Josephson Effect in NbS2 van der Waals Junctions
Abstract
Van der Waals (vdW) Josephson junctions can possibly accelerate the development of advanced superconducting device that utilizes the unique properties of two-dimensional (2D) transition metal dichalcogenide (TMD) superconductors such as spin-orbit coupling, spin-valley locking. Here, we fabricate vertically stacked NbS2/NbS2 Josephson junctions using a modified all-dry transfer technique and characterize the device performance via systematic low-temperature transport measurements. The experimental results show that the superconducting transition temperature of the NbS2/NbS2 Josephson junction is 5.84 K, and the critical current density reaches 3975 A/cm2 at 2K. Moreover, we extract a superconducting energy gap meV, which is considerably smaller than that expected from the single band s-wave Bardeen-Cooper-Schrieffer (BCS) model ( meV).
keywords:
Josephson junction, two-dimensional materials, Van der Waal heterostructures, NbS2, superconducting energy gap0.1 TOC GRAPHIC
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/35aa00ba-9d31-4801-b081-cae7bb3bf9a0/x1.png)
The Josephson junction is one of the basis of quantum metrology, quantum communication, and quantum computing. The Josephson junctions consisted of superconductor/oxide insulator/superconductor, such as Al/Al2O3/Al, are the mainstream scheme which leads to some of the most celebrated breakthrough in the fields such as the demonstration of quantum supremacy1. However, the in-homogeneity of the insulating layer and defect residing within the layer plague the device performance2, 3. For instance, these imperfections can result in noticeable decoherence in a superconducting qubit and Josephson microwave source4, 5. It is therefore necessary to explore alternative platforms for realizing Josephson junction which can possibly overcome the aforementioned shortcomings. In this regard, the layered transition metal dichalcogenides (TMD), such as NbSe2, has been one of the leading platforms, and stimulates the developments of Josephson junctions consisting of two superconducting TMD flakes, where misalignment and misorientation cause the difference in the order parameter for the two flakes and give rise to the Josephson relation (the conceptual illustration of vdW Josephson junction is shown in Figure S1)6, 7.
However, a comprehensive understanding of TMD-based Josephson junction is still lacking. One of the reasons is that for the superconducting TMDs, their superconductivity coexists with charge-density wave (CDW) states8. For instance, O. Naaman et al. found that in Pb/I/NbSe2 junction, the superconducting energy gap of NbSe2 is smaller than expected, which could be attributed to the interplay between the superconductivity and the coexisting CDW in this material9. This can potentially complicate the implementation of all TMD-based superconducting circuit. Fortunately, 2H-NbS2 shares the same crystal structure with NbSe2 and it is the only superconducting TMD material with no CDW order observed experimentally10. Understanding the behavior of NbS2 Josephson junction may provide a universal insight to the key properties of other TMD-based Josephson junctions. However, high-quality NbS2 Josephson junctions are less studied11. The reasons are twofold: First, the exfoliated NbS2 thin flakes are not stable and suffers rapid oxidization in ambient12. Second, the interlayer contaminants within vdW junctions may lead to the degradation of device quality13.
In this letter, we report the observation of Josephson effect in high quality NbS2/NbS2 vdW junctions for the first time. The junction is obtained by a modified all-dry transfer technique under inert environmental conditions. The critical current density is as high as 3975 A/cm2 at 2 K, which may meet the need for applications including high output on-chip microwave source5, high gain Josephson amplifier14. The extracted superconducting energy gap is smaller than the value predicted by the BCS theory. Our results call for more detailed investigation on the superconducting mechanism in NbS2 and other TMD materials in general, which will promote the development of TMD superconducting circuits.

High-quality 2H-NbS2 single crystals were grown by chemical vapor transport using iodine as a transport agent and had a typical Tc of 6.13 K15. Figure 1(a) shows the top and side view of the layered structure of 2H-NbS2, respectively. Vertical NbS2/NbS2 junction devices are fabricated by a modified dry-transfer technique in a Nitrogen-filled glove box to avoid contamination and oxidization. Water and oxygen are evacuated below 1 ppm for the most optimized device quality. The detailed device fabrication processes are illustrated in Figure 1(b). First, the bulk crystals are mechanically exfoliated and placed on top of silicone elastomer polydimethylsiloxane (PDMS) substrates using adhesive tape. It is necessary to emphasize that the PDMS substrates are cleaned with isopropanol alcohol (IPA) before transferring the NbS2 flakes onto it, ensuring a clean interface13. Then flakes of uniform surface and desired geometry are chosen and sequentially stacked onto a SiO2/Si substrate with pre-patterned Au/Ti electrodes. To remove any resist residues in the contact areas before transfer, the wafers were cleaned with acetone, IPA, deionized water and oxygen plasma at 300 W for 5 min. To avoid the effect of heating on the surface degradation of the NbS2 flakes, the entire device fabrication process was performed at room temperature. The thickness of all the NbS2 flakes used for the devices in this work are between 20 and 60 nm, to ensure negligible suppression Tc (Figure S2), because the Tc of NbS2 is known to decrease in flakes <10 nm12. Figure 1(c) shows a transmission electron microscope (TEM) image of the vertical cross-section of the junction, an artificial interface (equivalent to the insulating layer) with a thickness of 3 nm resides between the top and bottom NbS2 flakes. Some non-pristine interface layers and artificial interface are likely to be NbOx, which may be caused by the oxidization processing during taking the TEM images. To obtain the TEM images, the device was cleaved by focused ion beam and exposed in the ambient environment, and the oxidization came into play. The air-sensitivity of NbS2, especially compared to NbSe2, renders the difficulty in obtaining a high-quality junction. However, we would like to stress that the oxidized interface does not necessarily occur in the measured sample where it is kept in high vacuum condition. Figure 1(d) is a schematic illustration of the junction structure with a four-terminal measurement configuration and the typical junction areas span from 7.8 to 50 .

We first measured the temperature dependence of four-terminal resistance R(T) of a Josephson junction (JJ1) in the presence of a 10 d.c. excitation current at zero magnetic field and observed the superconducting transition in Figure 2(a). The critical temperature Tc of this device is determined to be 5.84 K. Note that Tc is defined as the temperature where resistance drops to 10 of the normal-state resistance. The transition window is as narrow as 0.2 K, without any additional kinks as shown in the inset of Figure 2(a). The clean transition highlights the cleanness of the junction area thanks to our specifically designed transfer protocol.
Next, the current-voltage (I-V) curve of JJ1 measured at 2 K is plotted in Figure 2(b), which displays a pronounced hysteretic feature, a signature of an under-damped Josephson junction. We checked that similar hysteresis behavior can be obtained at different current sweeping rates, which rules out the possibility that the observed hysteresis arises from Joule heating. This suggests that the vdW interface decouples the superconducting order parameters and creates a weak link that allows the flow of a Josephson supercurrent. We observed the same hysteretic feature in several other Josephson junctions (Figure S3). Besides, the switching current Is is about 1.26 mA in Figure 2(b), which corresponds to a current density of 3975 A/cm2. The large switching current and current density are indication of strong Josephson coupling, which would be beneficial for applications including superconducting quantum qubit16 and Josephson microwave source5. The Stewart-McCumber parameter17 of the device is estimated to be (where mA is the re-trapping current) and the junction capacitance calculated by equation (where =2.90 is normal state resistance of the Josephson junction extracted from I-V curve) is 0.23 . The characteristic parameters obtained from other junctions are presented in Table S1. It is interesting to point out that both the switching current and re-trapping current can be asymmetric, and the asymmetry lager than the fluctuation of Is (the results of other junctions are shown in Figure S3 and the switching current statistics of JJ3 can be found in Figure S4). This observation might be caused by the superconducting diode effect (SDE)18, 19 or the self-field effect arising from the superconducting electrodes20, 21, 22. However, for our NbS2 Josephson junctions, the asymmetry is insensitive to the out-of-plane magnetic field, which contrary to the predicted behavior of SDE arising from Ising spin-orbital coupling (the dominate type spin-orbital coupling in NbS2). As a result, we believe that the nonreciprocal supercurrent is not caused by SDE. Instead, we attribute the observation to the self-field effect. Interestingly the asymmetric switching current has not been reported in the more intensively studied NbSe26.

We further examined the field effect of switching current for our Josephson junctions by applying an in-plane magnetic field. Figure 3 shows the contour plot of voltage as a function of bias current and external parallel field measured at a temperature of 1.8 K in JJ4 when current is swept from 0 to 150 . A periodic modulation resembling the expected Fraunhofer pattern is observed. The minima of Is(B) do not reach zero, which may be due to the flux focusing23, 24, 25, 26 and the uneven interface in the vicinity of the electrodes (see Figure S5 of the Supporting information for the cross-sectional TEM image), like the results reported in MgB2 break junctions27 and high-Tc superconductor grain boundary junction.24, 28 The asymmetric data is most likely caused by the bias current induced magnetic field.29, 22 Additionally, the London penetration depth () of NbS2 can be obtained from , where is the magnetic flux quantum, is the period of the oscillation in the Fraunhofer pattern, is width of the junction perpendicular to the field direction, and d is the interface thickness. From the Fraunhofer pattern, the modulation period is averaged to be around 45 mT after ignoring some small oscillations and with , and , which corresponds to a is calculated to be 9 nm, an order smaller than in NbS2 bulk30. We did not observe a well-defined periodic modulation in junctions with larger overlap area, such as JJ1, mainly due to the in-homogeneity in the junction interface.

The superconducting mechanism in TMD material is under debate, here we provide further insight to this topic by measuring the superconducting energy gap of NbS2. We measured the I-V curves of the JJ1 at temperatures between 2 K and 6 K (see the curves in Figure 4(a)). Figure 4(b) shows the temperature dependence of Is, in which the solid curve is calculated by the Ambegaokar-Baratoff (AB) theory31, in which Is(T) is expressed as
(1) |
Where . Here and Tc are fitting parameters. Tc is found to be 5.83 K and in good agreement with the that of bulk NbS2 ( K), indicating no degradation of superconductivity at the vdW Josephson junction. The fitted superconducting energy gap meV is much smaller than the value of the standard BCS ratio meV. Previous studies point out that NbS2 exhibits two superconducting energy gaps11, 32, 33, 34. It is particularly interesting to note that the small gap falls below the BCS value, as confirmed by different measurement techniques including Andreev reflection spectroscopy11, scanning tunneling microscopy and spectroscopy32, specific heat measurements33, ac-calorimetry measurements34. The reported small gap ranges from 0.53 meV to 0.56 meV, in good agreement with our results. In NbSe2, it is speculated that the discrepancy between the measured gap and the BCS value is originated from the interaction between CDW and superconductivity9. Hence, our work in conjunction with previous studies suggest that even in the absence of CDW, the standard BCS theory cannot capture the superconducting behavior in NbS2 and TMD superconductor in general may arise from mechanisms other than CDW.
In conclusion, we successfully fabricated vdW-stacked vertical NbS2/NbS2 Josephson junctions by using a modified all-dry transfer technique. We confirmed the stacked structure indeed exhibits a pronounced Josephson effect via systemically transport measurements. The Josephson coupling was found to be strong, manifested as the large critical current, rendering the platform suitable for single flux quantum circuit. We also extract superconducting gap of 0.58 meV which is considerably smaller than the BCS prediction, this may give hints on the long-standing arguments on superconducting mechanism in NbS2 and other TMD in general.
We acknowledge the support from the National Natural Science Foundation of China (12074134) and the Analysis and Testing Center of Huazhong University of Science and Technology.
The following files are available free of charge.
Concept of a vdW Josephson junction; RT curve of NbS2 Bulk and flake device; I-V curves of additional NbS2/NbS2 junctions; Switching current statistics of JJ3; TEM image (PDF)
References
- Arute et al. 2019 Arute, F.; Arya, K.; Babbush, R.; Bacon, D.; Bardin, J. C.; Barends, R.; Biswas, R.; Boixo, S.; Brandao, F. G.; Buell, D. A., et al. Quantum supremacy using a programmable superconducting processor. Nature 2019, 574, 505–510
- Zeng et al. 2015 Zeng, L.; Nik, S.; Greibe, T.; Krantz, P.; Wilson, C.; Delsing, P.; Olsson, E. Direct observation of the thickness distribution of ultra thin AlOx barriers in Al/AlOx/Al Josephson junctions. J. Phys. D: Appl. Phys. 2015, 48, 395308
- McDermott 2009 McDermott, R. Materials origins of decoherence in superconducting qubits. IEEE Trans. Appl. Supercond. 2009, 19, 2–13
- Kim et al. 2021 Kim, S.; Terai, H.; Yamashita, T.; Qiu, W.; Fuse, T.; Yoshihara, F.; Ashhab, S.; Inomata, K.; Semba, K. Enhanced coherence of all-nitride superconducting qubits epitaxially grown on silicon substrate. Commun. Mater. 2021, 2, 1–7
- Yan et al. 2021 Yan, C.; Hassel, J.; Vesterinen, V.; Zhang, J.; Ikonen, J.; Grönberg, L.; Goetz, J.; Möttönen, M. A low-noise on-chip coherent microwave source. Nat. Electron 2021, 4, 885–892
- Yabuki et al. 2016 Yabuki, N.; Moriya, R.; Arai, M.; Sata, Y.; Morikawa, S.; Masubuchi, S.; Machida, T. Supercurrent in van der Waals Josephson junction. Nat. Commun. 2016, 7, 1–5
- Farrar et al. 2021 Farrar, L. S.; Nevill, A.; Lim, Z. J.; Balakrishnan, G.; Dale, S.; Bending, S. J. Superconducting quantum interference in twisted van der Waals heterostructures. Nano Lett. 2021, 21, 6725–6731
- Xi et al. 2015 Xi, X.; Zhao, L.; Wang, Z.; Berger, H.; Forró, L.; Shan, J.; Mak, K. F. Strongly enhanced charge-density-wave order in monolayer NbSe2. Nat. Nanotechnol. 2015, 10, 765–769
- Naaman et al. 2003 Naaman, O.; Dynes, R.; Bucher, E. Josephson Effect in Pb/I/NbSe2 Scanning Tunneling Microscope Junctions. Int. J. Mod. Phys. B 2003, 17, 3569–3574
- Naito and Tanaka 1982 Naito, M.; Tanaka, S. Electrical transport properties in 2H-NbS2, -NbSe2, -TaS2 and -TaSe2. J. Phys. Soc. Jpn. 1982, 51, 219–227
- Majumdar et al. 2020 Majumdar, A.; VanGennep, D.; Brisbois, J.; Chareev, D.; Sadakov, A. V.; Usoltsev, A.; Mito, M.; Silhanek, A. V.; Sarkar, T.; Hassan, A., et al. Interplay of charge density wave and multiband superconductivity in layered quasi-two-dimensional materials: The case of 2H-NbS2 and 2H-NbSe2. Phys. Rev. Mater. 2020, 4, 084005
- Yan et al. 2019 Yan, R.; Khalsa, G.; Schaefer, B. T.; Jarjour, A.; Rouvimov, S.; Nowack, K. C.; Xing, H. G.; Jena, D. Thickness dependence of superconductivity in ultrathin NbS2. Appl. Phys. Express 2019, 12, 023008
- Schwartz et al. 2019 Schwartz, J. J.; Chuang, H.-J.; Rosenberger, M. R.; Sivaram, S. V.; McCreary, K. M.; Jonker, B. T.; Centrone, A. Chemical identification of interlayer contaminants within van der Waals heterostructures. ACS Appl. Mater. Interfaces 2019, 11, 25578–25585
- Narla et al. 2014 Narla, A.; Sliwa, K.; Hatridge, M.; Shankar, S.; Frunzio, L.; Schoelkopf, R.; Devoret, M. Wireless josephson amplifier. Appl. Phys. Lett. 2014, 104, 232605
- Lian et al. 2017 Lian, H.; Wu, Y.; Xing, H.; Wang, S.; Liu, Y. Effect of stoichiometry on the superconducting transition temperature in single crystalline 2H-NbS2. Phys. C 2017, 538, 27–31
- Clarke and Wilhelm 2008 Clarke, J.; Wilhelm, F. K. Superconducting quantum bits. Nature 2008, 453, 1031–1042
- Tinkham 2004 Tinkham, M. Introduction to superconductivity; Dover Publications: Mineaola, NY, 2004
- Ando et al. 2020 Ando, F.; Miyasaka, Y.; Li, T.; Ishizuka, J.; Arakawa, T.; Shiota, Y.; Moriyama, T.; Yanase, Y.; Ono, T. Observation of superconducting diode effect. Nature 2020, 584, 373–376
- Narita et al. 2022 Narita, H.; Ishizuka, J.; Kawarazaki, R.; Kan, D.; Shiota, Y.; Moriyama, T.; Shimakawa, Y.; Ognev, A. V.; Samardak, A. S.; Yanase, Y., et al. Field-free superconducting diode effect in noncentrosymmetric superconductor/ferromagnet multilayers. Nat. Nanotechnol. 2022, 17, 823–828
- Barone et al. 1975 Barone, A.; Johnson, W.; Vaglio, R. Current flow in large Josephson junctions. J. Appl. Phys. 1975, 46, 3628–3632
- Schwidtal and Finnegan 1969 Schwidtal, K.; Finnegan, R. D. Barrier-Thickness Dependence of the dc Quantum Interference Effect in Thin-Film Lead Josephson Junctions. J. Appl. Phys. 1969, 40, 2123–2127
- Barone and Paterno 1982 Barone, A.; Paterno, G. Physics and applications of the Josephson Effect; Wiley: NY, 1982
- Gu et al. 1979 Gu, J.; Cha, W.; Gamo, K.; Namba, S. Properties of niobium superconducting bridges prepared by electron-beam lithography and ion implantation. J. Appl. Phys. 1979, 50, 6437–6442
- Rosenthal et al. 1991 Rosenthal, P. A.; Beasley, M.; Char, K.; Colclough, M.; Zaharchuk, G. Flux focusing effects in planar thin-film grain-boundary Josephson junctions. Appl. Phys. Lett. 1991, 59, 3482–3484
- Island et al. 2016 Island, J. O.; Steele, G. A.; Van Der Zant, H. S.; Castellanos-Gomez, A. Thickness dependent interlayer transport in vertical MoS2 Josephson junctions. 2D Mater. 2016, 3, 031002
- Paajaste et al. 2015 Paajaste, J.; Amado, M.; Roddaro, S.; Bergeret, F.; Ercolani, D.; Sorba, L.; Giazotto, F. Pb/InAs nanowire Josephson junction with high critical current and magnetic flux focusing. Nano Lett. 2015, 15, 1803–1808
- Tao et al. 2003 Tao, H.-J.; Li, Z.-Z.; Xuan, Y.; Ren, Z.-A.; Che, G.-C.; Zhao, B.-R.; Zhao, Z.-X. Josephson effects in MgB2 break junctions and MgB2/Nb point contacts. Phys. C: Supercond. 2003, 386, 569–574
- Nicoletti and Villegier 1997 Nicoletti, S.; Villegier, J.-C. Electrical behavior of YBa2Cu3O7-x grain boundary junctions under low magnetic field. J. Appl. Phys. 1997, 82, 303–308
- Kim et al. 2017 Kim, M.; Park, G.-H.; Lee, J.; Lee, J. H.; Park, J.; Lee, H.; Lee, G.-H.; Lee, H.-J. Strong proximity Josephson coupling in vertically stacked NbSe2–graphene–NbSe2 van der Waals junctions. Nano Lett. 2017, 17, 6125–6130
- Leroux et al. 2012 Leroux, M.; Rodiere, P.; Cario, L.; Klein, T. Anisotropy and temperature dependence of the first critical field in 2H-NbS2. Phys. B: Condens. Mater. 2012, 407, 1813–1815
- Ambegaokar and Baratoff 1963 Ambegaokar, V.; Baratoff, A. Tunneling between superconductors. Phys. Rev. Lett. 1963, 10, 486
- Guillamón et al. 2008 Guillamón, I.; Suderow, H.; Vieira, S.; Cario, L.; Diener, P.; Rodiere, P. Superconducting density of states and vortex cores of 2H-NbS2. Phys. Rev. Lett. 2008, 101, 166407
- Kačmarčík et al. 2010 Kačmarčík, J.; Pribulová, Z.; Marcenat, C.; Klein, T.; Rodière, P.; Cario, L.; Samuely, P. Specific heat measurements of a superconducting NbS2 single crystal in an external magnetic field: energy gap structure. Phys. Rev. B 2010, 82, 014518
- Kačmarčík et al. 2010 Kačmarčík, J.; Pribulová, Z.; Marcenat, C.; Klein, T.; Rodière, P.; Cario, L.; Samuely, P. Studies on two-gap superconductivity in 2H-NbS2. Phys. C 2010, 470, S719–S720