Introduction

The semiconductor diode is a fundamental component in modern electronics due to the non-reciprocal responses1. Analogous non-reciprocal charge transport in superconductors - namely superconducting diode effect (SDE) - has great potential for superconducting quantum electronics, since Josephson junctions (JJs) and superconducting quantum interference devices (SQUIDs) have been key components of superconducting quantum devices2,3. SDE in the JJs - namely Josephson diode effect (JDE) - and in intrinsic superconductors has been theoretically proposed in various systems with broken time-reversal and inversion symmetries4,5,6,7,8,9,10,11,12. Field-induced and field-free SDE/JDE has been experimentally observed in various superconductors13,14,15,16,17,18,19,20, supercurrent interferometers21,22,23,24, and other systems25,26,27,28,29. Interestingly, asymmetric supercurrent interferometers or SQUIDs with a non-sinusoidal current-phase relation (CPR) provide a good platform to realize JDE (Fig. 1a) with ultra-small magnetic fields and ultra-low power consumption21,22, both of which are crucial elements for applications at ultra-low temperatures. Moreover, it is quite feasible to promote their efficiency by varying the generic configuration9, which is compatible with state-of-the-art lithography technology, and integrate them into large-scale superconducting quantum circuits.

Fig. 1: Mechanisms for SDE/JDE.
figure 1

a JDE in the asymmetric SQUID formed by two asymmetric JJs. b An asymmetric SQUID formed by asymmetric edge states in Ta2Pd3Te5 JJ and its device configuration. The upper and lower edges are the palladium and tantalum-tellurium atomic chain, respectively. The inset shows the photomicrograph of device S2, and the white scale bar corresponds to 1 μm. cf Simulated CPR based on the minimal model. JDE only exists in (d) with different supercurrents, a nonzero magnetic field (ϕ/ϕ0 ≠ 0), and higher harmonics in the CPR (αn ≠ 0). The red and pink arrows represent the maximum(Ic+) and minimum(Ic) current in the CPR, respectively. g Band structure of monolayer Ta2Pd3Te5. Edge states (blue and red lines) exist near the Fermi level and are marked in the corresponding edges of (b). h \(\left\vert {I}_{b}\right\vert -\left\vert V\right\vert\) curves for positive and negative current sweep at Bz = 8.4 mT. i, j Alternating switching between the superconducting and normal states at Bz = 8.4 mT and 10 mK in device S2.

Essentially, the appearance of JDE is related to the difference between maximum(Ic+) and minimum(Ic) current in the CPR. The conventional CPR is an odd function of superconducting phase difference φ between superconducting leads, entailing a zero supercurrent at φ = 0, which shifts [meaning I(φ = 0) ≠ 0] when both time-reversal and chiral symmetries are broken21,30,31. To achieve JDE, other mechanisms to cause the CPR deviating from standard sinusoidal I(φ) = Icsin(φ) will also be introduced. The characteristics of CPR, which are usually inferred by the interference pattern of SQUIDs and the Shapiro steps due to AC Josephson effect, as well as the newly-discovered JDE with non-reciprocal critical current, have been widely harnessed to detect the anomalous superconductivity32,33, such as proximitized helical/chiral topological edge states. In turn, in a SQUID constructed by two asymmetric edge states of the topological insulator, JDE with an unconventional CPR could be easily realized with highly-tunable efficiency5 and relatively easy fabrication (Fig. 1b). The relatively small number of edge supercurrent channels, in comparison to the case of bulk states transport, may contribute to the small critical current and power consumption33,34.

More precisely, the interferometer formed by two asymmetric edge supercurrents (Fig. 1b) can be viewed as a SQUID with two different JJs and induces JDE with requirements similar to the asymmetric SQUID9 in Fig. 1a. According to Souto et al.9, a minimal model with n JJs concatenated in an interferometer array is applied to account for the origin of JDE, and the CPR only takes the first and second harmonic contribution into consideration. The current is written as \(I(\varphi,\phi /{\phi }_{0})=\mathop{\sum }_{1}^{n=2}{I}_{n}sin(\varphi+2\pi (n-1)\phi /{\phi }_{0})+{\alpha }_{n}{I}_{n}sin(2\varphi+ 4\pi (n- 1)\phi / {\phi }_{0})\) for two-JJs in parallel, where In (αnIn) is the amplitude of the first (second) harmonic content for the nth JJ, αn is the current coefficient of the higher harmonic, ϕ is the magnetic flux and ϕ0 = h/2e is the magnetic flux quantum. In Fig. 1c–f, the simulations of the CPR using the aforementioned formula serves the purpose of illustrating three key requirements that contribute to the difference between Ic+ (red arrow) and Ic (pink arrow). First, an external magnetic field (ϕ/ϕ0 ≠ 0) is necessary to break time-reversal symmetry and induce nonzero current at φ = 0 in the CPR (Fig. 1d–f). Second, the supercurrent of each edge, including all the harmonic contributions, should be different (see Fig. 1d, e and detailed simulations in Supplementary Fig. 2). This condition can be realized by two interfering edge states with different dispersion5. It can also be met, for example, by two superconducting quantum point contacts with different transmission coefficients9, which may be linked to various factors such as different edge states, disorder situations, among others. All of these can cause the different supercurrent of edges in our devices, which is actually quite unique in topological materials. Finally, at least one edge channel should be transmissive so that the CPR acquires a higher harmonic (αn ≠ 0)9,35 (see Fig. 1d, f).

Results

Interfering JDE

In this work, highly efficient JDE, residing in the tilting supercurrent pattern, is reported in a Ta2Pd3Te5 edge interferometer at very small Bz and is driven by low power consumption to achieve a highly stable rectification effect. It is also accompanied by fractional Shapiro steps, revealing the higher harmonic contributions of the non-negligible transmission channels. The choice of Ta2Pd3Te5 is based on the following reasons: (1) Ta2Pd3Te5 is a van der Waals material with quasi-one-dimensional (1D) chains36 and can be easily mechanically exfoliated. (2) As a 2D topological insulator, where multiple layers could be viewed as simple stacking of monolayers due to very weak inter-layer coupling37,38, it hosts excitonic insulator states39,40,41 and edge states observed by scanning tunneling microscopy40,42 and transport measurements38,43. Notably, owing to the anisotropic bonding energy of 1D chains, the edges possess different atomic chains (Pd and Ta-Te atomic chains), resulting in asymmetric dispersion, as shown by the red and blue lines/arrows in Fig. 1b–g. (3) Superconductivity can be easily achieved in this system through doping or high-pressure techniques44,45,46. Therefore, by proximity effect, the induced supercurrent (Supplementary Fig. 1) carried by edge states of Ta2Pd3Te5 can be used to realize an asymmetric SQUID with the interfering JDE.

In Fig. 1h, JDE reaches up to 73% efficiency [\(\eta=2({I}_{c+}-\left\vert {I}_{c-}\right\vert )/({I}_{c+}+\left\vert {I}_{c-}\right\vert )\)] at an out-of-plane magnetic field Bz = 8.4 mT in \(\left\vert {I}_{b}\right\vert -\left\vert V\right\vert\) traces. The measurement is performed in the following sequence: (1) Zero-to-positive current sweep (0-p, red scatters); (2) Positive-to-zero current sweep (p-0, blue line); (3) Zero-to-negative current sweep (0-n, magenta scatters); (4) Negative-to-zero current sweep (n-0, cyan line). The retrapping current Ir+ (Ir) and switching current Ic (Ic+) are defined during the positive-to-negative (negative-to-positive) current sweep. Ir+ (\(\left\vert {I}_{r-}\right\vert\)) is nearly the same as Ic+ (\(\left\vert {I}_{c-}\right\vert\)), suggesting negligible capacitance in the JJ6,11,25. Unlike reciprocal transport (\({I}_{c+}=\left\vert {I}_{c-}\right\vert\) and \({I}_{r+}=\left\vert {I}_{r-}\right\vert\)), the difference between Ic+ (Ir+) and \(\left\vert {I}_{c-}\right\vert\) (\(\left\vert {I}_{r-}\right\vert\)) indicates the presence of JDE in this JJ.

The superconducting half-wave rectification is observed at Bz = 8.4 mT and 10 mK in Fig. 1i, j. The ‘on’/‘off’ (superconducting/normal) states are switched by alternating current bias (Ib = ±400 nA). Many cycles of these alternating measurements are conducted over a measurement time exceeding 1.5 h, demonstrating the high stability of this rectification device, which is crucial for its potential applications. Intriguingly, the estimated switching power (\({I}_{b}^{2}{R}_{N}\)) reaches the picowatt level (6.4 pW for device S2 with RN = 40 Ω, 0.56 pW for device S1 in Supplementary Fig. 1), which is four and eight orders of magnitude smaller than the field-free Josephson diode25 and another bulk superconducting diode13, respectively. This power is also close to or even lower than the power of nanowire/nanoflake SDE systems19,21,47,48 and thin film SDE systems15,18,49. The ultra-low switching power in Ta2Pd3Te5 Josephson diodes may be attributed to fewer edge supercurrent channels, making it a potential candidate for applications. Nonetheless, further technological advancements are required to optimize the efficiency, magnetic field conditions, and power consumption not only in this Josephson diode but also in other superconducting diodes.

To further study the origin of JDE, we measure the magnetic flux dependence of the JDE in the Ta2Pd3Te5 JJ. In Fig. 2a, dV/dI as a function of Bz and Ib displays a SQUID pattern for device S1, suggesting the supercurrent interference of two channels in this JJ. The sawtooth-shaped Ic pattern shows up, usually indicating the involvement of higher harmonic components due to non-negligible transmission channels50, consistent with findings from Supplementary Note 1. The position-dependent supercurrent density distribution (Fig. 2b) extracted from the SQUID pattern33 and the estimated enclosed area formed by two interference supercurrents (see Supplementary Note 3 and 9) illustrate the supercurrent mainly originates from edge states of the JJ, confirming the reported edge state in Ta2Pd3Te538,40,42. In Fig. 2c, the Bz-dependent Ic+ and \(\left\vert {I}_{c-}\right\vert\) extracted from Fig. 2a shows the asymmetric Josephson effect. Several features, such as the large deviation of Ic from zero at half flux quantum in Fig. 2c and JDE in Fig. 2d, indicate the asymmetric edge supercurrent channels in this JJ, as detailed in “Methods” section. Moreover, \(\Delta {I}_{c}={I}_{c+}-\left\vert {I}_{c-}\right\vert\) oscillates with Bz (orange curve in Fig. 2d), which can be called interfering JDE and is approximately described by the two-JJs model (black line in Fig. 2d, and see the detail in “Methods” section). This type of JDE has been theoretically predicted in Weyl semimetals with broken inversion symmetry and asymmetric helical edge states5. Furthermore, the oscillating Josephson diode efficiency (cyan line in Fig. 2d) shows a maximal efficiency η≈45% at 6.4 mT. It is intriguing that η can reach 10% at Bz = −0.5 mT. Also, the required field can be significantly reduced when the enclosed area of the SQUID loop is increased, which could be easily realized in the experimental setup21,22.

Fig. 2: Asymmetric Josephson effect and JDE in Ta2Pd3Te5 JJs.
figure 2

a SQUID pattern of device S1 at 15 mK. b Position-dependent supercurrent density distribution. Inset: A sketch of edge supercurrents-formed JJ. c Bz-dependent Ic+ and \(\left\vert {I}_{c-}\right\vert\). The switching current Ic+ and Ic are extracted by  ~15% of normal resistance. d Non-reciprocal critical current \(\Delta {I}_{c}={I}_{c+}-\left\vert {I}_{c-}\right\vert\), Josephson diode efficiency η and fitted ΔIc (black line) as a function of Bz. e Temperature-dependent Ic+, Ic, ΔIc and η for device S1 at Bz = 6.4 mT. The error bars primarily stem from the definition of the switching current, which is determined by 15% ± 5% drop of the normal resistance. f SQUID pattern of device S2 at 10 mK. g Supercurrent density distribution. The inset is a draft of the JJ with three-edge supercurrents. h Bz dependence of Ic+ and \(\left\vert {I}_{c-}\right\vert\). i Oscillating ΔIc, η and fitted ΔIc (black line). j Temperature-dependent Ic+, Ic, ΔIc and η for device S2 at Bz = 8.4 mT. The error bars are obtained by the same method as in (e).

Having demonstrated the JDE in Ta2Pd3Te5 edge interferometer, we now show how to further enhance its efficiency. Interestingly, compared with device S1, the SQUID pattern of device S2 in Fig. 2f reveals two main periods, leading to a more pronounced asymmetry in the pattern (Fig. 2h, i) and an increased efficiency with ηmax = 73% at 8.4 mT. The enhanced JDE can be explained by three-JJs in parallel. The formation of the middle edge state in the JJ may stem from the ladder-like structure induced by imperfect exfoliation (as indicated by the green arrow in the inset of Fig. 2g and supplementary Fig. 3d), which is a common occurrence during the exfoliation process of 2D flakes51. These three supercurrent channels form two sets of interference patterns (corresponding red and green dashed loops in the inset of Fig. 2g), leading to the asymmetric SQUID. The ΔIc can also be approximately fitted by the three-JJs model (black line in Fig. 2i, and see the detail in “Methods” section). The optimization of η in device S2 (Fig. 2j), compared to device S1 in Fig. 2e, seems to be achieved by increasing the number of JJs in parallel, which is in line with the theoretically predicted method9.

Here, we examine the main mechanisms to determine which one best fits our data. (1) Magnetism mechanism20 is initially excluded due to the absence of magnetism in the Ta2Pd3Te5 JJ. (2) The Rashba-spin orbital coupling (SOC) mechanism is excluded as it typically requires in-plane magnetic fields, and Ising-SOC is also considered implausible because it is usually found in 2D superconductors52. (3) JDE/SDE caused by finite-momentum Cooper pairing generally also needs in-plane magnetic fields8,14. The absence of obvious enhancement of the upper critical field under in-plane magnetic fields or the approximate butterfly interference pattern in our JJs14 (Supplementary Fig. 5) leads to the exclusion of finite-momentum Cooper pairing as a mechanism. (4) Nonlinear capacitance6,28 is absent in this system due to the symmetric superconducting electrodes and the lack of obvious hysteresis between switching current and retrapping current (Fig. 1h). (5) Self-inductance is also found to be too small to be considered (see Supplementary Note 8). Therefore, the main origin of Bz-induced JDE in the Ta2Pd3Te5 edge interferometer is attributed to asymmetric edge supercurrents, as detailed in the “Methods” section.

Fractional Shapiro steps

As discussed above, besides asymmetric edge supercurrents and time-reversal symmetry breaking, the remaining higher harmonic in the CPR is further studied to explain the JDE in our JJs. The Shapiro resonances have been widely used to reveal not only the exotic symmetry of Cooper pairing32 but also the higher harmonic or non-sinusoidal CPR9,53. The Shapiro steps appear in the IV curve at V = Vn ≡ nhf/2e (n is an integer) when the JJ is irradiated with the microwave. If the nth harmonic contributes to the CPR, m/nth Shapiro steps (m is an integer) will be present9,53. In Fig. 3a, fractional Shapiro steps emerge at a microwave frequency of f = 5 GHz, indicating the significant transparency of the JJ (Supplementary Note 1). The  ±1/2th Shapiro steps at half flux quantum (n + 1/2)ϕ0 are marked by black arrows, which is also observed in other works24,54. The characteristic dV/dI valleys formed by half-integer steps can be clearly observed at marked positions in Fig. 3b at Bz = 5.5 mT (corresponding to 3/2 ϕ0). In addition, half-integer Shapiro steps are also observed in device S2 with flux-dependent measurements (f = 4.64 GHz in Fig. 3c, d or 5.02 GHz in Supplementary Fig. 7) and power-dependent measurements at 3ϕ0/2 (Fig. 3e). These experiments indicate the existence of the second harmonic in the CPR, which is also supported by the rapid suppression observed in temperature-dependent ΔIc compared with Ic in device S1 (see Fig. 2e), because higher harmonics are quickly suppressed when temperature approaches Tc55. The rapid suppression of ΔIc in device S2 (Fig. 2j) seems less pronounced, possibly because its three-JJs model enhances the diode effect, therefore relatively weakening such suppression. In addition to their applications in Josephson diodes, the JJs with higher harmonics also have potential applications in 0 - π qubits, particularly if the first harmonic in the CPR can be effectively eliminated56,57,58.

Fig. 3: Fractional Shapiro steps under microwave in both Ta2Pd3Te5 devices.
figure 3

a Differential resistance (dV/dI) as a function of voltage characteristics and flux quantum at microwave power = −32 dBm, f = 5 GHz, and 30 mK for device S1. The  ±1/2th Shapiro steps are marked by black arrows. b dV/dI versus Ib at Bz = 5.5 mT (3/2 ϕ0). Valleys characterized by half-integer steps are observed clearly. c Flux quantum and voltage dependence of dV/dI at microwave power = 10 dBm, f = 4.64 GHz, and 10 mK for device S2. d Valleys characterized by the  −1/2th Shapiro step at 6.7 mT (3/2 ϕ0). e Microwave power dependence of dV/dI at 4.64 GHz and Bz = 6.7 mT in device S2. The  −1/2th Shapiro step is marked by black arrows. f Microwave power-dependent non-reciprocal critical current of the 0th Shapiro steps, which is extracted from (e). The switching currents here can be determined by peaks in the dV/dI - Ib curve, as shown in Supplementary Fig. 8c, and the error bars primarily originate from the broadening of these peaks.

Furthermore, some features in Fig. 3 require further explanation. First, the disappearance of half-integer Shapiro steps at zero field (Fig. 3a, c) may be attributed to the relatively small contribution of the second harmonic in the CPR and the low microwave frequency59. In device S1, a higher microwave frequency is utilized at zero field, and a small signal of half-integer steps appears in Supplementary Fig. 6d–f. No higher frequency is applied to detect half-integer steps at 0 T, considering the limitations of our microwave generator. Second, the integer Shapiro steps at 5.5 mT (3/2 ϕ0) appear weaker in Fig. 3b (or Supplementary Fig. 6c) compared to those at 0 T in Supplementary Fig. 6b. This is due to the application of a magnetic field at half the flux quantum, causing a reduction in the sharpness of the integer steps because of a decrease in the first harmonic of the CPR. Subsequently, the second harmonic or fractional Shapiro steps become visible at half the flux quantum (Fig. 3a), in agreement with theoretical predictions9,35 and experimental results24,54. Third, in device S2 (Fig. 3c), half-integer steps vanish at  ±ϕ0/2 due to the superposition of multiple SQUID patterns, which is slightly different from the single SQUID pattern observed in device S1. In the case of multiple SQUID patterns, the position of destructive interference for the first harmonic differs from that of a single SQUID pattern, leading to a notable contribution from the first harmonic and the absence of half-integer steps at  ±ϕ0/2 (see Supplementary Note 6).

The JDE is also studied as a function of microwave power. In Fig. 3f, switching current \({I}_{c+}^{0th}\) and \({I}_{c-}^{0th}\) of the 0th Shapiro step are extracted from Fig. 3e. The quantity \(\Delta {I}_{c}^{0th}={I}_{c+}^{0th}-\left\vert {I}_{c-}^{0th}\right\vert\) (orange curve, see the detail in Supplementary Fig. 8) displays little variation and gradually decreases with increasing microwave power. This stable observation of the JDE under microwave irradiation represents one of the initial examples of such behavior and is a crucial feature in quantum information applications, where microwaves are commonly utilized2.

Antisymmetric second harmonic transport

As another type of non-reciprocal behavior, the second harmonic resistance (R2w) is usually measured to characterize magnetochiral anisotropy (MCA)13,25,60, whose strength is often defined by the coefficient \(\gamma=\left\vert \frac{2{R}_{2w}}{{R}_{w}BI}\right\vert\). MCA typically exhibits antisymmetric behavior in R2w with respect to magnetic fields and is generally used to characterize noncentrosymmetric or chiral features. In normal conductors, MCA is usually a minor effect, generally due to the weak SOC. However, a significant enhancement of MCA is observed in noncentrosymmetric superconductors, potentially offering additional insights into factors such as disparities between the Fermi energy and superconducting gap, the Rashba parameter, and spin-orbit splitting, among other factors60,61. Moreover, the antisymmetric oscillations in B-dependent R2w have been observed in an ionic liquid-gated chiral nanotube exhibiting the Little-Parks effect62, indicating its characteristic chiral symmetry.

Therefore, the second harmonic resistance in our supercurrent interferometer is measured to clarify its asymmetric features. In Fig. 4a, b, the first harmonic resistance (Rw) and R2w of device S2 are measured using lock-in techniques. Below Iac = 100 nA, a typical pair of antisymmetric peaks in R2w is observed at critical fields (near the green dashed line in Fig. 4b). This is consistent with the previously reported antisymmetric peaks of R2w with respect to B in noncentrosymmetric superconductors13,60, suggesting the existence of asymmetry in our JJ. With increasing Iac, the antisymmetric oscillations in R2w become more pronounced due to the quantum interference of asymmetric edge supercurrents, similar to the behavior observed in chiral nanotubes62. At around 500 nA, the oscillating R2w (black line in Fig. 4b) is analogous to the ΔIc (orange line), and the similar behavior of device S1 can be seen in Fig. 4d (see the detail in Supplementary Note 7). This similarity suggests the antisymmetric oscillations in R2w may be related to the polarity of JDE. At higher Iac, the superconductivity is substantially suppressed and self-heating effects may begin to influence the second harmonic resistance, which is not further analyzed.

Fig. 4: Antisymmetric second harmonic transport of Ta2Pd3Te5 JJs.
figure 4

a Rw versus Bz at different currents. b Bz-dependent R2w at various currents. Typical pairs of antisymmetric peaks (R2w) are marked by dashed lines in matching colors. R2w at 500 nA (black line) shows a similar behavior to ΔIc (orange line). The green dashed line indicates the regions near the critical fields, while the other dashed line marks the areas near the oscillating peaks or valleys. c Typical \(\left\vert {R}_{2w}\right\vert\) as a function of current at 20 mK for different oscillating peaks. The average \(\left\vert {R}_{2w}\right\vert\) values are obtained using pairs of antisymmetric peaks near the dashed lines in (b). The error bars primarily originate from discrepancies between average \(\left\vert {R}_{2w}\right\vert\) and the observed values. d Antisymmetric R2w and ΔIc as a function of Bz in device S1. The R2w is measured at Iac = 70 nA.

In addition, supercurrent interference-induced R2w in our JJs appears more complex. In Fig. 4c, the current-dependent R2w (red, magenta, and violet lines) extracted from the same oscillating peaks below critical fields does not intersect the origin, suggesting deviations from the R2w = \(\frac{1}{2}\gamma {R}_{w}IB\) formula generally observed in other systems63,64,65,66. Therefore, although the antisymmetric oscillations in R2w at large Iac display partial antisymmetric features of MCA, the 2R2w/RwIB calculated using the typical formula below the critical field may not accurately describe γ or MCA. The MCA in interference systems appears more complex and requires further investigation.

Discussion

The Ta2Pd3Te5 edge interferometer reveals a significant JDE under small out-of-plane magnetic fields. The JDE is accompanied by the enhanced antisymmetric second harmonic transport, which deserves further research. Moreover, the presence of asymmetric Josephson critical current alongside fractional Shapiro steps demonstrates the importance of the higher harmonic content in the JDE. It is noted that power consumption requires attention if SDE/JDE is to be applied at low temperatures in the future, and further optimization is necessary not only regarding mechanism but also in terms of device size. The Josephson diode efficiency can be further enhanced by concatenating interferometer loops. These findings offer a promising approach to exploring JDE with significant diode efficiency, ultra-low switching power at small magnetic fields, and stable JDE under microwave irradiation, rendering them potentially valuable for practical applications.

Methods

Device fabrication

Single crystals Ta2Pd3Te5 were prepared by the self-flux method36. Ta2Pd3Te5 thin films were obtained through mechanical exfoliating bulk samples onto Si substrates with the 280-nm-thick SiO2 on top, and then coated with PMMA in a glove box at 4000 rpm for 60 s, followed by annealing at 100 C for 120 s. Multi-terminal electrical contacts were patterned after electron-beam exposure and subsequent development. Ti/Al (5 nm/60 nm) electrodes were deposited after Ar etching in order to remove the oxidized layer. After the lift-off step, the device was coated with hexagonal boron nitride (hBN) to prevent oxidation. The entire fabrication process took place in a nitrogen atmosphere glove box. The rectangular film, with a long side along the b-axis of the crystal, was easily obtained due to its quasi-1D nature38. The dimensions of device S2 (device S1) are as follows: thickness 37 nm (16 nm), width 1.9 μm (1.3 μm), and length 245 nm (400 nm), respectively.

Transport measurements

The electrical transport measurements were carried out in cryostats (Oxford instruments dilution refrigerator). The DC current bias measurements were applied using a Keithley 2612 current source. Both the first- (Rw) and second harmonic (R2w) resistance were measured through a standard low-frequency (7–11 Hz) lock-in technique (LI5640, NF Corporation). The phase of the first- and second harmonic signal was set to be 0° and 90°, respectively. The current of the magnet was applied by a Keithley 2400 current source in order to control the magnetic field accurately. Before the measurement, the annealing process with 300 °C for  ~10 min was performed to obtain good contacts in the Ta2Pd3Te5 JJs. IV curves were obtained through the numerical integration process and the origin data were all dV/dI − Ib curves.

Transport measurements under microwave radiation

The electrical transport measurements under microwave radiation were carried out using PSG-A Series Signal generator (Agilent E8254A). The main microwave frequency between 4.64 and 5.02 GHz was applied due to the large absorption effect of microwave for devices S1 and S2. The dissipation of the microwave circuit existed during the measurement of device S2, while another microwave circuit for device S1 was good. So the applied power of the microwave was a little large but without inducing obvious thermal effect.

Band structure calculations

The first-principles calculations were carried out based on the density functional theory (DFT) with the projector augmented wave (PAW) method, as implemented in the Vienna ab initio simulation package (VASP)37. The generalized gradient approximation (GGA) in the form of the Pardew-Burke-Ernzerhof (PBE) function was employed for the exchange-correlation potential. The kinetic energy cutoff for plane wave expansion was set to 400 eV, and a 1 × 9 × 1 k-mesh was adopted for the Brillouin zone sampling in the self-consistent process. The thicknesses of the vacuum layer along the x and z axis were set to  >20 Å.

Simulations of JDE

A minimal model with n JJs in parallel is used to explain the JDE9, where only the first and second harmonic contributions in the CPR are considered. The supercurrent is approximatively described as \(I(\varphi,\phi /{\phi }_{0})=\mathop{\sum }_{1}^{n}{I}_{n}sin(\varphi+2\pi {\zeta }_{n}\phi /{\phi }_{0})+ {\alpha }_{n}{I}_{n}sin(2\varphi+ 4\pi {\zeta }_{n}\phi / {\phi }_{0})\), where ζn is the modified ratio and ζ1 = 0, ζ2 = 1 (ζ1 = 0, ζ2 = 1, ζ3 = 0.23) for the simulation of our two (three) JJs in parallel. ζ3 represents the enclosed area ratio of two corresponding SQUID patterns in device S2. Preliminary switching currents for different channels are estimated in both devices without considering the second harmonic contribution, in order to simulate CPR and JDE. For device S1, the parameters used for simulating ΔIc are: I1 = 75 nA, I2 = 15 nA, α1 = 0.3 and α2 = 0.5, while for device S2, the parameters are: I1 = 330 nA, I2 = 120 nA, I3 = 158 nA, α1 = 0.6, α2 = 0.1 and α3 = 0.2. The simulated ϕ/ϕ0-dependent ΔIc can approximatively explain the experimental data in both devices, as shown by black lines in Fig. 2d, i. The slight discrepancy between the simulation and experiment may arise from the omission of higher-order harmonics in the CPR, a reduction in critical current, and a potential variation of α with increasing magnetic field, or other contributing factors.

Asymmetric edge supercurrents

We will discuss edge supercurrents, their asymmetry, and the origin of such asymmetry in our cases, using device S1 as an example. Firstly, the SQUID-like pattern observed in Fig. 2a exhibits several interference features of two edge supercurrent channels, as supported by previous reports38,40,42. In this pattern, both the width and amplitude of the center lobe are approximately equal to those of the other lobes. This is obviously contrast to a typical Fraunhofer pattern induced by the bulk transport33,34, where the width of the center lobe is twice as large as that of the other lobes, while the amplitude of the center lobe is usually significantly larger than that of the other lobes.

Secondly, there are several features supporting asymmetric edge supercurrents in our devices: (1) Without considering the higher harmonics of CPR, we calculated two edge supercurrent channels (I1 and I2) from the SQUID pattern in Supplementary Note 3, and in device S1, the ratio between the supercurrents I1:I2 is approximately 3.5:1 (I1:I2:I3~6:1:2 for device S2); (2) After considering the second harmonic, from a simple simulation using a two-JJs model in Supplementary Fig. 2e, the deviation of Ic from zero at half flux quantum alone cannot support asymmetric interference channels. However, the coexistence of this deviation and the JDE can support asymmetric supercurrents, as shown in Supplementary Fig. 2d, g, h, and device S1 satisfies this coexistence condition; (3) The position-dependent supercurrent density distribution shows a significantly asymmetric supercurrent distribution. However, it should be noted that such asymmetric supercurrent density distribution can also exist in a symmetric system33,34 due to the relatively imperfect experimental data.

Finally, the origin of asymmetric edge supercurrents is further discussed: (1) The DFT calculations in Fig. 1g support the existence of asymmetric edge states, which contribute to asymmetric edge supercurrents and JDE5; (2) Different transmission coefficients9 of two edge JJs may also lead to asymmetric edge supercurrents, and can be caused by the different edge disorders, different coupling with the Al contacts, and other factors. Such conditions were not intentionally controlled in our devices.

Discussion of switching power

From a practical perspective, the power consumption of devices should be carefully considered at low temperatures. Although the current cooling power of commercial dilution refrigerators generally reaches the microwatt level at 10 mK67, a Joule excitation power higher than a picowatt (or nanowatt) may raise the temperature of the device or sensor above 10 mK (or 100 mK) when the mixing chamber is kept at 10 mK68,69. The self-heating effect is sometimes noticeable and can also be applied. For example, a microheater with a power output of a few picowatts can generate a temperature gradient as large as several millikelvins per micron and can be further used to conduct Seebeck or Nernst effect measurements in 2D devices at millikelvin base temperatures70. Therefore, ultra-low power consumption is essential not only for basic measurements but also for applications in large-scale superconducting quantum circuits, which need ultra-low temperatures to enhance the precision of quantum state measurements.

The edge-induced JDE in Ta2Pd3Te5 may offer distinct advantages in terms of power consumption. The edges of the topological system exhibit a low number of conducting channels compared to bulk states transport. This characteristic can result in a lower critical current and reduced power consumption33,34.