Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

RETRACTED ARTICLE: Quantized Majorana conductance

A Retraction to this article was published on 08 March 2021

An Addendum to this article was published on 29 April 2020

This article has been updated

Abstract

Majorana zero-modes—a type of localized quasiparticle—hold great promise for topological quantum computing1. Tunnelling spectroscopy in electrical transport is the primary tool for identifying the presence of Majorana zero-modes, for instance as a zero-bias peak in differential conductance2. The height of the Majorana zero-bias peak is predicted to be quantized at the universal conductance value of 2e2/h at zero temperature3 (where e is the charge of an electron and h is the Planck constant), as a direct consequence of the famous Majorana symmetry in which a particle is its own antiparticle. The Majorana symmetry protects the quantization against disorder, interactions and variations in the tunnel coupling3,4,5. Previous experiments6, however, have mostly shown zero-bias peaks much smaller than 2e2/h, with a recent observation7 of a peak height close to 2e2/h. Here we report a quantized conductance plateau at 2e2/h in the zero-bias conductance measured in indium antimonide semiconductor nanowires covered with an aluminium superconducting shell. The height of our zero-bias peak remains constant despite changing parameters such as the magnetic field and tunnel coupling, indicating that it is a quantized conductance plateau. We distinguish this quantized Majorana peak from possible non-Majorana origins by investigating its robustness to electric and magnetic fields as well as its temperature dependence. The observation of a quantized conductance plateau strongly supports the existence of Majorana zero-modes in the system, consequently paving the way for future braiding experiments that could lead to topological quantum computing.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Quantized Majorana zero-bias peak.
Figure 2: Quantized Majorana conductance plateau.
Figure 3: Majorana peak splitting.
Figure 4: Quantized Majorana plateau reproduced, and temperature dependence.
Figure 5: Trivial zero-bias peaks from Andreev bound states.

Similar content being viewed by others

Change history

References

  1. Kitaev, A. Y. Unpaired Majorana fermions in quantum wires. Phys. Uspekhi 44, 131–136 (2001)

    Article  ADS  Google Scholar 

  2. Mourik, V. et al. Signatures of Majorana fermions in hybrid superconductor-semiconductor nanowire devices. Science 336, 1003–1007 (2012)

    Article  ADS  CAS  Google Scholar 

  3. Law, K. T., Lee, P. A. & Ng, T. K. Majorana fermion induced resonant Andreev reflection. Phys. Rev. Lett. 103, 237001 (2009)

    Article  ADS  CAS  Google Scholar 

  4. Flensberg, K. Tunneling characteristics of a chain of Majorana bound states. Phys. Rev. B 82, 180516 (2010)

    Article  ADS  Google Scholar 

  5. Wimmer, M., Akhmerov, A. R., Dahlhaus, J. P. & Beenakker, C. W. J. Quantum point contact as a probe of a topological superconductor. New J. Phys. 13, 053016 (2011)

    Article  ADS  Google Scholar 

  6. Lutchyn, R. M. et al. Realizing Majorana zero modes in superconductor-semiconductor heterostructures. Preprint at https://arxiv.org/abs/1707.04899 (2017)

  7. Nichele, F. et al. Scaling of Majorana zero-bias conductance peaks. Phys. Rev. Lett. 119, 136803 (2017)

    Article  ADS  Google Scholar 

  8. Lutchyn, R. M., Sau, J. D. & Das Sarma, S. Majorana fermions and a topological phase transition in semiconductor-superconductor heterostructures. Phys. Rev. Lett. 105, 077001 (2010)

    Article  ADS  Google Scholar 

  9. Oreg, Y., Refael, G. & von Oppen, F. Helical liquids and Majorana bound states in quantum wires. Phys. Rev. Lett. 105, 177002 (2010)

    Article  ADS  Google Scholar 

  10. Sengupta, K., Žutic´, I., Kwon, H.-J., Yakovenko, V. M. & Das Sarma, S. Midgap edge states and pairing symmetry of quasi-one-dimensional organic superconductors. Phys. Rev. B 63, 144531 (2001)

    Article  ADS  Google Scholar 

  11. Majorana, E. A symmetric theory of electrons and positrons. Soryushiron Kenkyu 63, 149 (1981) [transl.]; Nuovo Cimento 14, 171–184 (1937)

    Article  CAS  Google Scholar 

  12. Read, N. & Green, D. Paired states of fermions in two dimensions with breaking of parity and time-reversal symmetries and the fractional quantum Hall effect. Phys. Rev. B 61, 10267–10297 (2000)

    Article  ADS  CAS  Google Scholar 

  13. Deng, M. T. et al. Majorana bound state in a coupled quantum-dot hybrid-nanowire system. Science 354, 1557–1562 (2016)

    Article  ADS  CAS  Google Scholar 

  14. Gül, Ö. et al. Ballistic Majorana nanowire devices. Nature Nanotech. 13, 192–197 (2018)

    Article  ADS  Google Scholar 

  15. Pientka, F., Kells, G., Romito, A., Brouwer, P. W. & von Oppen, F. Enhanced zero-bias Majorana peak in the differential tunneling conductance of disordered multisubband quantum-wire/superconductor junctions. Phys. Rev. Lett. 109, 227006 (2012)

    Article  ADS  Google Scholar 

  16. Prada, E., San-Jose, P. & Aguado, R. Transport spectroscopy of NS nanowire junctions with Majorana fermions. Phys. Rev. B 86, 180503 (2012)

    Article  ADS  Google Scholar 

  17. Lin, C.-H., Sau, J. D. & Das Sarma, S. Zero-bias conductance peak in Majorana wires made of semiconductor/superconductor hybrid structures. Phys. Rev. B 86, 224511 (2012)

    Article  ADS  Google Scholar 

  18. Rainis, D., Trifunovic, L., Klinovaja, J. & Loss, D. Towards a realistic transport modeling in a superconducting nanowire with Majorana fermions. Phys. Rev. B 87, 024515 (2013)

    Article  ADS  Google Scholar 

  19. Liu, C.-X., Sau, J. D. & Das Sarma, S. Role of dissipation in realistic Majorana nanowires. Phys. Rev. B 95, 054502 (2017)

    Article  ADS  Google Scholar 

  20. Krogstrup, P. et al. Epitaxy of semiconductor–superconductor nanowires. Nat. Mater. 14, 400–406 (2015)

    Article  ADS  CAS  Google Scholar 

  21. Gazibegovic, S. et al. Epitaxy of advanced nanowire quantum devices. Nature 548, 434–438 (2017)

    Article  ADS  CAS  Google Scholar 

  22. Gül, Ö. et al. Hard superconducting gap in InSb nanowires. Nano Lett. 17, 2690–2696 (2017)

    Article  ADS  Google Scholar 

  23. Zhang, H. et al. Ballistic superconductivity in semiconductor nanowires. Nat. Commun. 8, 16025 (2017)

    Article  ADS  CAS  Google Scholar 

  24. Takei, S., Fregoso, B. M., Hui, H.-Y., Lobos, A. M. & Das Sarma, S. Soft superconducting gap in semiconductor Majorana nanowires. Phys. Rev. Lett. 110, 186803 (2013)

    Article  ADS  Google Scholar 

  25. Stanescu, T. D., Tewari, S., Sau, J. D. & Das Sarma, S. To close or not to close: the fate of the superconducting gap across the topological quantum phase transition in Majorana-carrying semiconductor nanowires. Phys. Rev. Lett. 109, 266402 (2012)

    Article  ADS  Google Scholar 

  26. Mishmash, R. V., Aasen, D., Higginbotham, A. P. & Alicea, J. Approaching a topological phase transition in Majorana nanowires. Phys. Rev. B 93, 245404 (2016)

    Article  ADS  Google Scholar 

  27. Liu, C.-X., Setiawan, F., Sau, J. D. & Das Sarma, S. Phenomenology of the soft gap, zero-bias peak, and zero-mode splitting in ideal Majorana nanowires. Phys. Rev. B 96, 054520 (2017)

    Article  ADS  Google Scholar 

  28. Liu, C.-X., Sau, J. D., Stanescu, T. D. & Das Sarma, S. Andreev bound states versus Majorana bound states in quantum dot–nanowire-superconductor hybrid structures: trivial versus topological zero-bias conductance peaks. Phys. Rev. B 96, 075161 (2017)

    Article  ADS  Google Scholar 

  29. Das Sarma, S., Sau, J. D. & Stanescu, T. D. Splitting of the zero-bias conductance peak as smoking gun evidence for the existence of the Majorana mode in a superconductor–semiconductor nanowire. Phys. Rev. B 86, 220506 (2012)

    Article  ADS  Google Scholar 

  30. Zazunov, A., Egger, R. & Levy Yeyati, A. Low-energy theory of transport in Majorana wire junctions. Phys. Rev. B 94, 014502 (2016)

    Article  ADS  Google Scholar 

  31. Setiawan, F., Liu, C.-X., Sau, J. D. & Das Sarma, S. Electron temperature and tunnel coupling dependence of zero-bias and almost-zero-bias conductance peaks in Majorana nanowires. Phys. Rev. B 96, 184520 (2017)

    Article  ADS  Google Scholar 

  32. Kells, G., Meidan, D. & Brouwer, P. W. Near-zero-energy end states in topologically trivial spin–orbit coupled superconducting nanowires with a smooth confinement. Phys. Rev. B 86, 100503 (2012)

    Article  ADS  Google Scholar 

  33. Lee, E. J. H. et al. Spin-resolved Andreev levels and parity crossings in hybrid superconductor–semiconductor nanostructures. Nat. Nanotech. 9, 79–84 (2014)

    Article  ADS  CAS  Google Scholar 

  34. Pikulin, D. I., Dahlhaus, J. P., Wimmer, M., Schomerus, H. & Beenakker, C. W. J. A zero-voltage conductance peak from weak antilocalization in a Majorana nanowire. New J. Phys. 14, 125011 (2012)

    Article  ADS  Google Scholar 

  35. Stanescu, T. D. & Tewari, S. Disentangling Majorana fermions from topologically trivial low-energy states in semiconductor Majorana wires. Phys. Rev. B 87, 140504 (2013)

    Article  ADS  Google Scholar 

  36. Prada, E., Aguado, R. & San-Jose, P. Measuring Majorana nonlocality and spin structure with a quantum dot. Phys. Rev. B 96, 085418 (2017)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

We thank M. Wimmer and Ö. Gül for discussions. This work has been supported by the European Research Council, the Dutch Organization for Scientific Research, the Office of Naval Research, the Laboratory for Physical Sciences and Microsoft Corporation Station-Q.

Author information

Authors and Affiliations

Authors

Contributions

H.Z., D.X., G.W., N.v.L., J.D.S.B. and M.W.A.d.M. fabricated the devices, performed electrical measurements and analysed the experimental data. S.G., J.A.L., D.C., R.L.M.O.h.V., P.J.v.V., S.K., M.A.V., M.P., D.J.P., B.S., J.S.L., C.J.P. and E.P.A.M.B. grew the nanowires with epitaxial Al and performed the nanowire deposition. C.-X.L. and S.D.S. performed the numerical simulations. The manuscript was written by H.Z. and L.P.K. with comments from all authors.

Corresponding authors

Correspondence to Hao Zhang or Leo P. Kouwenhoven.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Additional information

Reviewer Information Nature thanks M. Franz and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Figure 1 Apparent ‘soft gap’ due to large Andreev reflection.

a, Differential conductance dI/dV of the device in Figs 1, 2, 3 (device A) as a function of bias voltage at zero magnetic field. The tunnel-gate voltage is tuned to more negative from the top curve to the bottom curve. The transmission probability of the tunnel barrier is tuned from large (black curve) to small (orange curve). In the low transmission regime (orange curve), where the above-gap conductance (about 0.03 × 2e2/h) is much less than 2e2/h, dI/dV is proportional to the density of states in the proximitized wire part, resolving a hard superconducting gap. In the high transmission regime (black curve), where the above-gap conductance is comparable with 2e2/h, the finite sub-gap conductance is due to large Andreev reflection. This ‘soft gap’ is not from dissipation, and does not affect the quantized ZBP height as shown in c. b, Re-plot of the two extreme curves from a, for clarity. c, Waterfall plot of Fig. 1b, showing all the individual curves from 0 T to 1 T in steps of 0.02 T. The curves are offset vertically by 0.066 × 2e2/h for clarity. The curve at 0 T and the red curve at 0.88 T correspond to the curves in Fig. 1c (left panel).

Extended Data Figure 2 Thermal-broadened ZBP in low transmission regime.

a, Differential conductance dI/dV of device D, as a function of B, showing a stable ZBP. b, Vertical line-cuts at 0 T, 0.88 T and 0.94 T. At B = 0 T, the above-gap conductance (approximately 0.05 × 2e2/h) is much less than 2e2/h, which means that the device is in the low transmission regime, and thus shows a hard gap. The tiny sub-gap conductance is due to the small Andreev reflection and the noise background of the measurement equipment. The low transmission leads to a narrow ZBP width, which is negligible compared with the thermal width of 3.5kBT. Thus, thermal averaging suppresses the ZBP height below the quantized value. The sub-gap conductance at finite B (for example, 0.88 T or 0.94 T), where the ZBP appears, is the same as the sub-gap conductance at zero field, indicating that the gap remains hard at high magnetic field where the Majorana state is present. c, The zoom-in curves show that the FWHM of the ZBP is about 28 μeV, which is consistent with the combined effect of the thermal broadening (3.5kBT ≈ 6 μeV at 20 mK), the lock-in bias voltage excitation (5 μeV) and broadening from tunnelling. This shows that the thermal broadening does indeed dominate over tunnel broadening. d, Waterfall plot of a with vertical offset of 0.01 × 2e2/h for clarity.

Extended Data Figure 3 Simulation of temperature dependence on the quantized ZBP.

a, False-colour scanning electron micrograph of device B with data shown in Fig. 4. Scale bar is 1 μm. The length of the Al section is about 0.9 μm. We calculate the dI/dV curve at high temperature by convolution of the derivative of the Fermi distribution function with the dI/dV curve at base temperature of 20 mK: , where T is temperature, V is bias voltage, and is the Fermi distribution function. Because we use the dI/dV curve at 20 mK as the zero-temperature data, our model only works for T sufficiently larger than 20 mK, that is, T > 50 mK. b, Comparison between the experimental data (left, taken from Fig. 4d) and theory simulations, for different temperatures. c, Several typical curves at different temperatures; black traces are the experimental data, and the red curves are the theory simulations with no fitting parameters. The agreement between simulation and experiment indicates that thermal averaging effect is the dominating effect that smears out the ZBP at high temperature. d, Temperature dependence of the ZBP taken from our theory model: Fig. 1c (right panel). The temperature varies from 25 mK to 700 mK in steps of 23 mK.

Extended Data Figure 4 Perfect ballistic Andreev transport in InSb–Al nanowires.

a, False-colour scanning electron micrograph of the device in Fig. 5 (device C). Scale bar is 500 nm. Electrical contacts and top gates are Cr/Au. Lower panel shows the device schematic and measurement set-up. The two top-gates (tunnel-gate and super-gate) are separated from the nanowire by 30-nm-thick SiN dielectric. The global back gate is p-doped Si covered by 285-nm-thick SiO2 dielectric. b, Differential conductance dI/dV, as a function of bias voltage (V) and tunnel-gate voltage at zero field. No localization effect (conductance resonances or quantum-dot-induced Coulomb blockade) is observed. c, Vertical line-cuts from b at tunnel-gate voltage of −0.18 V (lower panel) and −0.12 V (upper panel), showing a hard superconducting gap in the low transmission regime (lower panel) and strong Andreev enhancement in the open regime (upper panel). d, Horizontal line-cuts from c for V = 0 mV (pink, sub-gap conductance, GS) and V = 0.45 mV (green, above-gap conductance, GN). The blue curve is the calculated sub-gap conductance using GS = 4e2/h × T2/(2 − T)2, where the transmission T is extracted from the above-gap conductance: GN = (2e2/h) × T. e, Sub-gap conductance GS as a function of GN (black dots) and the theory prediction (red curve): GS = 2GN2/(2 − GN)2, with GS and GN in unit of 2e2/h. Both d and e show perfect agreement between theory and experiment. This indicates that the sub-gap conductance is indeed dominated by the Andreev reflection, that is, without contributions from sub-gap states. f, Magnetic field dependence of the hard gap. Lower panel shows the zero-bias line-cut. The gap remains hard up to 1 T, where the bulk superconducting gap closes.

Extended Data Figure 5 Majoranas versus trivial Andreev bound states.

a, b, Schematics of a Majorana nanowire device. The only difference between the left column (Majorana) and right column (ABS) is the chemical potential, as shown in c and d. c, d, Potential profile in the device. The tunnel barrier height is 10 meV and the width is 10 nm. The dot potential shape is x/ldot), for x between 0 and 0.3 μm, where the length of the dot (ldot) is 0.3 μm, and VD is the dot depth which can be tuned by the nearby gate, that is, the tunnel-gate. The rest of the flat nanowire segment is 1 μm long. We assume a pairing potential  = 0.2 meV, with a spin–orbit coupling of 0.5 eV Å. We set the Zeeman energy to be 1 meV, so the chemical potential of 0.5 meV (left) corresponds to the topological regime, and 1.2 meV (right) corresponds to the trivial regime, based on the topological condition , where μ is chemical potential. e, f, Spatial distribution of the Majorana and ABS wavefunctions in the topological and trivial regime. In the topological regime, two spatially well separated Majoranas (red and black) are localized at the two ends of the topological section. In the trivial regime, the Andreev bound state, which can be considered as two strongly overlapped Majoranas (red and black), is localized near the tunnel barrier. g, h, The Majorana ZBP remains non-split against the change of dot potential, regardless of the energy of the dot level. The green arrow indicates one bound state in the dot, whose wavefunction |Ψ2| is shown in e (green curve). When this dot level moves down, it is repelled from zero energy, where the Majorana ZBP remains bound to zero (inset of i). On the contrary, the ABS-induced ZBP is not robust at all and only shows up at the crossing points of two Andreev levels. This is because the tunnel-gate tunes the dot potential, which therefore affects the energy of the localized ABS near the tunnel barrier. i, j, The Majorana ZBP height shows a quantized plateau at 2e2/h by tuning the dot potential with tunnel-gate. The ZBP height drops from the quantized value (inset) when the ABS-dot level moves towards zero, which effectively squeezes the ZBP-width such that the thermal averaging effect starts to dominate. The ABS zero-bias conductance does not show a plateau, but instead varies between 0 and 4e2/h.

Extended Data Figure 6 Magnetic field dependence of trivial Andreev bound states.

a, Top panel is a re-plot of the trivial ABS data in Fig. 5a. Middle and bottom panels are the ZBP data at different back-gate voltages (labelled in the panels). b, Line-cuts of the ZBP data from a. The ZBP height varies with back-gate voltages and can exceed 2e2/h. The ZBP height at 2e2/h here is just a tuned coincidence.

Extended Data Figure 7 Specifics of devices.

We fabricated and tested many (over 60) devices out of which we selected 11 devices that showed good basic transport with all gates being fully functional. These were used for extensive measurements. Although most of these devices show ZBPs after tuning gate voltages and magnetic field, only two devices (presented in the main text: Figs 13 for device A and Fig. 4 for device B) show a quantized ZBP plateau. All other devices show trivial ZBPs similar to Fig. 5 (from device C). Scanning electron microscope images of devices A, B and C are shown in Fig. 1a, Extended Data Fig. 3a and Extended Data Fig. 4a, respectively. Here we show the scanning electron microscope images of the other eight devices, which we have explored extensively, but without finding a quantized ZBP plateau. Devices 1 and 2 are side-gate devices. Device 3 has a top tunnel-gate separated from the nanowire by 30-nm-thick SiN dielectric, and a global back-gate separated by 285-nm-thick SiO2. Devices 4 and 5 have tunnel-gate and super-gate on top separated from the nanowire by 30-nm-thick SiN dielectric. Devices 6 to 8 have two layers of top-gate. The bottom layer has a tunnel-gate separated by 30-nm-thick SiN dielectric while the top layer has super-gates separated by 30-nm-thick SiN from the bottom layer. The scale bar is 1 μm for all devices, except for device 2, which is 500 nm. It would be informative to perform Schrodinger–Poisson calculations on these different device geometries to determine the self-consistent potential landscape and find out which geometry suppresses a local potential dip near the tunnel barrier.

PowerPoint slides

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, H., Liu, CX., Gazibegovic, S. et al. RETRACTED ARTICLE: Quantized Majorana conductance. Nature 556, 74–79 (2018). https://doi.org/10.1038/nature26142

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature26142

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing