Abstract
In Josephson diodes the asymmetry between positive and negative current branch of the current-phase relation leads to a polarity-dependent critical current and Josephson inductance. The supercurrent nonreciprocity can be described as a consequence of the anomalous Josephson effect —a φ0-shift of the current-phase relation— in multichannel ballistic junctions with strong spin-orbit interaction. In this work, we simultaneously investigate φ0-shift and supercurrent diode efficiency on the same Josephson junction by means of a superconducting quantum interferometer. By electrostatic gating, we reveal a direct link between φ0-shift and diode effect. Our findings show that spin-orbit interaction in combination with a Zeeman field plays an important role in determining the magnetochiral anisotropy and the supercurrent diode effect.
Similar content being viewed by others
Introduction
In solids, spin-orbit interaction (SOI) makes it possible to control orbital degrees of freedom by acting on the electron spin, and vice-versa1,2. In superconductors3, the impact of SOI can be particularly spectacular, since it enables phenomena which go beyond the realm of conventional s-wave superconductors, as e.g. topological phases4, finite-momentum superconductivity5,6,7, Lifshitz invariant8,9,10, Ising superconductivity11, anomalous Josephson effect12,13,14,15,16, and intrinsic supercurrent diode effect17,18,19,20,21,22,23,24,25,26,27,28,29,30,31. In what follows, we shall focus on the last two effects and on their relation in Josephson junctions.
The anomalous Josephson effect manifests itself in a phase offset φ0 at zero current, I(φ0) = 0, in the current-phase relation (CPR)15,16,32,33,34,35,36. This also implies a finite supercurrent at zero phase difference I(φ = 0) ≠ 0. The effect requires the simultaneous breaking of both inversion and time-reversal symmetry37, which can be provided by SOI and Zeeman field, respectively.
The same symmetries need to be broken in order to observe the supercurrent diode effect (SDE), namely, the dependence of the critical current on the bias polarity. This effect can be trivially obtained, e.g., in asymmetric superconducting quantum interference devices (SQUIDs)38 or, more generally, when in a film an inhomogeneous supercurrent distribution is coupled to a flux. Recently, it was shown17,18,19,22,23,24,25,26,27,30,31 that supercurrent rectification can as well emerge as an intrinsic feature of homogeneous quasi-2D systems subjected to a Zeeman field. Such nontrivial SDE is a new precious probe of the condensate physics (in films)10 and of Andreev bound states (ABSs) in Josephson junctions27,30, including possible topological properties39,40,41.
Several mechanisms have been proposed to explain such intrinsic SDE in films10,42,43,44,45,46 and Josephson junctions47,48,49,50,51,52,53. In superconducting-normal-superconducting (SNS) junctions the supercurrent can be computed in terms of the ABSs in the N weak link. In experiments, ABSs can be directly probed by tunnel spectroscopy36,54,55, while their effect on the CPR can be deduced from IV-characteristics, inductance versus current measurements18,27,56 and SQUID experiments57.
Early observations on Josephson diodes were interpreted in terms of φ0-shift in ballistic systems with skewed CPR18,22,23,24,56. Within this picture, the SDE ultimately originates (as it does the φ0-shift) from SOI. An alternative model, proposed by Banerjee et al.30 based on the theory of ref. 50, explains the same effect in terms of a purely orbital mechanism. To date, it is not clear yet to which extent the two mechanisms (namely, the SOI-based and the purely orbital mechanism) contribute to the supercurrent rectification observed in experiments.
In this work, we make use of an asymmetric SQUID with mutually orthogonal junctions to directly measure both the anomalous φ0-shift and the SDE on the same junction. By gating, we can electrostatically control both effects and highlight their relation. Finally, by measuring the temperature dependence of the φ0-shift and of the diode efficiency we highlight the role of the higher harmonics of the CPR for the emergence of the SDE. We comment on our results in light of alternative models proposed in the literature and compare the temperature dependence of φ0 and the SDE to the predictions of a minimal theoretical model.
Figure 1a shows a scheme of our SQUID. The device is fabricated starting from a molecular beam epitaxy-grown heterostructure featuring an InGaAs/InAs/InGaAs quantum well, capped by a 5-nm-thick epitaxial Al film18,58,59,60. The quantum well hosts a 2D electron gas (2DEG) with a proximity-induced superconducting gap inherited from the Al film. By deep wet etching, we define an asymmetric SQUID loop. The actual geometry is shown in the false-color scanning electron microscopy image in Fig. 1b, where the turquoise areas indicate the pristine Al/InGaAs/InAs/InGaAs regions, while the gray areas refer to deeply etched (insulating) regions. To obtain the two normal (N) weak links, we selectively etch the Al film (yellow areas in Fig. 1b). The reference Josephson junction 1 (JJ1) is 28-μm-wide and 120-nm-long, whereas the Josephson junction 2 (JJ2) is 2.7-μm-wide and 100-nm-long. Finally, a gate is fabricated on top of JJ2, which allows us to control the electron density in the N-link and thus the critical current Ic,2 of this junction. The two junctions are mutually perpendicular, so that an in-plane magnetic field \({\overrightarrow{B}}_{{{{{{{{\rm{ip}}}}}}}}}\) parallel to the short junction, i.e., along \(\hat{y}\), (Fig. 1a, b) will induce magnetochiral effects18 in the short junction only, and not in the reference junction. Here, we take as positive \(\hat{z}\) direction that perpendicular to the 2DEG and directed from the substrate towards the Al (Fig. 1a), which corresponds to the direction opposite to the built-in electric field in the quantum well18,61 which provides SOI in the 2DEG18,61.
a (Left) Schematic illustration of the device. The SQUID consists of a large reference junction (JJ1) and of a small junction (JJ2) which is coupled to a gate. By applying a magnetic field along \(\hat{y}\), a φ0-shift is induced in the CPR of the latter junction, which can be controlled by electrostatic gating. (Right) Scheme of the topmost layers of the heterostructure. The black arrow indicates the direction of the positive current bias I. b False-color scanning electron microscopy image of the device taken before gate patterning. The pristine superconducting Al/InAs leads are highlighted in turquoise, the areas where Al is selectively etched in yellow (including the weak links, highlighted by the black arrows). The remaining parts in gray correspond to deeply etched regions, where both the Al film and the topmost semiconducting layers are etched. c The color plot shows the SQUID differential resistance versus out-of-plane field Bz and current I, for θ = − 90° and Bip = 100 mT (i.e., Bx = 0, By = −100 mT) at T = 40 mK. The white arrow indicates where anomalous phase shifts were measured (see panels d–f). d SQUID critical current Ic as a function of Bz, for θ and Bip as in c. The different curves refer to different gate voltages Vg. We define B0(Vg) as the crossing of each curve with the Vg = −2.5 V reference curve (black). e, f Corresponding measurements for the same Bip = 100 mT but, respectively, θ = 180° and θ = 90°. g Plot of Δφ0(Vg) ≡ 2πA[B0(Vg) − B0(Vg = −1.5 V)]/Φ0, for different \({\overrightarrow{B}}_{{{{{{{{\rm{ip}}}}}}}}}\) orientations, i.e., for different θ. Here, A is the loop area and Φ0 the flux quantum.
We measure differential resistance in a 4-terminal geometry as a function of DC current. In what follows, we indicate as Ic the (measured) SQUID critical current (see Methods) and as Ic,i (i = 1, 2) the (deduced) critical current in junction i. When needed, we use the superscript + (−) to indicate positive (negative) current from source to drain (Fig. 1a), i.e., supercurrent in the positive (negative) x direction in JJ2. For the in-plane field, we interchangeably use either Cartesian components, or magnitude and angle parametrization, i.e. \({\overrightarrow{B}}_{{{{{{{{\rm{ip}}}}}}}}}={B}_{x}\hat{x}+{B}_{y}\hat{y}={B}_{{{{{{{{\rm{ip}}}}}}}}}(\cos \theta,\sin \theta )\), see cartesian axes in Fig. 1a.
Results
Figure 1c shows the color plot of the differential resistance versus out-of-plane field Bz and DC current I, measured with an applied field By = − 100 mT at T = 40 mK. The fast oscillations have period 1.8 μT, corresponding to a flux quantum Φ0 = h/2e applied to the loop. Such oscillations are superimposed to the Fraunhofer pattern of the reference junction, whose central and first side lobes are visible. The plot displays an evident asymmetry around the Bz = 0 and I = 0 axes, while it is approximately point-inversion-symmetric around the origin, namely Ic(Bz) ≈ − Ic(−Bz). The dominant extrinsic SDE of the SQUID as a whole (since Ic(Bz) ≠ −Ic(Bz)) is a long-known consequence of screening in the loop, which must be kept distinct from the intrinsic SDE in single homogeneous junctions, which is the effect we are studying. As discussed in the Supplementary Information, screening in SQUIDs arises when there are current-dependent corrections to the fundamental SQUID relation
where γi it the gauge-invariant phase drop at the i-th junction, Φ0 is the flux quantum and Φ the flux through the loop. The corrections originate either from the additional flux induced by the SQUID current itself, or by the phase drop accumulated along the loop arms38,62. The former correction is proportional to the geometric loop inductance, while the latter is proportional to the kinetic inductance of the SQUID arms63. As discussed in the Methods, in our sample, screening is mostly dominated by the large kinetic inductance of the thin Al film, whose sheet inductance is L□ ~ 30 pH. The geometric loop inductance is comparatively small, see Supplementary Information. As shall be discussed below, screening effects hinder the measurement of both the absolute value of the anomalous shift φ0 and that of the diode efficiency. Indeed, the determination of the absolute φ0 is challenging even in SQUID devices with low screening and reference devices32,36. One of the difficulties is the fact that an accurate determination of φ0 requires a reproducible Bz control on the microtesla scale, while an in-plane field of the order of tens or hundreds of millitesla is swept.
The determination of the absolute φ0 is usually difficult in systems with large kinetic inductance. Similar to refs. 33,34,36, we shall measure the relative shift of the CPR with respect to that for large negative gate voltage. Figure 1d shows SQUID oscillations measured for different gate voltages Vg. The measurement is performed with an in-plane field Bip = 100 mT applied perpendicular to JJ2 (θ = −90°) at T = 40 mK. The oscillations are taken near the maximum of the tilted Fraunhofer pattern, indicated by the upper arrow in Fig. 1c. The black curve (Vg = −2.5 V) refers to a completely pinched-off JJ2 (Ic,2 = 0), and serves as zero-current baseline. The first curve for which oscillations are clearly visible is that for Vg = −1.5 V. We shall label as B0(Vg) the crossing with positive slope of each Ic(Vg) curve with the baseline Ic(Vg = −2.5 V), see Fig. 1d. As discussed in the Supplementary Information, at these crossings the supercurrent I2 in JJ2 vanishes, therefore its gauge invariant phase difference is, by definition, the anomalous shift φ0. We consider variations of φ0 with respect to the reference voltage Vg = −1.5 V36, namely, Δφ0 ≡ 2πAloop[B0(Vg) − B0(Vg = −1.5 V)]/Φ0, where Aloop = 1150 μm2 is the loop area.
Figure 1e and f show the results of the same measurements after two subsequent -90° sample rotations, namely, for θ = 180° and θ = 90°, respectively. From Fig. 1e we deduce that for \({\overrightarrow{B}}_{{{{{{{{\rm{ip}}}}}}}}}\parallel \overrightarrow{I}\), the curves cross the baseline with positive slope nearly at the same Bz (i.e., B0(Vg) is constant). Instead, for \({\overrightarrow{B}}_{{{{{{{{\rm{ip}}}}}}}}}\perp \overrightarrow{I}\), B0(Vg) monotonically increases (decreases) with Vg for negative (positive) sign of \({\hat{e}}_{z}\cdot ({\overrightarrow{B}}_{{{{{{{{\rm{ip}}}}}}}}}\times \overrightarrow{I})\). This is a clear signature of the magnetochiral nature of the anomalous Josephson effect18.
The variation of Δφ0(Vg) for different \({\overrightarrow{B}}_{{{{{{{{\rm{ip}}}}}}}}}\) orientations (i.e., for different θ with \(| {\overrightarrow{B}}_{{{{{{{{\rm{ip}}}}}}}}}|=100\) mT) is plotted in Fig. 1g. We stress that, since we subtract φ0(Vg = −1.5 V) (as in the definition of Δφ0), what is important in Fig. 1g is the monotonic increase or decrease of Δφ0 with the gate voltage. The graph clearly shows the proportionality of Δφ0 to18\(-{\hat{e}}_{z}\cdot ({\overrightarrow{B}}_{{{{{{{{\rm{ip}}}}}}}}}\times \overrightarrow{I})\), as expected by SOI-based models for the anomalous Josephson effect12. A graph of Δφ0 versus the angle θ is shown in Fig. S4 of the Supplemental Information. As expected, we find the sinusoidal dependence \(\Delta {\varphi }_{0} \sim -\sin (\theta )\). For all curves, the magnitude of ∣Δφ0∣ increases with Vg. The monotonic increase is expected, since in this type of InGaAs/InAs/InGaAs quantum wells, a positive gate voltage increases the built-in electric field61 responsible for the Rashba SOI18,61.
The main goal of our experiments is to establish a relation between the anomalous φ0-shift and intrinsic SDE by measuring both phenomena on the same junction. For this purpose, we investigate SQUID oscillations for both current bias polarities in order to deduce both the positive (\({I}_{c,2}^{+}\)) and the negative (\({I}_{c,2}^{-}\)) critical current of JJ2. Figure 2a shows the SQUID interference pattern measured in the vicinity of Bz = 0 for By = − 50 mT at Vg = 1.0 V (red), together with the reference curve at Vg = − 2.5 V (black, where Ic,2 = 0 and SQUID oscillations vanish). Figure 2b shows several \({I}_{c}^{+}({B}_{z})\) curves for different gate voltages Vg, measured at By = − 50 mT and T = 40 mK, as in Fig. 2a. We focus on one particular oscillation highlighted by the upper arrow in Fig. 2a. We stress that, owing to screening, one cannot extract the negative critical current in the CPR from the relative minima of \({I}_{c}^{+}({B}_{z})\). As discussed in the Supplementary Information, it is instead necessary to look at the maxima of \(| {I}_{c}^{-}(-{B}_{z})|\). The corresponding \({I}_{c}^{-}\) curves for the opposite Bz range (lower arrow in Fig. 2a) are shown in panel c. We use the baseline curve in Fig. 2a (black, Vg = −2.5 V) as a reference to extract \({I}_{c,2}^{+}\) from data in Fig. 2b and \({I}_{c,2}^{-}\) from Fig. 2c. For each Vg, \({I}_{c,2}^{+}\) corresponds to the maximum in Bz (see red arrow in Fig. 2b) of the difference \({I}_{c}^{+}({B}_{z},{V}_{g})-{I}_{c}^{+}({B}_{z},{V}_{g}=-2.5\,\,{{\mbox{V}}})\). \({I}_{c,2}^{-}\) is deduced in a similar way from \({I}_{c}^{-}({B}_{z},{V}_{g})-{I}_{c}^{-}({B}_{z},{V}_{g}=-2.5\,\,{{\mbox{V}}})\).
a Positive and negative SQUID critical current for an out-of-plane field Bz close to zero, T = 40 mK, and By = − 50 mT. The red (black) curve refers to a gate voltage Vg = 1.0 V (Vg = −2.5 V). The upper and lower arrow indicates the oscillation highlighted in panel b and c, respectively. They are located asymmetrically in Bz, eliminating the trivial diode effect of the background. b Positive critical current \({I}_{c}^{+}\) versus out-of-plane field Bz. The red arrow indicates \({I}_{c,2}^{+}\) for the Vg = 1 V-curve, see text. c Negative critical current \({I}_{c}^{-}\) versus Bz. The red arrow indicates \({I}_{c,2}^{+}\) for the Vg = 1 V-curve. The different curves in b and c correspond to different values of the gate voltage Vg. Measurements are performed at T = 40 mK and By = − 50 mT. d \({I}_{c,2}^{+}\) and \({I}_{c,2}^{-}\) versus Vg for By = − 50 mT. e–i Diode efficiency \(\eta \equiv 2({I}_{c}^{+}-| {I}_{c}^{-}| )/({I}_{c}^{+}+| {I}_{c}^{-}| )\) versus Vg, for By = −50 mT (e), −30 mT (f), 0 mT (g), 30 mT (h), 50 mT (i). j Slope of the diode efficiency dη/dVg, plotted versus By.
The resulting \({I}_{c,2}^{+}({V}_{g})\) and \({I}_{c,2}^{-}({V}_{g})\) are plotted in Fig. 2d. As a figure of merit for the supercurrent rectification, we use the supercurrent diode efficiency \(\eta \equiv ({I}_{c,2}^{+}-| {I}_{c,2}^{-}| )/\langle {I}_{c,2}\rangle\), with \(\langle {I}_{c,2}\rangle \equiv ({I}_{c,2}^{+}+| {I}_{c,2}^{-}| )/2\). The efficiency η(Vg) is plotted in Fig. 2e–i for different values of By. As explained in the Supplementary Information, η is subjected to a spurious vertical shift owing to SQUID screening effects. On the other hand, its slope as a function of Vg, i.e. ∂η/∂Vg remains unaffected by screening. Our crucial assumption is that, seen as functions of\(\overrightarrow{B}\) orT, η and ∂η/∂Vg are proportional, i.e., the larger η, the larger is its modulation of as a function of Vg. Thus, we shall use ∂η/∂Vg as a convenient figure of merit for the magnetochiral effect when considering field or temperature dependence.
Previous reports on Josephson diodes18,22,24,25 showed that the rectification efficiency is linear in By, but only up to a certain threshold (here at ∣By∣ = 50 mT): after that, a clear suppression is observed. Our data in Fig. 2j, deduced from data in Fig. 2e–i, show precisely the same behavior. The suppression at higher field was interpreted in ref. 27 as a signature of a 0-π-like transition, occurring when two minima in the Josephson energy-phase relation become degenerate.
The results shown so far provide strong evidence of the link between anomalous phase shift and SDE. Both effects are linear in By. Both Δφ0 and η can be modulated by a gate voltage in the accessible range Vg ∈ [-1.5 V, 1.0 V]. This is consistent with the SOI-based model10,18,24,53,56 of the SDE, where the supercurrent rectification is accompanied by the anomalous Josephson shift φ0. However, the φ0-shift is, per se, not a sufficient condition for the supercurrent nonreciprocity. To break the symmetry between positive and negative part of the CPR, multiple ballistic channels are needed14,27. In finite-width junctions with parabolic dispersion, each i-channel has a different φ0,i-shift, owing to the different Fermi velocity vF,i: in fact, for a single ballistic channel, \({\varphi }_{0,i}\propto {v}_{F,i}^{-2}\)12. The total CPR is the sum of all single-channel CPR contributions Ii(φ + φ0,i). The sum of skewed CPRs with different φ0,i-shift leads to an asymmetric total CPR, whose positive and negative branch are different. The skewness (i.e., the content of higher harmonics) of the individual CPRs is crucial, since otherwise the sum of sinusoidal CPRs would always lead to a sinusoidal –i.e., reciprocal– CPR. The relation between η and φ0 is evident from the comparison between Fig. 1g and Fig. 2e–i. On the other hand, since the SDE also relies on the presence of higher harmonics in the CPR18, we expect that η (and thus ∂η/∂Vg) will be highly sensitive to the junction transparency and temperature, as opposed to the anomalous shift φ0.
Figure 3a and b respectively show the temperature dependence of φ0 and η as obtained from a simple theoretical model (described in Methods and in refs. 27,53) for a Josephson junction in the short-ballistic limit with Zeeman interaction. The junction separates two semi-infinite superconducting 2DEGs with zero-temperature superconducting gap Δ*(0) ≈ 130 μeV60, and critical temperature Tc ≈ 2 K. We note that the anomalous shift φ0 is nearly T-independent, whereas the supercurrent rectification is strongly suppressed already for T > 100 mK.
a Computed temperature dependence of the (sign changed) anomalous phase shift φ0 for a Zeeman parameter of λZ = − 0.40 (which would correspond to By = − 100 mT for g-factor ∣g∣ = 12). We assume Tc = 2.0 K as in the experiments. Inset: full-range φ0(T) graph plotted for T up to T = Tc. b Computed temperature dependence of the supercurrent diode rectification efficiency − η for the same parameters as in a. Inset: full-range − η(T) graph up to T = Tc. c Measured Δφ0 (as defined in the text) at Vg = 1.0 V and By = − 100 mT, plotted as a function of temperature. d Temperature dependence of dφ0/dVg (as defined in the text) at By = 50 mT. For ease of comparison, all main graphs are plotted up to T = 0.8 K, the highest temperature for which Δφ0 and dφ0/dVg was measurable.
In Fig. 3c and d we show the measured temperature dependence of Δφ0 and that of ∂η/∂Vg, which we take as measures of φ0(T) and η(T), respectively. Δφ0(T) is measured at By = − 100 mT, (Bip = 100 mT, θ = − 90°), where the field magnitude is set large enough to limit screening effects to an acceptable level. Instead, ∂η/∂Vg(T) is measured at By = 50 mT, where the SDE is maximal, see Fig. 2j. Figure 3c shows that Δφ0 is temperature-independent within the experimental accuracy. In contrast, the (gate modulation of the) supercurrent rectification is clearly suppressed already at temperatures well below Tc, as shown in Fig. 3d. Both observations match the corresponding theory predictions.
A comment is in order about the sign and magnitude of the effects. Both our experimental data and analytical model show that if the Rashba SOI-inducing electric field is directed along \(-\hat{z}\) (as in ref. 61 and in the Supplementary Information of ref. 18), \({\overrightarrow{B}}_{{{{{{{{\rm{ip}}}}}}}}}\) along \(+\hat{y}\), and the positive current bias along \(+\hat{x}\), then φ0 < 0 [where I(φ0) = 0, ∂φI(φ0) > 0] and η > 0. Instead, the magnitude of φ0 and η predicted by ballistic theory12 is smaller than the one measured in our and in other experiments in the literature32,33. A possible explanation for this discrepancy could be disorder in and near the junction32, since diffusive models predict a much larger φ0. Another possibility is the enhancement of SOI due to interaction of quantum well electrons with the image charges that are formed in a nearby Al gate. The nontrivial property of the induced image-potential that depends on the electron density of 2DEG gives a feedback on SOI that superimposes with the innate Rashba SOI of the quantum well without metallic gate as demonstrated in refs. 64,65,66.
Discussion
The main goal of our study is to elucidate the physical mechanism behind the intrinsic SDE in single, homogeneous Josephson junctions. The effect has been so far explained by two different models: one14,18,56 is based on the combination of Rashba SOI plus Zeeman interaction (due to an external in-plane field or exchange interaction); the other is a purely orbital mechanism30,50 based on the finite Cooper pair momentum induced in the superconducting leads by the flux associated to the in-plane field30. This flux is finite if the parent superconducting film and the 2DEG are spatially separated.
The main difference between the two pictures is the expected dependence on Vg. Such dependence naturally emerges since Vg affects the band alignment and thus the Rashba coefficient αR. Both αR and the electron density n critically affect φ0, which, in turn, determines η in multichannel systems. In contrast, the orbital mechanism30 hardly depends on the gate voltage (Max Geier, Karsten Flensberg, private communication). As discussed in the Supplementary Information, the gate voltage also affects the magnetochiral anisotropy for the inductance18, an effect that is strictly related to the supercurrent rectification27. The observed strong gate dependence of φ0 and η indicates that the Rashba-based mechanism must certainly play an important role in the SDE. On the other hand, the orbital mechanism cannot be ruled out by our observations: it could still coexist with the spin-orbit-based mechanism.
Finally, we would like to stress that, even though we make use of a SQUID to link η to φ0, our point does not concern the (trivial and long known38) SDE of the asymmetric SQUID as a whole. Our focus is exclusively on the intrinsic SDE in a single, homogeneous junction (JJ2).
In conclusion, we have shown the coexistence of anomalous Josephson effect and supercurrent rectification by measuring both effects on the same Josephson junction embedded in a SQUID. The observed gate voltage and temperature dependence are compatible with a spin-orbit based picture where supercurrent rectification arises in multichannel junctions with anomalous shift φ0 and skewed current-phase relation.
Josephson diodes based on φ0-junctions are important for both fundamental research and applications, e.g. as sensors for readout of racetrack memory devices67,68. They are novel and powerful probes of symmetry breaking in 2D superconductors21,49,69 and possible probes of topological phase transitions39. A recent proposal70 suggested that the anomalous Josephson effect might be used in multiterminal junctions to obtain compact nonreciprocal devices as, e.g., circulators for rf-applications71.
Methods
Experimental methods
The heterostructure is grown by molecular beam epitaxy. The full layer sequence is reported in the Supplementary Information. The most relevant layers are the topmost ones, namely, the nominally 5-nm-thick Al film at the sample surface, a 10-nm-thick In0.75Ga0.25 As layer acting as a barrier, a 7-nm InAs layer hosting the 2DEG, followed by another In0.75Ga0.25As barrier of thickness 4 nm. Structures are defined by electron beam lithography. The selective etching of Al is performed by wet chemical etching using Transene D. Deep etching processes (where the 2DEG is removed altogether) is performed using a phosphoric acid-based solution.
Transport measurements are performed in a 4-point configuration using standard lock-in techniques. To determine the SQUID critical current \({I}_{c}^{+}\) and \({I}_{c}^{-}\) as defined in the text we take dV/dI = 6 Ω as a threshold.
To determine the correct offset for Bz we look at the symmetry of the plot of R = dV/dI versus I and Bz, see e.g., Fig. 1c. Since R(I, Bz) ≈ R(−I, −Bz), the center of inversion symmetry of the plot allows us to determine the applied out-of-plane field which corresponds to an effective Bz = 0.
Circuit model
A circuit model of the asymmetric SQUID device, which includes the effect of inductive screening in the loop and the reference junction is described in the Supplementary Information. The model can reproduce the experimentally found Ic(Bz) curves when using realistic parameters for the kinetic inductance of the aluminum electrodes. A detailed discussion of the model and simulation results are provided in the Supplemental Information. The most important findings from the simulations regarding our data evaluation are:
-
With our device parameters, inductive screening effects do not affect the extraction of φ0. When φ0 = 0, we find that all curves intersect the reference line (Ic,2 = 0) in the same point.
-
The critical current Ic,2 extracted from the SQUID oscillations is proportional to the actual critical current of JJ2. In general, the proportionality constant can be slightly different for the positive and negative critical current. For the evaluation of η(Vg) this means that we can only analyze dη/dVg, as η(Vg) will have an uncontrolled offset.
Theoretical methods
Our theoretical model, initially developed in refs. 27,53, describes the experimentally relevant system in terms of a short S–N–S Josephson junction that couples two semi-infinite s-wave superconducting (S) regions with inherently strong Rashba SOI through a thin delta-like normal-conducting (N) link. Nontrivial solutions of the 2D Bogoliubov–de Gennes equation72
with the single-electron Hamiltonian
determine the energies E and wave functions Ψ(x, y) of the Andreev bound states73, which are at the heart of the coherent Cooper-pair supercurrent transport along the \(\hat{x}\)-direction; \(\hat{\Delta }(x)\) corresponds to the s-wave superconducting pairing potential that we approximate by
where \({\Delta }^{*}(T)={\Delta }^{*}(0)\tanh (1.74\sqrt{{T}_{{{{{{{{\rm{c}}}}}}}}}/T-1})\) approximates the temperature-dependent proximity-induced superconducting gap [from the experimental data, the induced gap at zero temperature was estimated as Δ*(0) ≈ 130 μeV and the critical temperature as Tc ≈ 2 K] and φ indicates the phase difference between the two superconducting regions. The Rashba SOI that is present throughout the whole system is parameterized by αR, V0 and VZ represent the scalar (spin-independent) and Zeeman (spin-dependent) potentials inside the delta-like N link of thickness d—the magnetic field causing the Zeeman splitting is thereby aligned perpendicular to the current direction (i.e., along \(\hat{y}\))—, μ is the Fermi energy, m the (effective) quasiparticle mass, and \({\hat{\sigma }}_{0}\) and \({\hat{\sigma }}_{i}\) refer to the 2 × 2 identity and ith Pauli spin matrix, respectively.
After determining the Andreev-state energies En(φ) (with n labeling the n-th tranverse channel) as a function of the superconducting phase difference φ, we apply the quantum-mechanical current operator to the corresponding bound-state wave functions inside the N link to compute in the first step the Josephson CPRs I(φ) and obtain in the second step the direction-dependent critical currents necessary to quantify the SDE. In the simultaneous presence of SOI and Zeeman interaction, the bound-state energies depend on φ through En(φ) = Δ*(T)f(φ), where the generic function f(φ) is no longer antisymmetric with respect to φ, i.e., f( − φ) ≠ f(φ), reflecting the broken space-inversion and time-reversal symmetries, and the therefrom resulting nontrivial φ0-phase shifts. Note that the only impact of temperature on En(φ) is an effective rescaling (i.e., suppression with increasing temperature) of the superconducting-gap amplitude Δ*(T) according to \({\Delta }^{*}(T)={\Delta }^{*}(0)\tanh (1.74\sqrt{{T}_{{{{{{{{\rm{c}}}}}}}}}/T-1})\), whereas the qualitative shape of En(φ) is not altered by temperature. The total Josephson current is then given by
where the sum over n ensures to account for the current contributions of all bound states (i.e., from all transverse channels of the junction) and kB indicates the Boltzmann constant. The current at zero temperature can, in the simplest case, be extracted from the thermodynamic relation74
where E = ∑nEn and e is the positive elementary charge. The major temperature effect on the Josephson current originates, therefore, from the suppression of the higher-harmonic contributions in Eq. (6) due to the tanh term in Eq. (5).
The strengths of the Rashba SOI, the scalar (barrier), and the Zeeman potentials are measured by the dimensionless parameters λSOI = mαR/(ℏ2kF), Z = 2mV0d/(ℏ2kF), and λZ = 2mVZd/(ℏ2kF), respectively, where \({k}_{{{{{{{{\rm{F}}}}}}}}}=\sqrt{2m\mu }/\hslash\) refers to the Fermi wave vector (for the experimental parameters, kF ≈ 3 × 108 m−1). In agreement with our earlier studies27,60, we assume Z = 0.5—mimicking an average junction transparency of \(\overline{\tau }=1/[1+{(Z/2)}^{2}] \, \approx \, 0.94\)—and λSOI = 0.661—corresponding to Rashba SOI αR ≈ 15 meV nm. For a typical g-factor of ∣g*∣ ≈ 12, the Zeeman parameter λZ = −0.40 used for our theoretical calculations in the main text corresponds to the magnetic field By ≈ −100 mT.
Data availability
The data that support the findings of this study are available at the online depository EPUB of the University of Regensburg, with the identifier DOI: 10.5283/epub.57878.
Code availability
The computer codes that support the theoretical results, the plots within this paper and other findings of this study are available from the corresponding author upon request.
References
Winkler, R. Spin-orbit Coupling Effects in Two-Dimensional Electron and Hole Systems, Physics and Astronomy Online Library No. no. 191 (Springer, 2003).
Manchon, A., Koo, H. C., Nitta, J., Frolov, S. M. & Duine, R. A. New perspectives for Rashba spin–orbit coupling. Nat. Mater. 14, 871 (2015).
Smidman, M., Salamon, M. B., Yuan, H. Q. & Agterberg, D. F. Superconductivity and spin-orbit coupling in non-centrosymmetric materials: a review. Rep. Prog. Phys. 80, 036501 (2017).
Sato, M. & Ando, Y. Topological superconductors: a review. Rep. Prog. Phys. 80, 076501 (2017).
Yuan, N. F. Q. & Fu, L. Topological metals and finite-momentum superconductors. Proc. Natl Acad. Sci. USA 118, e2019063118 (2021).
Hart, S. et al. Controlled finite momentum pairing and spatially varying order parameter in proximitized HgTe quantum wells. Nat. Phys. 13, 87 (2017).
Chen, A. Q. et al. Finite momentum Cooper pairing in three-dimensional topological insulator Josephson junctions. Nat. Commun. 9, 3478 (2018).
Edelstein, V. M. The Ginzburg - Landau equation for superconductors of polar symmetry. J. Phys.: Condens. Matter 8, 339 (1996).
Fuchs, L. et al. Anisotropic vortex squeezing in synthetic rashba superconductors: a manifestation of lifshitz invariants. Phys. Rev. X 12, 041020 (2022).
Kochan, D., Costa, A., Zhumagulov, I., & Žutić, I. Phenomenological Theory of the Supercurrent Diode Effect: The Lifshitz Invariant. https://arxiv.org/abs/2303.11975 (2023).
Lu, J. M. et al. Evidence for two-dimensional Ising superconductivity in gated MoS2. Science 350, 1353 (2015).
Buzdin, A. Direct coupling between magnetism and superconducting current in the josephson φ0 junction. Phys. Rev. Lett. 101, 107005 (2008).
Reynoso, A. A., Usaj, G., Balseiro, C. A., Feinberg, D. & Avignon, M. Anomalous Josephson current in junctions with spin polarizing quantum point contacts. Phys. Rev. Lett. 101, 107001 (2008).
Yokoyama, T., Eto, M. & Nazarov, Y. V. Anomalous Josephson effect induced by spin-orbit interaction and Zeeman effect in semiconductor nanowires. Phys. Rev. B 89, 195407 (2014).
Szombati, D. B. et al. Josephson φ0-junction in nanowire quantum dots. Nat. Phys. 12, 568 (2016).
Strambini, E. et al. A Josephson phase battery. Nat. Nanotechnol. 15, 656 (2020).
Ando, F. et al. Observation of superconducting diode effect. Nature 584, 373 (2020).
Baumgartner, C. et al. Supercurrent rectification and magnetochiral effects in symmetric Josephson junctions. Nat. Nanotechnol. 17, 39 (2022).
Wu, H. et al. The field-free Josephson diode in a van der Waals heterostructure. Nature 604, 653 (2022).
Díez-Mérida, J. et al. Symmetry-broken josephson junctions and superconducting diodes in magic-angle twisted bilayer graphene. Nat. Commun. 14, 2396 (2023).
Lin, J.-X. et al. Zero-field superconducting diode effect in small-twist-angle trilayer graphene. Nat. Phys. 18, 1221 (2022).
Turini, B. et al. Josephson diode effect in high-mobility insb nanoflags. Nano Lett. 22, 8502 (2022).
Pal, B. et al. Josephson diode effect from Cooper pair momentum in a topological semimetal. Nat. Phys. 18, 1228–1233 (2022).
Jeon, K.-R. et al. Zero-field polarity-reversible Josephson supercurrent diodes enabled by a proximity-magnetized Pt barrier. Nat. Mater. 21, 1008 (2022).
Lotfizadeh, N. et al. Superconducting Diode Effect Sign Change in Epitaxial Al-InAs Josepshon Junctions https://doi.org/10.48550/ARXIV.2303.01902 (2023).
Mazur, G. P. et al. The Gate-tunable Josephson Diode https://arxiv.org/abs/2211.14283 (2022).
Costa, A. et al. Sign reversal of the Josephson inductance magnetochiral anisotropy and 0–π-like transitions in supercurrent diodes. Nat. Nanotechnol. 18, 1266–1272 (2023).
Bauriedl, L. et al. Supercurrent diode effect and magnetochiral anisotropy in few-layer NbSe2. Nat. Commun. 13, 4266 (2022).
Yun, J. et al. Magnetic proximity-induced superconducting diode effect and infinite magnetoresistance in a van der Waals heterostructure. Phys. Rev. Res. 5, L022064 (2023).
Banerjee, A. et al. Phase Asymmetry of Andreev Spectra From Cooper-Pair Momentum. https://doi.org/10.48550/ARXIV.2301.01881 (2023).
Sundaresh, A., Väyrynen, J. I., Lyanda-Geller, Y. & Rokhinson, L. P. Diamagnetic mechanism of critical current non-reciprocity in multilayered superconductors. Nat. Commun. 14, 1628 (2023).
Assouline, A. et al. Spin-Orbit induced phase-shift in Bi2Se3 Josephson junctions. Nat. Commun. 10, 126 (2019).
Mayer, W. et al. Gate controlled anomalous phase shift in Al/InAs Josephson junctions. Nat. Commun. 11, 212 (2020).
Dartiailh, M. C. et al. Phase signature of topological transition in Josephson junctions. Phys. Rev. Lett. 126, 036802 (2021).
Zhang, B. et al. Evidence of φ0-Josephson junction from skewed diffraction patterns in Sn-InSb nanowires. arXiv https://arxiv.org/abs/2212.00199 (2022a).
Haxell, D. Z. et al. Zeeman and orbital driven phase transitions in planar josephson junctions. arXiv https://arxiv.org/abs/2306.01514 (2023).
Krive, I. V., Gorelik, L. Y., Shekhter, R. I. & Jonson, M. Chiral symmetry breaking and the Josephson current in a ballistic superconductor-quantum wire-superconductor junction. Low Temp. Phys. 30, 398 (2004).
Barone, A. & Paterno, G. Physics and Applications of the Josephson Effect, 2nd edn. (John Wiley and Sons, Inc., 1982).
Legg, H. F., Laubscher, K., Loss, D., & Klinovaja, J. Parity protected superconducting diode effect in topological josephson junctions. arXiv https://arxiv.org/abs/2301.13740 (2023).
Pientka, F. et al. Topological superconductivity in a planar josephson junction. Phys. Rev. X 7, 021032 (2017).
Scharf, B., Pientka, F., Ren, H., Yacoby, A. & Hankiewicz, E. M. Tuning topological superconductivity in phase-controlled Josephson junctions with Rashba and Dresselhaus spin-orbit coupling. Phys. Rev. B 99, 214503 (2019).
Daido, A., Ikeda, Y. & Yanase, Y. Intrinsic superconducting diode effect. Phys. Rev. Lett. 128, 037001 (2022).
Yuan, N. F. Q. & Fu, L. Supercurrent diode effect and finite-momentum superconductors. Proc. Natl Acad. Sci. USA 119, e2119548119 (2022).
He, J. J., Tanaka, Y. & Nagaosa, N. A phenomenological theory of superconductor diodes. N. J. Phys. 24, 053014 (2022).
Ilić, S. & Bergeret, F. S. Theory of the supercurrent diode effect in rashba superconductors with arbitrary disorder. Phys. Rev. Lett. 128, 177001 (2022).
Hu, J.-X., Sun, Z.-T., Xie, Y.-M. & Law, K. T. Josephson diode effect induced by valley polarization in twisted bilayer graphene. Phys. Rev. Lett. 130, 266003 (2023).
Grein, R., Eschrig, M., Metalidis, G. & Schön, G. Spin-dependent cooper pair phase and pure spin supercurrents in strongly polarized ferromagnets. Phys. Rev. Lett. 102, 227005 (2009).
Zhang, Y., Gu, Y., Li, P., Hu, J. & Jiang, K. General theory of josephson diodes. Phys. Rev. X 12, 041013 (2022).
Scammell, H. D., Li, J. I. A. & Scheurer, M. S. Theory of zero-field superconducting diode effect in twisted trilayer graphene. 2D Materials 9, 025027 (2022).
Davydova, M., Prembabu, S. & Fu, L. Universal josephson diode effect. Sci. Adv. 8, eabo0309 (2022).
Fu, P.-H. et al. Field-effect josephson diode via asymmetric spin-momentum locking states. arXiv https://arxiv.org/abs/2212.01980 (2023).
Lu, B., Ikegaya, S., Burset, P., Tanaka, Y., and Nagaosa, N. Josephson diode effect on the surface of topological insulators. arXiv https://arxiv.org/abs/2211.10572 (2022)
Costa, A., Fabian, J., & Kochan, D. Microscopic study of the Josephson supercurrent diode effect in 2DEG-based Josephson junctions. arXiv https://arxiv.org/abs/2303.14823 (2023b)
Fornieri, A. et al. Evidence of topological superconductivity in planar josephson junctions. Nature 569, 89 (2019).
Banerjee, A. et al. Signatures of a topological phase transition in a planar Josephson junction. arXiv http://arxiv.org/abs/2201.03453 (2022).
Baumgartner, C. et al. Effect of Rashba and Dresselhaus spin–orbit coupling on supercurrent rectification and magnetochiral anisotropy of ballistic Josephson junctions. J. Phys.: Condens. Matter 34, 154005 (2022).
Nichele, F. et al. Relating andreev bound states and supercurrents in hybrid josephson junctions. Phys. Rev. Lett. 124, 226801 (2020).
Shabani, J. et al. Two-dimensional epitaxial superconductor-semiconductor heterostructures: a platform for topological superconducting networks. Phys. Rev. B 93, 155402 (2016).
Wickramasinghe, K. S. et al. Transport properties of near surface InAs two-dimensional heterostructures. Appl. Phys. Lett. 113, 262104 (2018).
Baumgartner, C. et al. Josephson inductance as a probe for highly ballistic semiconductor-superconductor weak links. Phys. Rev. Lett. 126, 037001 (2021).
Zhang, T. et al. Mobility exceeding 100 000 cm2/V s in modulation-doped shallow InAs quantum wells coupled to epitaxial aluminum. Phys. Rev. Mater. 7, 056201 (2023).
Tinkham, M. Introduction to Superconductivity: Second Edition (Dover Books on Physics) (Vol i), 2nd edn (Dover Publications, 2004)
Goswami, S. et al. Quantum interference in an interfacial superconductor. Nat. Nanotechnol. 11, 861 (2016).
McLaughlan, J. R., Llewellyn-Samuel, E. M. & Crampin, S. Spin-orbit splitting of image states. J. Phys.: Condens. Matter 16, 6841 (2004).
Gindikin, Y. & Sablikov, V. A. Image-potential-induced spin-orbit interaction in one-dimensional electron systems. Phys. Rev. B 95, 045138 (2017).
Nakazawa, T., Takagi, N., Kawai, M., Ishida, H. & Arafune, R. Rashba splitting in an image potential state investigated by circular dichroism two-photon photoemission spectroscopy. Phys. Rev. B 94, 115412 (2016).
Hess, R., Legg, H. F., Loss, D., & Klinovaja, J. Josephson transistor from the superconducting diode effect in ___domain wall and skyrmion magnetic racetracks. arXiv https://arxiv.org/abs/2308.04817 (2023)
Sinner, A. et al. Superconducting diode sensor. arXiv https://arxiv.org/abs/2306.11145 (2023).
Volkov, P. A. et al. Josephson diode effects in twisted nodal superconductors. arXiv https://arxiv.org/abs/2307.01261 (2023).
Virtanen, P. & Heikkilä, T. T. Nonreciprocal Josephson linear response. Phys. Rev. Lett. 132, 046002 (2024).
Leroux, C.et al. Nonreciprocal devices based on voltage-tunable junctions. arXiv https://doi.org/10.48550/ARXIV.2209.06194 (2022).
De Gennes, P. G. Superconductivity of Metals and Alloys (Addison Wesley, 1989)
Andreev, A. F. Electron spectrum of the intermediate state of superconductors. Zh. Eksp. Teor. Fiz. 49, 655 (1966).
Kulik, I. O. Macroscopic quantization and the proximity effect in S-N-S junctions. Zh. Eksp. Teor. Fiz. 57, 1745 (1969).
Acknowledgements
We thank Max Geier, Abhishek Banerjee, and Karsten Flensberg for fruitful discussions. Work at Regensburg University was funded by the by EU’s HORIZON-RIA Programme under Grant No. 101135240 (JOGATE), by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) through Project-ID 314695032—SFB 1277 (Subprojects B05, B07, and B08)—and Project-ID 454646522—Research grant “Spin and magnetic properties of superconducting tunnel junctions” (A.C. and J.F.). D.K. acknowledges partial support from the project IM-2021-26 (SUPERSPIN) funded by the Slovak Academy of Sciences via the programme IMPULZ 2021 and Grant DESCOM VEGA 2/0183/21.
Funding
Open Access funding enabled and organized by Projekt DEAL.
Author information
Authors and Affiliations
Contributions
S.R., T.A. and J.B. fabricated the devices and performed initial transport characterization of the hybrid superconductor/semiconductor wafer. S.R. and T.A. performed the measurements with the SQUID device. S.R., C.S. and N.P. conceived the experiment. A.C., D.K. and J.F. formulated the theoretical model. A.C. performed the numerical simulations of the temperature dependence of the SDE and φ0. N.P., S.R., C.S., A.C. and D.K. wrote the manuscript. T.L., S.G. and G.C.G. designed the heterostructure and conducted MBE growth. S.R., T.A. and N.P. analyzed the data. M.J.M. supervised research activities at Purdue.
Corresponding author
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks Alexandre Assouline, Lingfei Zhao and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Reinhardt, S., Ascherl, T., Costa, A. et al. Link between supercurrent diode and anomalous Josephson effect revealed by gate-controlled interferometry. Nat Commun 15, 4413 (2024). https://doi.org/10.1038/s41467-024-48741-z
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-024-48741-z