+ All Categories
Home > Documents > Majorana's wires

Majorana's wires

Date post: 03-Dec-2016
Category:
Upload: marcel
View: 217 times
Download: 1 times
Share this document with a friend
4
NATURE NANOTECHNOLOGY | VOL 8 | MARCH 2013 | www.nature.com/naturenanotechnology 149 commentary Majorana’s wires Marcel Franz Experiments on nanowires have shown evidence of solid-state analogues of the particles predicted by Ettore Majorana more than 70 years ago. Although stronger confirmation is still to come, these first observations have already fuelled expectations of fundamental results and potential applications in quantum information technology. W hen in 1937 Ettore Majorana discovered a purely real-valued solution 1 to the celebrated Dirac equation, he could not have foreseen the whirlwind of activity that would follow — some 70 years later — and not in particle physics, which was his domain, but in nanoscience and condensed-matter physics. Majorana fermions, as the particles described by these solutions became known, are curious objects. e recent storm of activity in condensed-matter physics has focused on the ‘Majorana zero modes’, that is, emergent Majorana particles occurring at exactly zero energy that have a remarkable property of being their own antiparticles 2,3 . Mathematically, this property is expressed as an equality between the particle’s creation and annihilation operators γ  = γ. Any ordinary fermion can be thought of as being composed of two Majorana fermions. An interesting situation arises when a single Majorana particle can be spatially separated from its partner and independently probed. Observation of such an ‘unpaired’ Majorana particle in a solid-state system would clearly fulfil a longstanding intellectual challenge. Furthermore, Majorana zero modes are believed to exhibit the so called non-Abelian exchange statistics 4,5 , which endows them with a technological potential as building blocks of future quantum memory immune against many sources of decoherence that plague other such proposed devices. Recent advances in our understanding of solids with strong spin–orbit coupling, combined with the progress in nanofabrication, put the physical realization of the Majorana particles within reach. In fact, signatures consistent with their existence in quantum wires coupled to conventional superconductors in a set-up schematically depicted in Fig. 1 have been reported by several groups 6–10 . e theory behind these devices is very well understood — it is rooted in the standard band theory of solids and the Bardeen–Cooper–Schrieffer theory of superconductivity — and there is no doubt that Majorana zero modes should appear under the right conditions. e key question that remains to be answered is: Have the right conditions been achieved in existing devices? Why wires? Historically, there has been a number of proposals to engineer and detect Majorana zero modes in two-dimensional (2D) solid-state systems, including the fractional quantum Hall liquids 11 , interacting quantum spin systems 12 , spin-polarized p-wave superconductors 13 , and more recently interfaces between topological insulators 14 or semiconductors 15,16 and ordinary superconductors. Despite significant progress, especially in the quantum Hall liquids, none of these proposals has seen a decisive experimental confirmation. Instead, the focus over the past two years has shiſted to 1D structures — quantum wires — that are thought to possess several distinct advantages when it comes to fabrication and subsequent detection of the Majorana zero modes. In quantum wires Majoranas occur either at the wire end or at a domain wall between topological and non-topological regions of the wire. is facilitates relatively easy detection compared with 2D systems where Majoranas are in the cores of magnetic vortices or other topological defects, which can be located essentially anywhere in the sample and are thus difficult to find. e second key advantage is the expected paucity, relative to the 2D systems, Gates Superconductor Metal B Nanowire Figure 1 | A typical experimental set-up for the Majorana zero mode detection in a nanowire. The nanowire is placed on a substrate equipped with gates and contacted from above by superconducting and normal metal electrodes. The red arrows indicate the applied magnetic field B. © 2013 Macmillan Publishers Limited. All rights reserved
Transcript
Page 1: Majorana's wires

NATURE NANOTECHNOLOGY | VOL 8 | MARCH 2013 | www.nature.com/naturenanotechnology 149

commentary

Majorana’s wiresMarcel Franz

Experiments on nanowires have shown evidence of solid-state analogues of the particles predicted by Ettore Majorana more than 70 years ago. Although stronger confirmation is still to come, these first observations have already fuelled expectations of fundamental results and potential applications in quantum information technology.

When in 1937 Ettore Majorana discovered a purely real-valued solution1 to the celebrated Dirac

equation, he could not have foreseen the whirlwind of activity that would follow — some 70 years later — and not in particle physics, which was his domain, but in nanoscience and condensed-matter physics. Majorana fermions, as the particles described by these solutions became known, are curious objects. The recent storm of activity in condensed-matter physics has focused on the ‘Majorana zero modes’, that is, emergent Majorana particles occurring at exactly zero energy that have a remarkable property of being their own antiparticles2,3. Mathematically, this property is expressed as an equality between the particle’s creation and annihilation operators γ† = γ. Any ordinary fermion can be thought of as being composed of two Majorana fermions. An interesting situation arises when a single Majorana particle can be spatially separated from its partner and independently probed. Observation of such an ‘unpaired’ Majorana particle in a solid-state system would clearly fulfil a longstanding intellectual challenge. Furthermore, Majorana zero modes are believed to exhibit the so called non-Abelian exchange statistics4,5, which endows them with a technological potential as building blocks of future quantum memory immune against many sources of decoherence that plague other such proposed devices.

Recent advances in our understanding of solids with strong spin–orbit coupling, combined with the progress in nanofabrication, put the physical realization of the Majorana particles within reach. In fact, signatures consistent with their existence in quantum wires coupled to conventional superconductors in a set-up schematically depicted in Fig. 1 have been reported by several groups6–10. The theory behind these devices is very

well understood — it is rooted in the standard band theory of solids and the Bardeen–Cooper–Schrieffer theory of superconductivity — and there is no doubt that Majorana zero modes should appear under the right conditions. The key question that remains to be answered is: Have the right conditions been achieved in existing devices?

Why wires?Historically, there has been a number of proposals to engineer and detect Majorana zero modes in two-dimensional (2D) solid-state systems, including the fractional quantum Hall liquids11, interacting quantum spin systems12, spin-polarized p-wave superconductors13, and more recently interfaces between topological insulators14 or semiconductors15,16 and ordinary

superconductors. Despite significant progress, especially in the quantum Hall liquids, none of these proposals has seen a decisive experimental confirmation. Instead, the focus over the past two years has shifted to 1D structures — quantum wires — that are thought to possess several distinct advantages when it comes to fabrication and subsequent detection of the Majorana zero modes. In quantum wires Majoranas occur either at the wire end or at a domain wall between topological and non-topological regions of the wire. This facilitates relatively easy detection compared with 2D systems where Majoranas are in the cores of magnetic vortices or other topological defects, which can be located essentially anywhere in the sample and are thus difficult to find. The second key advantage is the expected paucity, relative to the 2D systems,

Gates

Superconductor

Metal

BNanowire

Figure 1 | A typical experimental set-up for the Majorana zero mode detection in a nanowire. The nanowire is placed on a substrate equipped with gates and contacted from above by superconducting and normal metal electrodes. The red arrows indicate the applied magnetic field B.

© 2013 Macmillan Publishers Limited. All rights reserved

Page 2: Majorana's wires

150 NATURE NANOTECHNOLOGY | VOL 8 | MARCH 2013 | www.nature.com/naturenanotechnology

commentary

of various low-energy excitations that could interfere with the detection of Majorana zero modes. This is due to the 1D confinement of electronic and other degrees of freedom in the wire geometry. Finally, there have been significant recent advances in fabrication and manipulation of clean quantum wires that allow for an unprecedented level of control and analysis in a wide range of settings.

The prototype on which all the current 1D devices, both proposed and realized, are based is the Kitaev chain17 (Fig. 2). The model, elegant in its simplicity, describes spinless electrons hopping between the sites of a 1D tight-binding chain and subject to superconducting pairing with p-wave symmetry (that is, on the bonds connecting the neighbouring sites). Kitaev observed that depending on the model parameters, the hopping amplitude t and the superconducting (SC) pairing amplitude Δ, the system can be in two distinct phases. If we think of the fermions on each lattice site j as being composed of two Majorana fermions, cj = γj1 + iγj2, then the trivial phase can be depicted as in the top panel of Fig. 2, and has Majorana fermions bound in pairs on each site. In the other phase, Majoranas on the neighbouring sites bind to form an ordinary fermion, leaving an unpaired Majorana at each end of the chain as illustrated in the bottom panel. This is the topological SC phase that underlies all the recent proposals for engineering Majorana fermions in 1D devices.

Physical realizations of the Kitaev chainTwo basic realizations of the Kitaev chain exist. One is based on quantum wires made of a semiconductor with strong spin–orbit coupling such as InSb or InAs, and the other employs wires made of a 3D topological insulator such as Bi2Se3. In both cases, superconductivity in the wire must be induced by the proximity effect in a set-up schematically depicted in Fig. 1.

Magnetic field is used to produce effectively spinless electrons in the wires as required by the Kitaev paradigm.

We discuss first the semiconductor wire implementation18,19 that has attracted the most attention by far over the past two years. The well-known excitation spectrum of such a wire is illustrated in Fig. 3a: the Rashba spin–orbit coupling separates the parabolic bands for two spin projections and the Zeeman field opens up a gap VZ = gB near k = 0 leading to an effectively spinless 1D metal when the chemical potential μ lies in the Zeeman gap (g is the magnetic g-factor, B is the applied magnetic field and k is momentum). This is the key condition that must be met to realize a 1D topological superconductor; more generally one requires an odd number of Fermi points in the right half of the Brillouin zone17. Superconducting order in such a wire gives rise to the topological phase when the following condition for Δ is satisfied: VZ > √Δ—2 +— μ—2 . In this regime the semiconductor wire realizes the Kitaev chain paradigm and will have Majorana zero modes localized at its ends.

The condition on VZ, Δ and μ listed above imposes some considerable constraints on the physical realization of the topological phase20. For typical values of the magnetic g-factor (15 and 50 for InAs and InSb wires, respectively) and for the magnetic fields of a few Tesla one obtains VZ @ 1–10 K. Tuning μ with this accuracy and ensuring that it is also sufficiently homogeneous so that the condition is satisfied everywhere along the length of the wire represents a significant experimental challenge. Also, the smallness of VZ restricts the experimental window for Majorana fermion observation and manipulation to low temperatures T << VZ. Yet, several groups have reported signatures in wires consistent with the existence of Majorana zero modes6–10. If true, this is a remarkable achievement, although as we discuss below

there exist alternative interpretations of these experiments that do not involve Majorana zero modes.

The above-mentioned constraint is relaxed in quantum wires made of a 3D topological insulator21,22. The underlying physics here is quite different and relies on the topologically protected surface states that are the hallmark of these materials23,24. It is easy to show by an explicit calculation25 that the spectrum of such surface states in a wire whose cross-section is threaded by magnetic flux (n + ½)Φ0 , with n an integer and Φ0 = hc/e the magnetic flux quantum, where h is Planck’s constant, c is the speed of light and e is the electron charge, has the form illustrated in Fig. 3b. It consists of a pair of non-degenerate linearly dispersing gapless modes and a set of doubly degenerate gapped modes. The important property of this spectrum is that the number of Fermi points in the right half of the Brillouin zone is odd for any value of μ as long as it lies inside the bulk bandgap, which is ~300 meV in the Bi2Se3 family of materials. Thus, such a wire conforms to Kitaev’s paradigm and will exhibit Majorana zero modes when superconducting order is induced in it by the proximity effect21. Furthermore, unlike in the semiconductor wires, superconducting order in this set-up is expected to be robust against the effects of non-magnetic disorder22.

As of this writing the existence of coherent surface states in topological insulator wires has been established26 and the superconducting proximity effect has been demonstrated27. However, signatures of Majorana zero modes have not yet been reported. The key difficulty seems to lie in the fact that as in most bulk topological insulators μ in the wires is pinned in the conduction band thus obscuring the universal physics of the surface modes.

Majorana yes or no?The easiest way to experimentally observe the Majorana zero modes in wires is by using tunnelling spectroscopy. The tunnelling conductance dI/dV into the wire’s end should show a superconducting gap with a peak at the zero bias when the Majorana mode is present and no peak when it is absent. Furthermore, under ideal conditions (that is, low temperature and weak disorder), the zero-bias peak conductance should be quantized28 at 2e2/h and the SC gap should close at the phase transition between the topological and non-topological phase. The existing experiments6–9 performed in a set-up similar to Fig. 1 generally observe a ‘soft SC gap’ (that is, reduced but non-zero

Figure 2 | Two phases of the Kitaev chain. Trivial phase (top) has Majorana fermions (blue spheres) bound in pairs located on the same site of the physical lattice, represented by translucent spheres. In the topological phase (bottom) Majorana femions are bound in pairs located on the neighbouring sites leading to two unpaired Majoranas at both ends, represented by the red spheres.

© 2013 Macmillan Publishers Limited. All rights reserved

Page 3: Majorana's wires

NATURE NANOTECHNOLOGY | VOL 8 | MARCH 2013 | www.nature.com/naturenanotechnology 151

commentary

density of states at low bias) with a non-quantized zero-bias peak emerging and then disappearing as the magnetic field is increased.

The zero-bias peak is interpreted as an evidence for the Majorana zero mode. There is typically no sign of SC-gap closing associated with the appearance of the zero-bias peak. This last feature at least can be understood from theoretical models29, which tend to show unambiguous gap closing for tunnelling into the bulk of the wire but, under a wide range of conditions, no apparent gap closing for tunnelling into the wire end. A control experiment showing dI/dV for both the middle and the end of the wire would thus be desirable to resolve this particular issue.

The big question of course remains whether or not the observed non-quantized zero-bias peaks6–9 reflect the Majorana zero mode. Recent theoretical work30 showed that in the presence of disorder such non-quantized zero-bias peaks generically appear even when the wire is in the non-topological phase. These do not correspond to the Majorana modes but instead to ordinary Andreev bound states that exist close to zero energy. To make things more complicated it turns out30 that such Andreev states can have similar dependence on the applied magnetic field as the Majorana modes so the experimentally observed appearance and disappearance of the peaks as a function of B does not really constitute an unambiguous evidence for the Majorana modes, as was originally thought.

Another class of experiments tests the 4π-periodic Josephson effect that is predicted to occur between two SC wires in the topological phase17. The Josephson effect between two ordinary superconductors involves Cooper pair tunnelling and the current shows the characteristic I(φ) = ICsinφ behaviour, which is 2π-periodic in the relative phase variable φ. For two topological superconductors the existence of the Majorana zero modes enables single-electron tunnelling, which introduces a sin(φ/2) component to the tunnelling current. This effect was probed by Rokhinson et al.10 who measured the Shapiro steps that occur when a junction is exposed to a radiofrequency electromagnetic field. The 4π-periodic Josephson component is manifested by the doubling of the Shapiro step height, which was indeed observed when the wires were subjected to a static magnetic field. The 4π-periodic Josephson effect was thought to represent a more reliable signature of Majorana fermions in this context than

the tunnelling spectroscopy. Recently, however, it was shown theoretically31 that the 4π-periodic Josephson effect can actually arise under certain conditions even for Josephson junctions formed of ordinary superconductors with no Majorana zero modes. Thus, the experimental result10 weighs in favour of Majorana interpretation but does not constitute a definitive proof.

The road aheadTheoretical models provide rather unambiguous predictions for the existence of Majorana zero modes in nanowires made of semiconductors with strong spin–orbit coupling and of topological insulators. The experimental data currently available shows features broadly consistent with the expectations for Majorana fermions in semiconductor wires but a number of

puzzles and challenges remain. Among these the most troublesome perhaps is that it has been almost too easy to observe these elusive modes. Given what we know about the purity of the wires, the interfaces and the effect of gating it should have been more difficult to tune μ to lie inside the Zeeman gap that is just a few Kelvins wide. This is illustrated in the density of states plot in Fig. 3a: it shows that if the electron density in the wire is set at random (as would be the case for a wire in direct contact with a superconductor) the chances for μ accidentally landing in the Zeeman gap are rather small.

Nevertheless the data of several different groups6–9 using variants of the basic set-up show the (non-quantized) zero-bias peak that appears in the range of magnetic fields consistent with the Majorana prediction.

E(k)

E(k)

E

E

a

b

k

k

VZ

DOS

Figure 3 | Electron-excitation spectra in semiconductor and topological insulator nanowires. a, The characteristic spectrum as a function of momentum k of a semiconductor wire with Rashba and Zeeman coupling. To achieve the topological phase, μ must lie in the green-shaded region of width VZ. b, The spectrum of a topological insulator wire in a parallel magnetic field. The red dashed lines correspond to non-degenerate bands whereas the blue solid lines are doubly degenerate. The right panels show the corresponding density of states (DOS).

© 2013 Macmillan Publishers Limited. All rights reserved

Page 4: Majorana's wires

152 NATURE NANOTECHNOLOGY | VOL 8 | MARCH 2013 | www.nature.com/naturenanotechnology

commentary

What does this mean? It is possible that for some unknown reason the Majorana zero modes are more stable and occur under a wider range of conditions than theoretically expected, or that we have been lucky in terms of achieving the necessary conditions in the existing devices. It is also possible that the non-quantized zero-bias peaks reflect ordinary Andreev states instead of Majorana zero modes30. Then there is the 4π-periodic Josephson effect10 that now seems to be a necessary but not a unique consequence of the Majorana zero modes31 adding into the evidence.

Where do we go from here? It is now thought that an observation of a stable quantized 2e2/h zero-bias peak would constitute a smoking gun proof of the Majorana zero mode; it is generally held true that mimicking this type of exact quantization would be difficult. Experimentally, such an observation using present day devices would require an unrealistically low temperature but can perhaps be achieved in future devices. Absent a single smoking gun experiment the final proof of the Majorana existence will most likely involve a body of additional experimental and theoretical work aimed at verifying various aspects of the Majorana zero modes. These include the already mentioned gap closing at the phase transition to and from the topological phase, better theoretical understanding of the ‘soft gap’ phenomenon observed using tunnelling spectroscopy, as well as several theoretically proposed tests, argued to be

less ambiguous, but so far unrealized. These include probes of non-locality inherent to the pair of spatially separated Majorana zero modes32–35 and more ambitiously direct tests of their non-Abelian exchange statistics36. Finally, it is likely that in the near future similar experiments will be conducted in topological insulator nanowires. If the signatures of zero modes are observed in these as well then this would add to the evidence in favour of Majorana fermions. Owing to the less restrictive conditions that should prevail in these systems it might be possible to access different regimes and observe, for instance, the quantized zero-bias peak conductance more easily. Given our theoretical understanding of the problem and the mounting experimental evidence it is becoming increasingly difficult to foresee a scenario in which the Majorana zero modes would not underlie at least some of the recent reports. Nevertheless, 75 years after Majorana’s historic prediction, the race for the unambiguous detection of these elusive particles continues. ❐

Marcel Franz is at the Department of Physics and Astronomy, University of British Columbia, Vancouver, British Columbia V6T 1Z1, Canada. e-mail: [email protected]

References1. Majorana, E. Nuovo Cimento 14, 171–184 (1937).2. Wilczek, F. Nature Phys. 5, 614–618 (2009).3. Franz, M. Physics 3, 24 (2010).4. Stern, A. Nature 464, 187–193 (2010).5. Nayak, C. et al. Rev. Mod. Phys. 80, 1083–1159 (2008).

6. Mourik, V., Zuo, K., Frolov, S. M., Plissard, S. R., Bakkers, E. P. A. M. & Kouwenhoven, L. P. Science

336, 1003–1007 (2012).7. Deng, M. T. et al. Nano Lett. 12, 6414–6419 (2012).8. Das, A. et al. Nature Phys. 8, 887–895 (2012).9. Finck, A. D. K., Van Harlingen, D. J., Mohseni, P. K., Jung, K. & Li, X. Preprint at http://arXiv.org/abs/1212.1101 (2012). 10. Rokhinson, L. P., Liu, X. & Furdyna, J. K. Nature Phys.

8, 795–799 (2012).11. Moore, G. & Read, N. Nucl. Phys. B 360, 362–396 (1991).12. Kitaev, A. Ann. Phys. 303, 2–30 (2003).13. Read, N. & Green, D. Phys. Rev. B 61, 10267–10297 (2000).14. Fu, L. & Kane, C. L. Phys. Rev. Lett. 100, 096407 (2008).15. Sau, J. D., Lutchyn, R. M., Tewari, S. & Das Sarma, S. Phys. Rev.

Lett. 104, 040502 (2010).16. Alicea, J. Phys. Rev. B 81, 125318 (2010).17. Kitaev, A. Y. Phys. Usp. 44, 131–136 (2001).18. Lutchyn, R. M., Sau, J. D. & Das Sarma, S. Phys. Rev. Lett.

105, 077001 (2010).19. Oreg, Y., Refael, G. & von Oppen, F. Phys. Rev. Lett.

105, 177002 (2010).20. Alicea, J. Rep. Prog. Phys. 75, 076501 (2012).21. Cook, A. M. & Franz, M. Phys. Rev. B 84, 201105(R) (2011).22. Cook, A. M., Vazifeh, M. M. & Franz, M. Phys. Rev. B

86, 155431 (2012).23. Moore, J. E. Nature 464, 194–198 (2010).24. Hasan, M. Z. & Kane, C. L. Rev. Mod. Phys. 82, 3045–3067 (2010).25. Rosenberg, G., Guo, H-M. & Franz, M. Phys. Rev. B

82, 041104(R) (2010).26. Peng, H. et al. Nature Mater. 9, 225–229 (2010).27. Zhang, D. et al. Phys. Rev. B 84, 165120 (2011).28. Wimmer, M., Akhmerov, A. R., Dahlhaus, J. P. &

Beenakker, C. W. J. New J. Phys. 13, 053016 (2011).29. Stanescu, T. D., Tewari, S., Sau, J. D. & Das Sarma, S. Phys. Rev.

Lett. 109, 266402 (2012).30. Liu, J., Potter, A. C., Law, K. T. & Lee, P. A. Phys. Rev. Lett.

109, 267002 (2012).31. Sau, J. D., Berg, E. & Halperin, B. I. Preprint at http://arXiv.org/

abs/1206.4596 (2012). 32. Fu, L. Phys. Rev. Lett. 104, 056402 (2010).33. Das Sarma, S., Stanescu, T. D. & Sau, J. D. Phys. Rev. B

86, 220506 (2012). 34. Sau, J. D., Swingle, B. & Tewari, S. Preprint at http://arXiv.org/

abs/1210.5514 (2012). 35. Liu, J., Zhang, F-C. & Law, K. T. Preprint at http://arXiv.org/

abs/1212.5879 (2012). 36. Alicea, J., Oreg, Y., Refael, G., von Oppen, F. & Fisher, M. P. A.

Nature Phys. 7, 412–417 (2011).

Skyrmions on the trackAlbert Fert, Vincent Cros and João Sampaio

Magnetic skyrmions are nanoscale spin configurations that hold promise as information carriers in ultradense memory and logic devices owing to the extremely low spin-polarized currents needed to move them.

Skyrmions are named after the nuclear physicist Tony Skyrme, who developed a nonlinear field theory for interacting

pions in the 1960s and showed that topologically stable field configurations occur as particle-like solutions1. The word skyrmion is now used to denote similar mathematical objects in many different contexts, from elementary particles to

liquid crystals, Bose–Einstein condensates and quantum Hall magnets. Magnetic skyrmions are chiral spin structures with a whirling configuration (Fig. 1a–d). As their structure cannot be continuously deformed to a ferromagnetic or other magnetic state, skyrmions are topologically protected and relatively stable structures, in comparison with, for example, magnetic

vortices or bubbles. These spin textures were first observed in 20092–4, and have since been intensively investigated because they reveal novel types of chiral magnetic orders induced by the spin–orbit coupling (SOC). Moreover, they are appealing because of their potential applications in novel spintronic devices, for example, information storage or logic devices based

© 2013 Macmillan Publishers Limited. All rights reserved


Recommended