel of Figure 2b, we show bias spectroscopy at VRB=1.37 Vsubscript𝑉RBtimes-1.37voltV_{\mathrm{RB}}=$-1.37\text{\,}\mathrm{V}$italic_V start_POSTSUBSCRIPT roman_RB end_POSTSUBSCRIPT = start_ARG - 1.37 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG (square in Figure 2a) for varying VSDsubscript𝑉SDV_{\mathrm{SD}}italic_V start_POSTSUBSCRIPT roman_SD end_POSTSUBSCRIPT and VPGsubscript𝑉PGV_{\mathrm{PG}}italic_V start_POSTSUBSCRIPT roman_PG end_POSTSUBSCRIPT. At VPGsubscript𝑉PGV_{\mathrm{PG}}italic_V start_POSTSUBSCRIPT roman_PG end_POSTSUBSCRIPT values of -1.81 V and -1.808 V, the state crosses Δ0subscriptΔ0\Delta_{0}roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, signalling the changes in ground state parity as seen from 2a.

In the bottom panel of Figure 2c, we show spectroscopy at VRB=1.385 Vsubscript𝑉RBtimes-1.385voltV_{\mathrm{RB}}=$-1.385\text{\,}\mathrm{V}$italic_V start_POSTSUBSCRIPT roman_RB end_POSTSUBSCRIPT = start_ARG - 1.385 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG (circle in Figure 2a). We see an evolution of the state features into a characteristic eye-shape indicating the formation of YSR states [34, 43, 44, 45, 46] on the hybridized QD. The negative differential conductance is attributed to probing sub-gap features with a coherence peak [47]. In the bottom panel of Figure 2d, we perform bias spectroscopy at VRB=1.4025 Vsubscript𝑉RBtimes-1.4025voltV_{\mathrm{RB}}=$-1.4025\text{\,}\mathrm{V}$italic_V start_POSTSUBSCRIPT roman_RB end_POSTSUBSCRIPT = start_ARG - 1.4025 end_ARG start_ARG times end_ARG start_ARG roman_V end_ARG (triangle in Figure 2a) showing no parity change, which we interpret as the QD filling remaining in a singlet ground state upon loading an additional hole.

In QD-SC systems, where the charging energy of the QD U𝑈Uitalic_U is larger than the SC order parameter ΔΔ\Deltaroman_Δ, quasiparticles in the SC can bind to the dot by the exchange interaction and give rise to sub-gap excitations in the form of YSR states. Such systems can be modelled using a zero-bandwidth (ZBW) model that describes a quantum dot coupled to a single superconducting orbital [37, 48], which predicts which ground state the system prefers depending on the degree of hybridization between the SC and the QD. By solving the ZBW for where the energy of the singlet equals the one of the doublet, a singlet doublet phase transition diagram can be realized. In the top panels of Figure 2b-d such a phase transition has been illustrated and dashed lines have been inserted at ΓSsubscriptΓS\Gamma_{\mathrm{S}}roman_Γ start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT values based on extracted coupling values (70μeV70μeV70~{}\upmu\mathrm{eV}70 roman_μ roman_eV, 110μeV110μeV110~{}\upmu\mathrm{eV}110 roman_μ roman_eV, and 150μeV150μeV150~{}\upmu\mathrm{eV}150 roman_μ roman_eV from b-d respectively), using the aforementioned minimal ZBW model (see Supplementary Information). As the hybridization energy ΓSsubscriptΓS\Gamma_{\mathrm{S}}roman_Γ start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT increases, it becomes less favorable to maintain the |Dket𝐷\ket{D}| start_ARG italic_D end_ARG ⟩ ground state. The vertical dashed lines indicate a qualitative correspondence between the experimental barrier gate RB controlling ΓSsubscriptΓS\Gamma_{\mathrm{S}}roman_Γ start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT, and the calculated ΓSsubscriptΓS\Gamma_{\mathrm{S}}roman_Γ start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT in the phase diagram. The magenta stippled lines in the bottom panels serve as guides to the eye to indicate where the phase transition occurs.

II.2 Magnetic field characterization

Refer to caption
Figure 3: (a)] Bias spectroscopy as a function of out-of-plane magnetic field Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT. We extract a critical field Bc,subscript𝐵𝑐perpendicular-toB_{c,{\perp}}italic_B start_POSTSUBSCRIPT italic_c , ⟂ end_POSTSUBSCRIPT = 0.91±0.05plus-or-minus0.05\pm 0.05± 0.05 T. (b-c) Bias spectroscopy at low ΓSsubscriptΓS\Gamma_{\mathrm{S}}roman_Γ start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT, for out-of-plane magnetic field B=1subscript𝐵perpendicular-to1B_{\perp}=1italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 1 T, and B=0subscript𝐵perpendicular-to0B_{\perp}=0italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 0 T respectively. Top panels show extracted charging energies from Coloumb diamonds below.
Refer to caption
Figure 4: (a-c) Bias spectroscopy displayed in logarithmic scale at magnetic field strengths of 0 mT, 250 mT and 350 mT respectively. Sub-gap energy splitting is visible in b) and c). The superimposed orange markers in Figures a-c serve as a qualitative comparison between the measured transport data and calculations using a zero-bandwidth model, inputting a Δ0=72μsubscriptΔ072μ\Delta_{0}=72\,\,\upmuroman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 72 roman_μeV, hybridization energy ΓSsubscriptΓS\Gamma_{\text{S}}roman_Γ start_POSTSUBSCRIPT S end_POSTSUBSCRIPT=110 μμ\upmuroman_μeV, charging energy U𝑈Uitalic_U = 1.61.61.61.6 meV, as well as a Zeeman energy of EzSC=0μsuperscriptsubscript𝐸𝑧𝑆𝐶0μE_{z}^{SC}=0\upmuitalic_E start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_S italic_C end_POSTSUPERSCRIPT = 0 roman_μeV on the SC and EzQD=0μsuperscriptsubscript𝐸𝑧𝑄𝐷0μE_{z}^{QD}=0\,\,\upmuitalic_E start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_Q italic_D end_POSTSUPERSCRIPT = 0 roman_μeV on the QD, EzSC=44μsuperscriptsubscript𝐸𝑧𝑆𝐶44μE_{z}^{SC}=44\,\,\upmuitalic_E start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_S italic_C end_POSTSUPERSCRIPT = 44 roman_μeV on the SC and EzQD=44μsuperscriptsubscript𝐸𝑧𝑄𝐷44μE_{z}^{QD}=44\,\,\upmuitalic_E start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_Q italic_D end_POSTSUPERSCRIPT = 44 roman_μeV on the QD, and EzSC=56μsuperscriptsubscript𝐸𝑧𝑆𝐶56μE_{z}^{SC}=56\,\,\upmuitalic_E start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_S italic_C end_POSTSUPERSCRIPT = 56 roman_μeV on the SC and EzQD=56μsuperscriptsubscript𝐸𝑧𝑄𝐷56μE_{z}^{QD}=56\,\,\upmuitalic_E start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_Q italic_D end_POSTSUPERSCRIPT = 56 roman_μeV on the QD, respectively. (d) Magnetic field sweep at VPGsubscript𝑉PGV_{\mathrm{PG}}italic_V start_POSTSUBSCRIPT roman_PG end_POSTSUBSCRIPT = -1.8015 V . We extract a g-factor for the out-of-plane magnetic field splitting gsubscript𝑔perpendicular-tog_{\perp}italic_g start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 1.5±0.2plus-or-minus1.50.21.5\pm 0.21.5 ± 0.2. (e-g) Bias spectroscopy of rotating magnetic field at total magnetic field strength |B|𝐵|B|| italic_B | = 400 mT and VPGsubscript𝑉PGV_{\mathrm{PG}}italic_V start_POSTSUBSCRIPT roman_PG end_POSTSUBSCRIPT = -1.8017 V. Strong g𝑔gitalic_g-tensor anisotropy is observed between out-of-plane and in-plane magnetic field orientations.

We now turn to the magnetic field dependence of the superconducting parent gap. Figure 3a shows bias spectroscopy of the quantum dot in the few hole and low ΓSsubscriptΓS\Gamma_{\mathrm{S}}roman_Γ start_POSTSUBSCRIPT roman_S end_POSTSUBSCRIPT regime, as a function of out-of-plane magnetic field strength Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT. We fit the closing of the superconducting gap according to Δ(B)Δ𝐵\Delta(B)roman_Δ ( italic_B ) = 2Δ01(B/Bc)2subscriptΔ01superscript𝐵subscript𝐵𝑐2\Delta_{0}\sqrt{1-(B/B_{c})^{2}}roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT square-root start_ARG 1 - ( italic_B / italic_B start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG [49], where Δ0subscriptΔ0\Delta_{0}roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the superconducting gap at zero magnetic field, and Bcsubscript𝐵𝑐B_{c}italic_B start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is the critical magnetic field. We extract a critical field of Bc,subscript𝐵𝑐perpendicular-toB_{c,\perp}italic_B start_POSTSUBSCRIPT italic_c , ⟂ end_POSTSUBSCRIPT = 0.91±0.05plus-or-minus0.910.050.91\pm 0.050.91 ± 0.05 T. This greatly exceeds the critical field measured in prior studies, where the critical out-of-plane magnetic field was measured to be approximately 50 mT [21]. This disparity could be due to the smaller size of the superconducting junctions and leads [50] measured in the present experiment compared to those in ref. [21], which may alleviate vortex formation. We also characterize the critical magnetic field strengths for the two in-plane axes in Supplementary Section II. Figures 3b and 3c show bias spectroscopy of the quantum dot in the low coupling limit, taken at 1 T and 0 T respectively. At 0 T the superconducting gap is present within the Coulomb diamonds (Figure 3b), while normal Coulomb diamonds are recovered at 1 T due to the breaking down of superconductivity (Figure 3c). In both cases we find an even-odd oscillation in the filling structure at both low and high field strengths as seen in the top panels of Figures 3b and 3c, depicting the addition energy of each Coulomb diamond below, consistent with that observed previously in germanium [51] and InAs [43] QDs. The charging energy of the quantum dot varies between even and odd periodicity, indicating that the quantum dot is in the low hole occupancy. The charging energy is typically between 1 and 1.8 meV, more than ten times the superconducting gap, supporting our interpretation that we are in the YSR regime.

Finally we study the transport spectrum under the influence of a magnetic field in the same electrostatic regime as Figure 2b-c. Figures 4a-c show bias spectroscopy measurements as a function of VPGsubscript𝑉PGV_{\mathrm{PG}}italic_V start_POSTSUBSCRIPT roman_PG end_POSTSUBSCRIPT portrayed in logarithmic scale for enhanced visibility, taken at a perpendicular magnetic field, at strengths of (a) 0 mT, (b) 250 mT, and (c) 350 mT (see Supplementary Information for data in non-logarithmic scale). At 0 mT a |Sket𝑆\ket{S}| start_ARG italic_S end_ARG ⟩-|Dket𝐷\ket{D}| start_ARG italic_D end_ARG ⟩-|Sket𝑆\ket{S}| start_ARG italic_S end_ARG ⟩ transition spectrum is observed, as seen in Figure 2b-c. A qualititive ZBW model of the system is plotted on top of the data as described previously (see Supplementary Information for details of the model), using a SC pairing energy of Δ0=72μeVsubscriptΔ072μeV\Delta_{0}=72~{}\upmu\mathrm{eV}roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 72 roman_μ roman_eV, a hybridization energy of ΓS=110μeVsubscriptΓS110μeV\Gamma_{\text{S}}=110~{}\upmu\mathrm{eV}roman_Γ start_POSTSUBSCRIPT S end_POSTSUBSCRIPT = 110 roman_μ roman_eV, and a QD charging energy of U𝑈Uitalic_U = 1.6 meV. As we increase the perpendicular field an energy splitting of the subgap states is observed in the singlet ground-state sectors 4b-c. Interestingly, the energy splitting seen in the even parity groundstates have a flat energy dispersion indicating either a spin splitting of the parent gap itself, or of a strongly coupled sub-gap state, which we attribute to spinful excited quasiparticles. To qualitatively model the data, we introduce a Zeeman splitting term into the ZBW model for both the SC and QD. We find that setting the g-factor of the QD and superconducting orbitals to be equal in magnitude is sufficient to phenomonenologically model the data potentially due to a g-factor renormalization as a result of hybridization [52, 53]. Additional data at lower magnetic field strength and lower coupling, supporting our hypothesis of g-factor renormalization, is reported in the Supplementary Information, as well as results from our ZBW model with varying system parameters.

In Figure 4d, we set the plunger gate voltage VPGsubscript𝑉PGV_{\mathrm{PG}}italic_V start_POSTSUBSCRIPT roman_PG end_POSTSUBSCRIPT to -1.8017 V (orange notch in Figure 4a) and perform bias spectroscopy as a function of Bsubscript𝐵perpendicular-toB_{\perp}italic_B start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT from 0 T to 0.7 T. We observe the magnetic field splitting of the superconducting coherence peaks, and extract an out-of-plane g-factor, gsubscript𝑔perpendicular-tog_{\perp}italic_g start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 1.5±0.2plus-or-minus1.50.21.5~{}\pm~{}0.21.5 ± 0.2. This value is several times larger than the g-factor measured for magnetic fields close to in-plane in planar germanium quantum wells [54, 55], and several times smaller than the value reported in ref. [56] for out of plane g-factors. Furthermore, in a regime of lower coupling for the YSR states, we have measured an out-of-plane g-factor of g=4.5±0.6subscript𝑔perpendicular-toplus-or-minus4.50.6g_{\perp}=4.5~{}\pm~{}0.6italic_g start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 4.5 ± 0.6 (see Supplementary Information, Figure S1) and a g=5.3±0.8subscript𝑔perpendicular-toplus-or-minus5.30.8g_{\perp}=5.3~{}\pm~{}0.8italic_g start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 5.3 ± 0.8 in out-of-gap Coulomb diamond spectroscopy (see Supplementary Information, Figure S2). These observations support the hypothesis that the g-factor gsubscript𝑔perpendicular-tog_{\perp}italic_g start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT either describes the quasiparticle coherence peaks, or is due to renormalization of the Ge/SiGe hole g𝑔gitalic_g-factor as a result of hybridization with the superconductor [52, 53].

Finally, we investigate anisotropy of the g-factor by performing bias spectroscopy as a function of magnetic field orientation. Figures 4e-g show bias spectroscopy at a magnetic field strength of |B|=400𝐵400|B|=400| italic_B | = 400 mT. The splitting is then investigated as a function of rotation angles. Here, the rotation angles θ𝜃\thetaitalic_θ, ϕitalic-ϕ\phiitalic_ϕ and α𝛼\alphaitalic_α are defined as shown in Figure 4e-g. A large anisotropy is measured; while g-tensor anisotropy is ubiquitous for heavy holes in strained planar Ge/SiGe wells, it is seldom observed for a superconducting gap edge. On the other hand, anisotropic g-tensors have been observed in heavy fermion bulk superconductors, which could additionally explain the non-dispersive splitting in Fig. 4b-c [57]. We stress that our use of the annealed poly-crystalline superconductor PtGeSi is a recent material development in itself, and the physics of superconductivity in these nanoscale thin films is not fully understood. However, this anisotropy could also be explained by a sub-gap state in the QD that is strongly coupled to the superconductor.

III Conclusion

We have demonstrated a quantum dot in Ge/SiGe proximitized by a superconducting lead and exhibiting robust YSR states. We find that the coupling between quantum dot and superconducting lead is highly tunable, as evidenced by the tunnel barrier controllability of the singlet or doublet nature of the YSR ground state at half-filling. Additionally, we have characterized the critical magnetic field strength, finding a robust out-of-plane critical field of 0.91similar-toabsent0.91\sim 0.91∼ 0.91 T. Finally, we observe Zeeman splitting of sub-gap states, which we explain using a modified zero-bandwidth Anderson impurity model. The ability to tunably and strongly couple a superconductor to a quantum dot, in combination with a robust critical magnetic field, demonstrates the feasability of our platform for hybrid germanium quantum information processing, including Andreev spin qubits and topological quantum computing, as well as the exploration of fundamental physics with superconductor-semiconductor devices. While the in-plane g-tensor component of holes in germanium planar wells is lower than in established group III-V platforms, it could be enhanced by confinement-induced g-factor engineering [58], and the out-of-plane g-factor could provide an alternative route to engineering topologically protected qubits using QD-SC chains [59]. Further work on RF-reflectometry-based measurements (as described in the Supplementary Information) could help achieve charge sensing and parity readout. Our demonstration in a group IV material, amenable to isotopic purification, constitutes a crucial building block for superconducting-semiconducting hybrid technologies and opens up previously inaccessible experimental directions.

These authors contributed equally.

IV Author Information

L.E.A.S and G.S. grew and supplied the Ge/SiGe heterostructures and developed processes for the PtGeSi contacts. L.L. and W.I.L.L. designed the devices. L.L. fabricated the devices. L.L. and W.I.L.L. performed the experiment in the dilution refrigerator along with D.v.D., and analysed the data. L.L. performed numerical simulations within the ZBW model with help from D.v.D and W.I.L.L. L.L and W.I.L.L. wrote the manuscript with input from D.v.D., M.V., G.S., F.K. and A.C. A.C. and F.K. supervised the project.

V Acknowledgments

We thank J. Paaske, V. Baran and G. Mazur for valuable discussions. This project has received funding from the European Research Council (ERC) as part of the project NONLOCAL under grant agreement No 856526, and through the IGNITE project under grant agreement No. 101069515 of the Horizon Europe Framework Programme. AC acknowledges support from the Inge Lehmann Programme of the Independent Research Fund Denmark. We acknowledge funding by the Casimir PhD Travel Grant.

VI Data Availability

Raw data and analysis scripts for all data included in this work are available at the Zenodo data repository at https://doi.org/10.5281/zenodo.11088753.

References

  • De Franceschi et al. [2010] S. De Franceschi, L. Kouwenhoven, C. Schoenenberger, and W. Wernsdorfer, Hybrid superconductor–quantum dot devices, Nature Nanotechnology 5, 703 (2010).
  • Bøttcher et al. [2018] C. Bøttcher, F. Nichele, and M. e. a. Kjaergaard, Superconducting, insulating and anomalous metallic regimes in a gated two-dimensional semiconductor–superconductor array, Nature Physics , 1138–1144 (2018).
  • Hays et al. [2021] M. Hays, V. Fatemi, D. Bouman, J. Cerrillo, S. Diamond, K. Serniak, T. Connolly, P. Krogstrup, J. Nygård, A. L. Yeyati, A. Geresdi, and M. H. Devoret, Coherent manipulation of an andreev spin qubit, Science 373, 430 (2021).
  • Bargerbos et al. [2023] A. Bargerbos, M. Pita-Vidal, R. Zitko, L. J. Splitthoff, L. Grunhaupt, J. J. Wesdorp, Y. Liu, L. P. Kouwenhoven, R. Aguado, C. K. Andersen, A. Kou, and B. van Heck, Spectroscopy of spin-split andreev levels in a quantum dot with superconducting leads, Phys. Rev. Lett. 131, 097001 (2023).
  • Leijnse and Flensberg [2012] M. Leijnse and K. Flensberg, Parity qubits and poor man’s majorana bound states in double quantum dots, Phys. Rev. B 86, 134528 (2012).
  • Kitaev [2001] A. Y. Kitaev, Unpaired majorana fermions in quantum wires, Physics-Uspekhi 4410.1070/1063-7869/44/10S/S29 (2001).
  • ten Haaf et al. [2023] S. L. ten Haaf, Q. Wang, A. M. Bozkurt, C.-X. Liu, I. Kulesh, P. Kim, D. Xiao, C. Thomas, M. J. Manfra, T. Dvir, et al., Engineering majorana bound states in coupled quantum dots in a two-dimensional electron gas, arXiv preprint arXiv:2311.03208  (2023).
  • Dvir et al. [2023] T. Dvir, G. Wang, N. van Loo, C.-X. Liu, G. P. Mazur, A. Bordin, S. L. Ten Haaf, J.-Y. Wang, D. van Driel, F. Zatelli, et al., Realization of a minimal kitaev chain in coupled quantum dots, Nature 614, 445 (2023).
  • Pino et al. [2024] D. M. Pino, R. S. Souto, and R. Aguado, Minimal kitaev-transmon qubit based on double quantum dots, Phys. Rev. B 109, 075101 (2024).
  • Wang et al. [2023] Q. Wang, S. L. D. ten Haaf, I. Kulesh, D. Xiao, C. Thomas, M. J. Manfra, and S. Goswami, Triplet correlations in Cooper pair splitters realized in a two-dimensional electron gas, Nature Communications 14, 4876 (2023).
  • Burkard et al. [2023] G. Burkard, T. D. Ladd, A. Pan, J. M. Nichol, and J. R. Petta, Semiconductor spin qubits, Rev. Mod. Phys. 95, 025003 (2023).
  • Scappucci et al. [2021] G. Scappucci, C. Kloeffel, F. A. Zwanenburg, D. Loss, M. Myronov, J.-J. Zhang, S. De Franceschi, G. Katsaros, and M. Veldhorst, The germanium quantum information route, Nature Reviews Materials 6, 926 (2021).
  • Muhonen et al. [2014] J. T. Muhonen, J. P. Dehollain, A. Laucht, F. E. Hudson, R. Kalra, T. Sekiguchi, K. M. Itoh, D. N. Jamieson, J. C. Mccallum, A. S. Dzurak, and A. Morello, Storing quantum information for 30 seconds in a nanoelectronic device, Nature Nanotechnology 9, 986 (2014).
  • Tao et al. [2003] M. Tao, S. Agarwal, D. Udeshi, N. Basit, E. Maldonado, and W. P. Kirk, Low Schottky barriers on n-type silicon (001), Applied Physics Letters 83, 2593 (2003).
  • Dimoulas et al. [2006] A. Dimoulas, P. Tsipas, A. Sotiropoulos, and E. K. Evangelou, Fermi-level pinning and charge neutrality level in germanium, Applied Physics Letters 89, 252110 (2006).
  • Nishimura et al. [2007] T. Nishimura, K. Kita, and A. Toriumi, Evidence for strong Fermi-level pinning due to metal-induced gap states at metal/germanium interface, Applied Physics Letters 91, 123123 (2007).
  • Aggarwal et al. [2021] K. Aggarwal, A. Hofmann, D. Jirovec, I. Prieto, A. Sammak, M. Botifoll, S. Marti-Sanchez, M. Veldhorst, J. Arbiol, G. Scappucci, J. Danon, and G. Katsaros, Enhancement of proximity-induced superconductivity in a planar ge hole gas, Phys. Rev. Res. 3, L022005 (2021).
  • Vigneau et al. [2019] F. Vigneau, R. Mizokuchi, D. C. Zanuz, X. Huang, S. Tan, R. Maurand, S. Frolov, A. Sammak, G. Scappucci, F. Lefloch, and S. De Franceschi, Germanium quantum-well josephson field-effect transistors and interferometers, Nano Letters 19, 1023 (2019).
  • Hendrickx et al. [2018] N. W. Hendrickx, D. P. Franke, A. Sammak, M. Kouwenhoven, D. Sabbagh, L. Yeoh, R. Li, M. L. V. Tagliaferri, M. Virgilio, G. Capellini, G. Scappucci, and M. Veldhorst, Gate-controlled quantum dots and superconductivity in planar germanium, Nature Communications 9, 2835 (2018).
  • Ridderbos et al. [2018] J. Ridderbos, M. Brauns, J. Shen, F. K. de Vries, A. Li, E. P. A. M. Bakkers, A. Brinkman, and F. A. Zwanenburg, Josephson effect in a few-hole quantum dot, Advanced Materials 30, 1802257 (2018).
  • Tosato et al. [2023] A. Tosato, V. Levajac, J.-Y. Wang, C. J. Boor, F. Borsoi, M. Botifoll, C. N. Borja, S. Marti-Sanchez, J. Arbiol, A. Sammak, M. Veldhorst, and G. Scappucci, Hard superconducting gap in germanium, Communications Materials 4, 23 (2023).
  • Valentini et al. [2024] M. Valentini, O. Sagi, and L. e. a. Baghumyan, Parity-conserving cooper-pair transport and ideal superconducting diode in planar germanium, Nat Commun 1510.1038/s41467-023-44114-0 (2024).
  • Zhuo et al. [2023] E. Zhuo, Z. Lyu, and X. S. et al., Hole-type superconducting gatemon qubit based on ge/si core/shell nanowires, npj quantum information 910.1038/s41534-023-00721-9 (2023).
  • Hinderling et al. [2024] M. Hinderling, S. ten Kate, M. Coraiola, D. Haxell, M. Stiefel, M. Mergenthaler, S. Paredes, S. Bedell, D. Sabonis, and F. Nichele, Direct microwave spectroscopy of andreev bound states in planar ge josephson junctions, arXiv preprint arXiv:2403.03800  (2024).
  • Itoh et al. [1993] K. Itoh, W. Hansen, E. Haller, J. Farmer, V. Ozhogin, A. Rudnev, and A. Tikhomirov, High purity isotopically enriched 70ge and 74ge single crystals: Isotope separation, growth, and properties, Journal of Materials Research 8, 1341–1347 (1993).
  • Vigneau et al. [2023] F. Vigneau, F. Fedele, A. Chatterjee, D. Reilly, F. Kuemmeth, M. F. Gonzalez-Zalba, E. Laird, and N. Ares, Probing quantum devices with radio-frequency reflectometry, Applied Physics Reviews 10, 021305 (2023).
  • Borsoi et al. [2024] F. Borsoi, N. W. Hendrickx, V. John, M. Meyer, S. Motz, F. van Riggelen, A. Sammak, S. L. de Snoo, G. Scappucci, and M. Veldhorst, Shared control of a 16 semiconductor quantum dot crossbar array, Nature Nanotechnology 19, 21 (2024).
  • Hendrickx et al. [2021] N. W. Hendrickx, W. I. L. Lawrie, M. Russ, F. van Riggelen, S. L. de Snoo, R. N. Schouten, A. Sammak, G. Scappucci, and M. Veldhorst, A four-qubit germanium quantum processor, Nature 591, 580 (2021).
  • Jirovec et al. [2021] D. Jirovec, A. Hofmann, A. Ballabio, P. M. Mutter, G. Tavani, M. Botifoll, A. Crippa, J. Kukucka, O. Sagi, F. Martins, et al., A singlet-triplet hole spin qubit in planar ge, Nature Materials 20, 1106 (2021).
  • Leijnse and Flensberg [2013] M. Leijnse and K. Flensberg, Coupling spin qubits via superconductors, Phys. Rev. Lett. 111, 060501 (2013).
  • Spethmann et al. [2024] M. Spethmann, S. Bosco, A. Hofmann, J. Klinovaja, and D. Loss, High-fidelity two-qubit gates of hybrid superconducting-semiconducting singlet-triplet qubits, Physical Review B 109, 085303 (2024).
  • Gyenis et al. [2021] A. Gyenis, A. Di Paolo, J. Koch, A. Blais, A. A. Houck, and D. I. Schuster, Moving beyond the transmon: Noise-protected superconducting quantum circuits, PRX Quantum 2, 030101 (2021).
  • Kjaergaard et al. [2020] M. Kjaergaard, M. E. Schwartz, J. Braumuller, P. Krantz, J. I.-J. Wang, S. Gustavsson, and W. D. Oliver, Superconducting qubits: Current state of play, Annual Review of Condensed Matter Physics 11, 369 (2020).
  • Kirsanskas et al. [2015] G. Kirsanskas, M. Goldstein, K. Flensberg, L. I. Glazman, and J. Paaske, Yu-shiba-rusinov states in phase-biased superconductor–quantum dot–superconductor junctions, Phys. Rev. B 92, 235422 (2015).
  • Jellinggaard et al. [2016] A. Jellinggaard, K. Grove-Rasmussen, M. H. Madsen, and J. Nygaard, Tuning yu-shiba-rusinov states in a quantum dot, Phys. Rev. B 94, 064520 (2016).
  • Bargerbos et al. [2022] A. Bargerbos, M. Pita-Vidal, R. Zitko, J. Avila, L. J. Splitthoff, L. Grunhaupt, J. J. Wesdorp, C. K. Andersen, Y. Liu, L. P. Kouwenhoven, R. Aguado, A. Kou, and B. van Heck, Singlet-doublet transitions of a quantum dot josephson junction detected in a transmon circuit, PRX Quantum 3, 030311 (2022).
  • Bauer et al. [2007] J. Bauer, A. Oguri, and A. C. Hewson, Spectral properties of locally correlated electrons in a bardeen–cooper–schrieffer superconductor, Journal of Physics: Condensed Matter 19, 486211 (2007).
  • Meng et al. [2009a] T. Meng, S. Florens, and P. Simon, Self-consistent description of andreev bound states in josephson quantum dot devices, Physical Review B 79, 224521 (2009a).
  • Meng et al. [2009b] T. Meng, S. Florens, and P. Simon, Self-consistent description of andreev bound states in josephson quantum dot devices, Phys. Rev. B 79, 224521 (2009b).
  • Lawrie et al. [2020] W. I. L. Lawrie, H. G. J. Eenink, N. W. Hendrickx, J. M. Boter, L. Petit, S. V. Amitonov, M. Lodari, B. Paquelet Wuetz, C. Volk, S. G. J. Philips, G. Droulers, N. Kalhor, F. van Riggelen, D. Brousse, A. Sammak, L. M. K. Vandersypen, G. Scappucci, and M. Veldhorst, Quantum dot arrays in silicon and germanium, Applied Physics Letters 116, 080501 (2020).
  • Sammak et al. [2019] A. Sammak, D. Sabbagh, N. W. Hendrickx, M. Lodari, B. Paquelet Wuetz, A. Tosato, L. Yeoh, M. Bollani, M. Virgilio, M. A. Schubert, P. Zaumseil, G. Capellini, M. Veldhorst, and G. Scappucci, Shallow and undoped germanium quantum wells: A playground for spin and hybrid quantum technology, Advanced Functional Materials 29, 1807613 (2019).
  • Lee et al. [2014] E. Lee, X. Jiang, and M. e. a. Houzet, Spin-resolved andreev levels and parity crossings in hybrid superconductor–semiconductor nanostructures, Nature Nanotechnology 9, 79–84 (2014).
  • Grove-Rasmussen et al. [2009] K. Grove-Rasmussen, H. I. Jorgensen, B. M. Andersen, J. Paaske, T. S. Jespersen, J. Nygaard, K. Flensberg, and P. E. Lindelof, Superconductivity-enhanced bias spectroscopy in carbon nanotube quantum dots, Phys. Rev. B 79, 134518 (2009).
  • Yu [1965] L. Yu, Bound states in paramagnetic impurity-containing superconductors, Chinese Jounral of Physics 21, 21 (1965).
  • Shiba [1968] H. Shiba, Classical Spins in Superconductors, Progress of Theoretical Physics 40, 435 (1968).
  • Rusinov [1969] A. I. Rusinov, Superconcductivity near a Paramagnetic Impurity, Letters to JETP 9, 146 (1969).
  • Andersen et al. [2011] B. M. Andersen, K. Flensberg, V. Koerting, and J. Paaske, Nonequilibrium transport through a spinful quantum dot with superconducting leads, Phys. Rev. Lett. 107, 256802 (2011).
  • Baran et al. [2023] V. V. Baran, E. J. Frost, and J. Paaske, Surrogate model solver for impurity-induced superconducting subgap states, Physical Review B 108, L220506 (2023).
  • Tinkham [2004] M. Tinkham, Introduction to Superconductivity, 2nd ed. (Dover Publications, 2004).
  • Fulde [1973] P. Fulde, High field superconductivity in thin films, Advances in Physics 22, 667 (1973).
  • van Riggelen et al. [2021] F. van Riggelen, N. W. Hendrickx, W. I. L. Lawrie, M. Russ, A. Sammak, G. Scappucci, and M. Veldhorst, A two-dimensional array of single-hole quantum dots, Applied Physics Letters 118, 044002 (2021).
  • de Moor et al. [2018] M. W. A. de Moor, J. D. S. Bommer, D. Xu, G. W. Winkler, A. E. Antipov, A. Bargerbos, G. Wang, N. van Loo, R. L. M. O. het Veld, S. Gazibegovic, D. Car, J. A. Logan, M. Pendharkar, J. S. Lee, E. P. A. M. Bakkers, C. J. Palmstrøm, R. M. Lutchyn, L. P. Kouwenhoven, and H. Zhang, Electric field tunable superconductor-semiconductor coupling in majorana nanowires, New Journal of Physics 20, 103049 (2018).
  • Antipov et al. [2018] A. E. Antipov, A. Bargerbos, G. W. Winkler, B. Bauer, E. Rossi, and R. M. Lutchyn, Effects of gate-induced electric fields on semiconductor majorana nanowires, Phys. Rev. X 8, 031041 (2018).
  • Hendrickx et al. [2020a] N. Hendrickx, D. Franke, and A. e. a. Sammak, Fast two-qubit logic with holes in germanium, Nature , 487–491 (2020a).
  • Hendrickx et al. [2020b] N. Hendrickx, D. Franke, and A. e. a. Sammak, A single-hole spin qubit, Nature Communications  (2020b).
  • Hendrickx et al. [2023] N. W. Hendrickx, L. Massai, M. Mergenthaler, F. Schupp, S. Paredes, S. W. Bedell, G. Salis, and A. Fuhrer, Sweet-spot operation of a germanium hole spin qubit with highly anisotropic noise sensitivity (2023), arXiv:2305.13150 [cond-mat.mes-hall] .
  • Chandra et al. [2013] P. Chandra, P. Coleman, and R. Flint, Hastatic order in the heavy-fermion compound uru2si2, Nature 493, 621 (2013).
  • Bosco et al. [2021] S. Bosco, M. Benito, C. Adelsberger, and D. Loss, Squeezed hole spin qubits in ge quantum dots with ultrafast gates at low power, Phys. Rev. B 104, 115425 (2021).
  • Laubscher et al. [2024] K. Laubscher, J. D. Sau, and S. Das Sarma, Germanium-based hybrid semiconductor-superconductor topological quantum computing platforms: Disorder effects, arXiv preprint arXiv:2404.16285  (2024).

VII Methods

VII.1 Fabrication

The device is fabricated on a Ge/SiGe heterostructure, which is grown on n-type Si(001) substrate using a reduced pressure chemical vapor deposition reactor. The stack comprises of a reverse graded Si0.2Ge0.8 virtual substrate, a 16 nm Ge quantum well, a 27 nm Si0.2Ge0.8 barrier, and a less than 1 nm Si sacrificial cap [41]. The device is fabricated in three electron-beam lithography defined layers separated by two oxide layers of similar-to\sim 7 nm Al2O3 grown by atomic layer deposition at 300superscript300300^{\circ}300 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPTC. The first layer forms the PtGeSi contacts after lithography; a wet etch is performed using buffered HF (6%similar-toabsentpercent6\sim 6\%∼ 6 %) solution to remove the sacrificial cap before a 15 nm Pt layer is deposited via e-gun evaporation at a pressure of 1×107cross-product1superscript1071\crossproduct 10^{-7}1 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT mbar. The Pt contacts then undergo rapid thermal annealing at 400superscript400400^{\circ}400 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPTC for 15 min in argon atmosphere. Barrier gates and plunger gates are deposited using e-gun evaporation and consist of a Ti (5 nm) sticking layer and a Pd layer (25 nm and 29 nm for the barrier and plunger layers respectively).

VII.2 Measurement

The device is measured inside a sample puck loaded into a Bluefors XLD dilution refrigerator, at a mixing chamber temperature of similar-to\sim 9 mK. The dilution refrigerator is equipped with a (1-1-6) T vector magnet, the sample chip being placed such that the 6 T direction is in-plane. We expect the main source of magnetic field misalignment to result from misalignment of the sample board inside the puck (QDevil QBoard sample holder) during sample loading. We estimate this error to be less than 5superscript55^{\circ}5 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT. Differential conductance measurements are taken using a Stanford-Research Systems SR860 lock-in amplifier at a frequency of 119 Hz, and amplitude of 2.5 μVμ𝑉\upmu Vroman_μ italic_V. A line resistance of 2.7 kΩΩ\Omegaroman_Ω is subtracted in all measurements. As we use a superconducting tunnel probe for our measurements, tunnelling events between the superconducting lead and the quantum dot manifest at an offset energy of Δ0=subscriptΔ0absent\Delta_{0}=roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 72 μμ\upmuroman_μeV. Bias spectroscopy measurements contain a small thermally induced bias offset of 18 μμ\upmuroman_μeV, which has been subtracted from all relevant data. DC gate voltages are applied via a QDevil QDAC-II voltage source.

In addition to lock-in measurements, radiofrequency reflectometry was performed as described in the Supplementary Information, using a FPGA with a built in microwave signal generator as well as signal demodulator (Quantum Machines OPX+). A tank circuit with a resonance frequency of 192.3 MHz was used, connected to the drain contact. The incident rf-signal is attenuated by 40 dB at various plates of the cryostat, and undergoes an additional attenuation of 20 dB from a directional coupler at the MXC plate (See Supplementary Section Figure 2). The reflected signal undergoes 40 dB of amplification at the 4 K stage via a HEMT amplifier, and passes through a DC block and bias-tee at the input of the demodulation circuit. A 50 ΩΩ\Omegaroman_Ω terminator is connected to the DC side of the bias-tee.