Play all audios:
ABSTRACT Majorana zero modes are leading candidates for topological quantum computation due to non-local qubit encoding and non-abelian exchange statistics. Spatially separated Majorana
modes are expected to allow phase-coherent single-electron transport through a topological superconducting island via a mechanism referred to as teleportation. Here we experimentally
investigate such a system by patterning an elongated epitaxial InAs-Al island embedded in an Aharonov-Bohm interferometer. With increasing parallel magnetic field, a discrete sub-gap state
in the island is lowered to zero energy yielding persistent 1_e_-periodic Coulomb blockade conductance peaks (_e_ is the elementary charge). In this condition, conductance through the
interferometer is observed to oscillate in a perpendicular magnetic field with a flux period of _h_/_e_ (_h_ is Planck’s constant), indicating coherent transport of single electrons through
the islands, a signature of electron teleportation via Majorana modes. SIMILAR CONTENT BEING VIEWED BY OTHERS ANOMALOUS UNIVERSAL CONDUCTANCE AS A HALLMARK OF NON-LOCALITY IN A
MAJORANA-HOSTED SUPERCONDUCTING ISLAND Article Open access 05 November 2022 ENHANCED MAJORANA STABILITY IN A THREE-SITE KITAEV CHAIN Article Open access 31 March 2025 CORRELATION BETWEEN TWO
DISTANT QUASIPARTICLES IN SEPARATE SUPERCONDUCTING ISLANDS MEDIATED BY A SINGLE SPIN Article Open access 24 April 2024 INTRODUCTION Initial experiments reporting signatures of Majorana zero
modes (MZMs) in hybrid superconductor–semiconductor nanowires focussed on zero-bias conductance peaks (ZBPs) using local tunneling spectroscopy1,2,3,4. Subsequently, Majorana islands
provided additional evidence of MZMs based on nearly 1_e_-spaced Coulomb blockade (CB) peaks5, and indicated a Rashba-like spin–orbit coupling with the spin–orbit field lying in-plane,
perpendicular to the wire axis6. Under some circumstances, these signatures can be mimicked by trivial modes7,8,9, motivating a new generation of experiments that explicitly probe non-local
properties, which are more difficult to mimic. For instance, non-locality of MZMs was recently investigated by measuring the energy splitting induced by the interaction of a quantum dot and
a zero-energy state in a hybrid nanowire10. Non-locality can also be accessed by interferometric measurements of a Majorana island, where CB couples separated MZMs and fixes fermion
parity11,12,13,14,15. In the topological regime, a Majorana island can coherently transfer a single-electron between its two ends through MZMs11,12. To demonstrate the effect, a Majorana
island can be embedded in the arm of an Aharonov–Bohm (AB) interferometer. If single-electron transport in both the reference arm and the Majorana island is coherent, conductance through the
interferometer is expected to show oscillations with a flux period _h_/_e_11,16. In addition, interferometry offers a way to distinguish between localized trivial modes and MZMs12,14. This
technique was used to investigate coherent transport in semiconductor quantum dots17,18,19,20. In this work, we study coherent single-electron transport through a Majorana island with an AB
interferometer. At high magnetic fields, we observe 1_e_ periodic CB peaks due to a discrete zero-energy state. In this case, we observe the conductance through the interferometer to
oscillate periodically for a flux _h_/_e_ piercing the AB ring, while oscillations are suppressed when the island shows a 2_e_ periodicity or is in the normal state. The observation of
conductance oscillations in the 1_e_ regime indicates coherent single-electron transport through the Majorana island, as predicted for electron teleportation mediated by MZMs11. RESULTS
MAJORANA ISLAND INTERFEROMETER Devices were fabricated using an InAs two-dimensional electron gas (2DEG) heterostructure covered by 8 nm of epitaxially grown Al21. The bare 2DEG (without Al)
showed a phase coherence length of _l__ϕ_ ~ 4 μm (see Supplementary Fig. 1). Figure 1a shows a micrograph of device 1 with a 1.2 μm long and 0.1 μm wide superconducting Al stripe formed by
wet etching. Ti/Au top-gates were evaporated on top of a 25 nm HfO2 dielectric grown by atomic layer deposition. We studied two lithographically similar interferometers with circumferences
_L_loop of 5.6 μm for device 1 and 5 μm for device 2. Applying a negative voltage, _V_pg, to the central gate serves two purposes. It depletes the 2DEG surrounding the Al wire to form both
the Majorana island and the AB ring center and also adjusts the chemical potential and charge occupancy of the island. Energizing all exterior gates confines the 2DEG into an AB
interferometer by connecting the Majorana island to a normal conducting reference arm. The resistance of the reference arm was independently tuned by a negative gate voltage _V_c. A
source-drain bias voltage (_V_sd) was applied to one lead and the resulting current and four-terminal voltage was recorded. The in-plane magnetic fields _B_∥ and _B_t, and perpendicular
field, _B_⊥, were controlled by a three-axis vector magnet. At low temperatures, tunneling of single electrons onto a Majorana island with a superconducting gap _Δ_ is suppressed by CB,
except at charge degeneracies. When the lowest sub-gap state energy, _E_0, exceeds the charging energy _E_c, ground-state degeneracies only occur between even-occupied states, resulting in
2_e_-periodic CB conductance peaks22. Odd-occupied ground states are lowered into the accessible spectrum by a Zeeman field, resulting in even–odd CB peak spacing when 0 < _E_0 < _E_c.
The difference in peak spacings between even and odd states, _S_ = _S_e − _S_o, is proportional to _E_05 (see Fig. 1b). For well-separated MZMs, _E_0 tends exponentially toward zero,
yielding 1_e_ periodic CB peaks with a discrete zero-bias state at consecutive charge degeneracy point5,23. Both observations are necessary for a MZM interpretation. When MZMs are not widely
separated, CB peak spacings oscillate with field and chemical potential5,6,7. CB SPECTROSCOPY We first investigated the Majorana island without interferometry by depleting a segment of the
reference arm (see Fig. 1a). Figure 1b shows zero-bias differential conductance _G_ = d_I_/d_V_ of the island as a function of parallel magnetic field _B_∥ and gate voltage _V_pg, which
controls the electron occupancy and chemical potential of the island. CB peaks are 2_e_ periodic at zero field and split around 2 T, becoming 1_e_ periodic as the sub-gap state moves toward
zero energy (see Fig. 3a for peak spacing analysis). Performing CB spectroscopy, that is, measuring _G_ as a function of both source-drain bias _V_sd and _V_pg reveals Coulomb diamonds (Fig.
1c–f). At low _B_∥, diamonds are 2_e_ periodic with distinct negative differential conductance (Fig. 1c), which transition to an even–odd peak spacing difference at moderate fields (Fig.
1d), similar to previous work on superconducting Coulomb islands5,6,22,24,25,26,27. At high fields, the 1_e_ periodic diamonds show a discrete ZBP for consecutive charge degeneracy points
that is well separated from the superconducting gap (Fig. 1e). This sub-gap feature remained at zero bias until the superconducting gap closure, and persists for 3 mV in _V_pg, corresponding
to an energy range of 0.8 meV. For systems with strong spin–orbit interaction, a helical gap of _g__μ_B_B_∥ is expected, where _μ_B and _g_ are the Bohr magneton and electron _g_-factor,
respectively. We estimate a helical gap of ~0.6 meV at _B_∥ = 2.5 T for a _g_-factor of 4, comparable to the span of the ZBP in _V_pg. The stability of the zero-bias state in both magnetic
field and in chemical potential is consistent with the MZM picture (see Supplementary Fig. 2)5, however, the observation of coherent single-electron transport is needed to draw conclusions
about non-locality. Below, we additionally show that the zero-bias state is sensitive to rotations of the in-plane field. The magnitude of _B_∥ where 1_e_ periodicity is observed is in
agreement with ZBPs measured in tunneling spectroscopy in InAs 2DEGs3. In contrast, as a function of _B_t the peak spacing showed an oscillating even–odd periodicity, and no consecutive ZBPs
were observed (see Supplementary Fig. 3d–f), as expected for extended modes in the wire6,9,28. The normal state 1_e_ regime of the Majorana island appears above _B_∥ ~ 3 T with _E_c = 80
μeV (Fig. 1f), where no discrete ZBPs are observed (see Supplementary Figs. 2 and 3a–c). INTERFEROMETRY AND COHERENT SINGLE-ELECTRON TRANSPORT The reference arm of the AB interferometer was
connected to the Majorana island by tuning _V_c from −8 to −3 V while _V_pg was compensated. This lifted the overall conductance by opening a path through the reference arm (see Fig. 2a-d).
Fig. 2e–h shows the conductance Δ_G_ through the full interferometer (with smooth background subtracted; see “Methods” section) as a function of _B_⊥ and gate voltage _V_pg, which control
the flux in the interferometer and occupancy of the island, respectively. Figure 2e shows small oscillations in Δ_G_(_B_⊥) at _B_∥ = 0 T for the 2_e_ periodic peaks. For _B_∥ = 2.2 T, where
the peak spacing is even–odd (Fig. 2f), the conductance showed moderate oscillations with a period Δ_B_⊥ = 1.5 mT. This periodicity corresponds to a single flux quantum _h_/_e_ threading an
area of ~2.7 μm2, consistent with the ring center area defined by _V_pg. This indicates coherent 1_e_ transport through both the reference arm and the Majorana island. At _B_∥ = 2.7 T, the
CB peak spacing is uniformly 1_e_, and oscillation amplitude is maximal (see Fig. 2g). When the Majorana island is driven normal, _B_∥ > 3 T, conductance oscillations are reduced,
becoming comparable to the low field oscillations (Fig. 2h). The appearance of strong _h_/_e_ periodic conductance oscillations in the 1_e_ regime of the island is a key experimental
signature of electron teleportation. The strength and periodicity of the oscillations are examined more quantitatively using Fourier power spectrum (PS) analysis (see “Methods” section). In
Fig. 2i–l, the PS of Δ_G_(_B_⊥) are shown. Increasing _B_∥ leads to a peak appearing around _f_ = 0.65 mT−1, the frequency expected for AB interference. The PS is maximized in the 1_e_
regime. To quantify the oscillations amplitude, \(\langle \tilde{A}\rangle\), we average the integrated PS (see “Methods” section). We next correlate the _B_∥ dependence of the oscillations
amplitude, \(\langle \tilde{A}\rangle (B_\parallel )\), with the _B_∥ dependence of the lowest sub-gap state, _E_0(_B_∥), of the island. The sub-gap energy is found from the difference
between even and odd CB peak spacings, averaged separately, 〈_S_〉 = 〈_S_e〉−〈_S_o〉 (see Fig. 1b). In Fig. 3a, 〈_S_〉 remains constant as a function of _B_∥ (indicating 2_e_ transport) until a
sub-gap state moves below _E_c, reaching zero at 2.2 T without overshoot (as expected for well separated MZMs in a long wire5,23). At low fields, where the CB periodicity is 2_e_, the
oscillation amplitude \(\langle \tilde{A}\rangle\) is small (Fig. 3b). When 〈_S_〉 approaches zero at high fields (_B_∥ > 2 T), \(\langle \tilde{A}\rangle\) exhibits a sharp increase that
coincides with the 2_e_ to 1_e_ transition. Above 3 T, the device is in the normal state and \(\langle \tilde{A}\rangle\) returned to the low value found in the 2_e_ regime. This comparison
shows that the oscillation amplitude is correlated with the energy of the lowest subgap state, and is maximal in the 1_e_ superconducting regime, as expected for electron teleportation.
Figure 3c, d shows a similar study for device 2. In Fig. 3c, 〈_S_〉 shows strong even–odd below 1 T, fluctuates around 〈_S_〉 = 0 between 1 and 2 T, then settles to 1_e_ (〈_S_〉 = 0) above 2 T.
CB spectroscopy reveals a discrete state that oscillates around zero bias in both _B_∥ and _V_pg without forming a stable 1_e_ periodic zero-bias peak (see Supplementary Fig. 5). These
oscillations about zero energy are compatible with hybridized Majorana modes due to wavefunction overlap resulting from the finite island length9,28. This overlap causes an energy splitting
that oscillates both in field and chemical potential5,6,27. Figure 3d shows that phase coherent transport first appears above 1 T and \(\langle \tilde{A}\rangle\) gradually increases until
reaching a maximum amplitude for 1_e_ peak spacing at 2.1 T, before diminishing in the normal state. In comparison to device 1, phase coherent transport appears when 〈_S_〉 oscillations about
zero, suggesting that hybridized Majorana modes characterized by an extended wavefunction may also contribute to coherent transport. The absence of energy oscillations may distinguish
non-local MZMs from hybridized Majorana modes (see Fig. 3a, c). We attribute the reduced oscillation amplitude in device 1 to result from the longer loop length _L_loop > _l__ϕ_, leading
to increased dephasing in the reference arm. We observe that conductance oscillations measured on opposite sides of a CB peak in the 1_e_ regime are out of phase (see yellow ticks in the
insets of Fig. 3) indicating a transmission phase shift of _π_ is acquired when the parity of the island is flipped. This demonstrates interferometric detection of island parity, which
offers a way to detect MZM parity as proposed by several recent topological qubit schemes13,29,30. In some cases, we found that the the phase shift is restored through the CB valley, such
that the same sides of adjacent CB peaks have the same phase. What determines whether there is phase recovery in the CB valley is not currently understood (see Supplementary Fig.
6)15,18,19,20. The angular dependence of the in-plane magnetic field is investigated by fixing the field magnitude _B_r = 2.5 T and rotating the field by an angle _α_ (see Fig. 1).
Theoretically, a rotation of the in-plane field towards the Rashba field direction is expected to close the topological gap31. Figure 4a shows 1_e_ periodic Coulomb diamonds at _B_∥ = 2.5 T
with a discrete ZBP at each charge degeneracy point (similar to Fig. 1e). Rotating by an angle _α_ = 5° lifted the discrete state from zero energy, leading to even-odd peak spacing; at _α_ =
10°, 1_e_ periodicity is recovered, though without a discrete ZBP. The observed sensitivity of the zero-energy state to in-plane field rotation is consistent with MZMs31. Small rotations
(∣_α_∣ < 7.5°) reduced the oscillation amplitude, \(\langle \tilde{A}\rangle\), as expected for even–odd periodicity (see Fig. 3). However, at larger angles (∣_α_∣ > 10°) where the
discrete ZBP is absent, a strong interference signal is observed (Fig. 4d). The observation of coherent transport in the absence of a discrete zero-energy state suggests trivial extended
modes are also phase coherent over the length of the island. Therefore, the additional information provided by bias spectroscopy is needed to distinguish teleportation from other coherent
transport mechanisms, as shown in Fig. 4a–c. We further studied the effect of different magnetic field directions. The results are shown in Supplementary Fig. 7. In summary, all three axes
showed coherent transport, with oscillation amplitude first increasing as 〈_S_〉 approached zero. This shows that the oscillation amplitude is dictated by the energy _E_0 in all field
directions and indicates that interference is not unique to a parallel magnetic field. Finally, we comment on the physical mechanism that correlates the oscillation amplitude to the energy
of _E_0. At low fields, the Majorana island favors an even parity where transport of electrons occurs as two sequential tunneling events on either end of the island22,25. The two electrons
acquire the condensate phase when forming a Cooper pair, which suppresses single-electron coherence. At moderate fields, a discrete sub-gap state is brought below _E_c and a single-electron
transport channel is opened, allowing coherent resonant tunneling through the Majorana island. When the discrete state is brought to zero-energy, the contribution of coherent transport is
increased due to electron teleportation. Finally, in the normal state, we interpret the reduction in interference signal to reflect the short coherence length in the diffusive Al wire. In
conclusion, we report signatures of single-electron teleportation via non-local MZMs using AB interference in combination with spectroscopy of a discrete zero-energy state. Our results also
reveal that coherent transport by topologically trivial modes extending over the full length of the Majorana island are allowed. These extended trivial modes may be precursors of topological
states in a finite length system that could transition into non-local MZMs by adjusting experimental parameters28. We have shown that interferometry accompanying bias spectroscopy revealing
stable 1_e_ periodic CB in magnetic field and chemical potential can discriminate non-local MZMs from extended modes (that display characteristic energy oscillations). Increasing the wire
length to greatly exceed the diffusive coherence length \(\xi =\sqrt{{\xi }_{0}\,{l}_{e}} \sim 1\) μm (for _Δ_ = 75 μeV at _B_∥ = 2.5 T), where _ξ_0 is the clean coherence length and _l_e ~
300 nm is the semiconducting mean free path will suppress 1_e_ transport via trivial extended modes32. The observation of coherent transport through the island rules out localized ABS at the
ends of the wire as the source of the studied zero-bias state. Indeed, transport flips the parity of localized modes and suppresses interference, while transport through MZMs preserves
island parity and coherent transport12,14. These localized modes could have been a possible interpretation of the previously observed ZBPs in single-end tunneling experiments8,9. These
results suggest that InAs–Al 2DEGs are a promising route towards more complex experiments related to the braiding or fusion of MZMs. We have established coherent transport and parity readout
from the transmission phase shifts in Majorana islands, two key results for future topological qubit networks13,29,30. Future devices will take advantage of improved material quality to
allow for increased wire lengths to suppress coherent trivial quasiparticle transport, allowing MZM contributions to be better separated from other potential contributions. METHODS WAFER
STRUCTURE The devices were fabricated on wafers grown by molecular beam epitaxy on a InP substrate. The wafer stack consists of a 1 μm graded In1−_x_Al_x_As insulating buffer, a 4 nm
In0.81Ga0.19As bottom barrier, a 5 nm InAs quantum well, and a top barrier consisting of 5 nm In0.9Al0.1As for device 1 and 10 nm In0.81Ga0.19As for device 2. A 7 nm film of epitaxial Al was
then grown in situ without breaking the vacuum of the chamber. The InAs 2DEGs were characterized with a Hall bar geometry (Al removed), which showed a peak mobility of _μ_ = 17,000 cm2 V−1
s−1 for an electron density of _n_ = 1.7 × 1012 cm−2 and _n_ = 7.5 × 1011 cm−2 for devices 1 and 2, respectively. DEVICE FABRICATION Devices were fabricated using standard electron beam
lithography techniques. The devices were electrically isolated by etching mesa structures by first removing the top Al film with Al etchant Transene D, followed by a deep III–V chemical wet
etch H2O:C6H8O7:H3PO4:H2O2 (220:55:3:3). Next, the Al film on the mesa was selectively etched with Al etchant Transene D to produce the Al strip. A 25 nm-thick layer of insulating HfO2 was
grown by atomic layer deposition at a temperature of 90 °C over the entire sample. Top gates of Ti/Au (5/25 nm) were then evaporated and connected to bonding pads with leads of Ti/Au (5/250
nm). MEASUREMENTS Electrical measurements were performed by standard lock-in techniques at 166 Hz by applying the sum of a variable dc bias voltage _V_sd and an ac excitation voltage of 3–10
μV applied to one of the top ohmic contacts as shown in Fig. 1a. The resulting current across the device was recorded by grounding a bottom ohmic via a low-impedance current-to-voltage
converter, and the four terminal voltage was measured by an ac voltage amplifier with an input impedance of 500 MΩ. All measurements were taken in a dilution refrigerator with a base
temperature of 20 mK and an electron temperature of 40 mK estimated by the temperature dependence saturation of ZBP conductance3. DATA ANALYSIS To highlight the oscillating components of the
differential conductance _G_(_B_⊥), a smooth background was subtracted with a low-degree polynomial Savitzky–Golay filter resulting in Δ_G_33. Analysis of the oscillations was performed by
first performing a fast Fourier transform _F_(_f_) of Δ_G_(_B_⊥) using a Hanning window then calculating the power spectral density _P__S_(_f_) = ∣_F_(_f_)∣2. The oscillation amplitude
\(\langle \tilde{A}\rangle\) was obtained by averaging integrated power spectra. The integration was limited to a band in frequency between _f_1 = 0.55 mT−1 and _f_2 = 0.75 mT−1, spanning
the range of a single flux quantum _Φ_0 = _h_/_e_ through the area _A_ defined by either the central gate (_A_1 = 2.25 μm2) or the exterior gates (_A_2 = 3.1 μm2), where _f_ = _A_/_Φ_0 (see
Fig. 2i–l). DATA AVAILABILITY The data that support the findings of this study are available from the corresponding authors on reasonable request. REFERENCES * Mourik, V. et al. Signatures
of Majorana fermions in hybrid superconductor–semiconductor nanowire devices. _Science_ 336, 1003–1007 (2012). Article ADS CAS Google Scholar * Deng, M. et al. Majorana bound state in a
coupled quantum-dot hybrid-nanowire system. _Science_ 354, 1557–1562 (2016). Article ADS CAS Google Scholar * Nichele, F. et al. Scaling of Majorana zero-bias conductance peaks. _Phys.
Rev. Lett._ 119, 136803 (2017). Article ADS Google Scholar * Zhang, H. et al. Quantized Majorana conductance. _Nature_ 556, 74–79 (2018). Article ADS CAS Google Scholar * Albrecht, S.
M. et al. Exponential protection of zero modes in Majorana islands. _Nature_ 531, 206–209 (2016). Article ADS CAS Google Scholar * O’Farrell, E. C. T. et al. Hybridization of sub-gap
states in one-dimensional superconductor/semiconductor Coulomb islands. _Phys. Rev. Lett._ 121, 256803 (2018). Article ADS Google Scholar * Chiu, C.-K., Sau, J. D. & Sarma, S. D.
Conductance of a superconducting Coulomb-blockaded Majorana nanowire. _Phys. Rev. B_ 96, 054504 (2017). Article ADS Google Scholar * Moore, C., Stanescu, T. D. & Tewari, S.
Two-terminal charge tunneling: disentangling Majorana zero modes from partially separated Andreev bound states in semiconductor–superconductor heterostructures. _Phys. Rev. B_ 97, 165302
(2018). Article ADS CAS Google Scholar * Prada, E. et al. From Andreev to Majorana bound states in hybrid superconductor–semiconductor nanowires. Preprint at arXiv:1911.04512 (2019). *
Deng, M.-T. et al. Nonlocality of Majorana modes in hybrid nanowires. _Phys. Rev. B_ 98, 085125 (2018). Article ADS CAS Google Scholar * Fu, L. Electron teleportation via Majorana bound
states in a mesoscopic superconductor. _Phys. Rev. Lett._ 104, 056402 (2010). Article ADS Google Scholar * Sau, J. D., Swingle, B. & Tewari, S. Proposal to probe quantum nonlocality
of Majorana fermions in tunneling experiments. _Phys. Rev. B_ 92, 020511 (2015). Article ADS Google Scholar * Vijay, S. & Fu, L. Teleportation-based quantum information processing
with Majorana zero modes. _Phys. Rev. B_ 94, 235446 (2016). Article ADS Google Scholar * Hell, M., Flensberg, K. & Leijnse, M. Distinguishing Majorana bound states from localized
Andreev bound states by interferometry. _Phys. Rev. B_ 97, 161401 (2018). Article ADS CAS Google Scholar * Drukier, C., Zirnstein, H.-G., Rosenow, B., Stern, A. & Oreg, Y. Evolution
of the transmission phase through a Coulomb-blockaded Majorana wire. _Phys. Rev. B_ 98, 161401 (2018). Article ADS CAS Google Scholar * Aharonov, Y. & Bohm, D. Significance of
electromagnetic potentials in the quantum theory. _Phys. Rev._ 115, 485 (1959). Article ADS MathSciNet Google Scholar * Yacoby, A., Schuster, R. & Heiblum, M. Phase rigidity and
_h_/2_e_ oscillations in a single-ring Aharonov–Bohm experiment. _Phys. Rev. B_ 53, 9583 (1996). Article ADS CAS Google Scholar * Schuster, R. et al. Phase measurement in a quantum dot
via a double-slit interference experiment. _Nature_ 385, 417–420 (1997). Article ADS CAS Google Scholar * Avinun-Kalish, M., Heiblum, M., Zarchin, O., Mahalu, D. & Umansky, V.
Crossover from ‘mesoscopic’ to ‘universal’ phase for electron transmission in quantum dots. _Nature_ 436, 529–533 (2005). Article ADS CAS Google Scholar * Edlbauer, H. et al.
Non-universal transmission phase behaviour of a large quantum dot. _Nat. Commun._ 8, 1710 (2017). Article ADS Google Scholar * Shabani, J. et al. Two-dimensional epitaxial
superconductor–semiconductor heterostructures: a platform for topological superconducting networks. _Phys. Rev. B_ 93, 155402 (2016). Article ADS Google Scholar * Hekking, F. W. J.,
Glazman, L. I., Matveev, K. A. & Shekhter, R. I. Coulomb blockade of two-electron tunneling. _Phys. Rev. Lett._ 70, 4138 (1993). Article ADS CAS Google Scholar * Van Heck, B.,
Lutchyn, R. M. & Glazman, L. I. Conductance of a proximitized nanowire in the Coulomb blockade regime. _Phys. Rev. B_ 93, 235431 (2016). Article ADS Google Scholar * Tuominen, M. T.,
Hergenrother, J. M., Tighe, T. S. & Tinkham, M. Experimental evidence for parity-based 2e periodicity in a superconducting single-electron tunneling transistor. _Phys. Rev. Lett._ 69,
1997 (1992). Article ADS CAS Google Scholar * Eiles, T. M., Martinis, J. M. & Devoret, M. H. Even–odd asymmetry of a superconductor revealed by the Coulomb blockade of Andreev
reflection. _Phys. Rev. Lett._ 70, 1862 (1993). Article ADS CAS Google Scholar * Higginbotham, A. P. et al. Parity lifetime of bound states in a proximitized semiconductor nanowire.
_Nat. Phys._ 11, 1017–1021 (2015). Article CAS Google Scholar * Shen, J. et al. Parity transitions in the superconducting ground state of hybrid InSb–Al Coulomb islands. _Nat. Commun._ 9,
4801 (2018). Article ADS Google Scholar * Stanescu, T. D., Lutchyn, R. M. & Sarma, S. D. Dimensional crossover in spin–orbit-coupled semiconductor nanowires with induced
superconducting pairing. _Phys. Rev. B_ 87, 094518 (2013). Article ADS Google Scholar * Karzig, T. et al. Scalable designs for quasiparticle-poisoning-protected topological quantum
computation with Majorana zero modes. _Phys. Rev. B_ 95, 235305 (2017). Article ADS Google Scholar * Plugge, S., Rasmussen, A., Egger, R. & Flensberg, K. Majorana box qubits. _New J.
Phys._ 19, 012001 (2017). Article ADS Google Scholar * Osca, J., Ruiz, D. & Serra, L. Effects of tilting the magnetic field in one-dimensional Majorana nanowires. _Phys. Rev. B_ 89,
245405 (2014). Article ADS Google Scholar * Tinkham, M. _Introduction to Superconductivity_ (Courier Corporation, 2004). * Savitzky, A. & Golay, M. J. Smoothing and differentiation of
data by simplified least squares procedures. _Anal. Chem._ 36, 1627–1639 (1964). Article ADS CAS Google Scholar Download references ACKNOWLEDGEMENTS This work was supported by Microsoft
Corporation, the Danish National Research Foundation, and the Villum Foundation. F.N. acknowledges support from European Research Commission, Grant No. 804273. We thank Karsten Flensberg,
Joshua Folk, Michael Hell, Andrew Higginbotham, Torsten Karzig, Panagiotis Kotetes, Martin Leijnse, Tommy Li, Alice Mahoney, and Ady Stern for useful discussions. AUTHOR INFORMATION Author
notes * F. Nichele Present address: IBM Research - Zurich, Sumerstrasse 4, 8803, Rschlikon, Switzerland * These authors contributed equally: A. M. Whiticar, A. Fornieri. AUTHORS AND
AFFILIATIONS * Center for Quantum Devices, Niels Bohr Institute, University of Copenhagen and Microsoft Quantum Lab Copenhagen, Universitetsparken 5, Copenhagen, 2100, Denmark A. M.
Whiticar, A. Fornieri, E. C. T. O’Farrell, A. C. C. Drachmann, C. M. Marcus & F. Nichele * Department of Physics and Astronomy and Microsoft Quantum Lab Purdue, Purdue University, West
Lafayette, IN, 47907, USA T. Wang, C. Thomas, S. Gronin, R. Kallaher, G. C. Gardner & M. J. Manfra * Birck Nanotechnology Center, Purdue University, West Lafayette, IN, 47907, USA T.
Wang, C. Thomas, S. Gronin, R. Kallaher, G. C. Gardner & M. J. Manfra * School of Materials Engineering, Purdue University, West Lafayette, IN, 47907, USA M. J. Manfra * School of
Electrical and Computer Engineering, Purdue University, West Lafayette, IN, 47907, USA M. J. Manfra Authors * A. M. Whiticar View author publications You can also search for this author
inPubMed Google Scholar * A. Fornieri View author publications You can also search for this author inPubMed Google Scholar * E. C. T. O’Farrell View author publications You can also search
for this author inPubMed Google Scholar * A. C. C. Drachmann View author publications You can also search for this author inPubMed Google Scholar * T. Wang View author publications You can
also search for this author inPubMed Google Scholar * C. Thomas View author publications You can also search for this author inPubMed Google Scholar * S. Gronin View author publications You
can also search for this author inPubMed Google Scholar * R. Kallaher View author publications You can also search for this author inPubMed Google Scholar * G. C. Gardner View author
publications You can also search for this author inPubMed Google Scholar * M. J. Manfra View author publications You can also search for this author inPubMed Google Scholar * C. M. Marcus
View author publications You can also search for this author inPubMed Google Scholar * F. Nichele View author publications You can also search for this author inPubMed Google Scholar
CONTRIBUTIONS T.W., C.T., S.G., R.K., G.C.G. and M.J.M. developed and grew the InAs/Al heterostructure. A.M.W. and A.F. designed and fabricated the devices. A.M.W. and A.F. performed the
measurements. Data analysis was done by A.M.W. and A.F with input from C.M.M. and F.N. C.M.M. and F.N. conceived the experiment. A.M.W. and A.F. interpreted the data with input from
E.C.T.O., A.C.C.D., C.M.M. and F.N. The manuscript was written by A.M.W., A.F., C.M.M. and F.N. with suggestions from all other authors. CORRESPONDING AUTHORS Correspondence to C. M. Marcus
or F. Nichele. ETHICS DECLARATIONS COMPETING INTERESTS The authors declare no competing interests. ADDITIONAL INFORMATION PEER REVIEW INFORMATION _Nature Communications_ thanks Ramón Aguado
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. SUPPLEMENTARY INFORMATION 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 license, and indicate if changes were made. The images or other third party material in this article are included in the
article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license 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 license, visit
http://creativecommons.org/licenses/by/4.0/. Reprints and permissions ABOUT THIS ARTICLE CITE THIS ARTICLE Whiticar, A.M., Fornieri, A., O’Farrell, E.C.T. _et al._ Coherent transport through
a Majorana island in an Aharonov–Bohm interferometer. _Nat Commun_ 11, 3212 (2020). https://doi.org/10.1038/s41467-020-16988-x Download citation * Received: 09 March 2020 * Accepted: 17 May
2020 * Published: 25 June 2020 * DOI: https://doi.org/10.1038/s41467-020-16988-x SHARE THIS ARTICLE Anyone you share the following link with will be able to read this content: Get shareable
link Sorry, a shareable link is not currently available for this article. Copy to clipboard Provided by the Springer Nature SharedIt content-sharing initiative