arXiv:2006.06651v1 [cond-mat.mes-hall] 11 Jun 2020

9
Creating Majorana modes from segmented Fermi surface Micha l Papaj and Liang Fu Department of Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA We present a new platform for creating Majorana bound states from 2D gapless superconducting state in spin-helical systems under the in-plane Zeeman field. Topological 1D channels are formed by quantum confinement of quasiparticles via Andreev reflection from the surrounding fully gapped superconducting region. Our proposal can be realized using narrow strips of magnetic insulators on top of proximitized 3D topological insulators. This setup has key advantages that include: small Zeeman fields, no required fine-tuning of chemical potential, removal of the low-energy detrimental states, and large attainable topological gap. Introduction.— Majorana bound states [1, 2] have en- grossed the condensed matter physics community for much of the last decade, offering promises of fascinat- ing phenomena of fundamental interest and potential ap- plications in topologically protected quantum computing [3, 4]. Over the years, multiple platforms of very di- verse variety have been proposed as hosts of Majorana bound states and intensively studied, including proximi- tized topological insulators [5], heavy metal surfaces [6], semiconductor nanowires [7–9], magnetic atom chains [10, 11], planar Josephson junctions in 2D electron gas [12] and iron superconductors [13–15]. Signatures of Ma- jorana bound states in all these platforms have been ob- served in multiple experiments [16–29]. Despite the remarkable progress on many frontiers, most of the existing material platforms have some dis- advantages that hamper their further investigation and prompt the continued search [30–32] for new systems that would resolve these issues. The multitude of issues plagu- ing current Majorana platforms are related to the mate- rial quality, the device fabrication process and the re- quired experimental conditions. For example, in iron su- perconductor FeSeTe the topological band inversion nec- essary for creating Majorana bound states requires alloy- ing, which results in disorder and inhomogeneity [20]. In the fabrication of semiconductor nanowires band bending near the crystal-vacuum interface may result in quasi- Majorana bound states that complicate the interpreta- tion of the zero bias peak [33, 34]. Moreover, in many platforms the appearance of Majorana bound states re- lies on strong external magnetic fields above 1 T (which is detrimental to superconductivity) or fine-tuning the chemical potential into the Zeeman gap. In this Letter we propose a new approach to creating Majorana bound states in 2D systems that can resolve these issues. Our proposal is based on a gapless super- conducting state of spin-helical electrons placed under an in-plane Zeeman field. Examples of such systems are presented in Fig. 1 and include 3D topological insula- tors (TI) in proximity to conventional superconductors [35–42] or superconductors with Rashba spin-orbit cou- pling. In such systems the interplay of superconductivity and Zeeman field leads to a ”segmented Fermi surface”: FIG. 1. Platforms for creating Majorana bound states from segmented Fermi surface, based on proximitized 3D topolog- ical insulator (such as Bi2Se3). Quasi-1D channel is formed due to: (a) narrow strip of magnetic insulator (such as EuS) with magnetization parallel to the strip axis (b) narrow region (yellow) with a smaller superconducting gap under external in-plane magnetic field. while a large part of the normal state Fermi surface is still gapped, its remainder is reconstructed into contours consisting of electron- and hole-dominated arcs [43]. A topological gap can be opened in a quantum confined quasi-1D channel of such a gapless superconducting state surrounded by regions with a full superconducting gap. Majorana bound states will emerge at the boundaries of thus formed 1D channel. As a concrete example, we focus on the setup of Fig. 1(a), where we propose to use narrow strips of mag- netic insulator such as EuS on top of a proximitized thin film of 3D TI such as Bi 2 Se 3 . Crucially, since the narrow strip region is surrounded by a superconducting (and not just insulating) gap, the number of low energy modes de- pends on the strip width W and the superconducting co- herence length ξ , independent of the chemical potential. As the Zeeman field is induced by exchange interaction with a magnetic insulator strip, it doesn’t impact the parent superconductor or destroy the proximity effect in the rest of the surface. Since our proposal relies on seg- mented (rather than full) Fermi surface, it results in the removal of the low energy states. This translates to large topological gaps that are crucial for topological quantum computing applications. Combination of all these fea- tures makes our proposal an attractive alternative to the existing systems. Model.— While our proposal based on segmented arXiv:2006.06651v1 [cond-mat.mes-hall] 11 Jun 2020

Transcript of arXiv:2006.06651v1 [cond-mat.mes-hall] 11 Jun 2020

Creating Majorana modes from segmented Fermi surface

Micha l Papaj and Liang FuDepartment of Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA

We present a new platform for creating Majorana bound states from 2D gapless superconductingstate in spin-helical systems under the in-plane Zeeman field. Topological 1D channels are formedby quantum confinement of quasiparticles via Andreev reflection from the surrounding fully gappedsuperconducting region. Our proposal can be realized using narrow strips of magnetic insulators ontop of proximitized 3D topological insulators. This setup has key advantages that include: smallZeeman fields, no required fine-tuning of chemical potential, removal of the low-energy detrimentalstates, and large attainable topological gap.

Introduction.— Majorana bound states [1, 2] have en-grossed the condensed matter physics community formuch of the last decade, offering promises of fascinat-ing phenomena of fundamental interest and potential ap-plications in topologically protected quantum computing[3, 4]. Over the years, multiple platforms of very di-verse variety have been proposed as hosts of Majoranabound states and intensively studied, including proximi-tized topological insulators [5], heavy metal surfaces [6],semiconductor nanowires [7–9], magnetic atom chains[10, 11], planar Josephson junctions in 2D electron gas[12] and iron superconductors [13–15]. Signatures of Ma-jorana bound states in all these platforms have been ob-served in multiple experiments [16–29].

Despite the remarkable progress on many frontiers,most of the existing material platforms have some dis-advantages that hamper their further investigation andprompt the continued search [30–32] for new systems thatwould resolve these issues. The multitude of issues plagu-ing current Majorana platforms are related to the mate-rial quality, the device fabrication process and the re-quired experimental conditions. For example, in iron su-perconductor FeSeTe the topological band inversion nec-essary for creating Majorana bound states requires alloy-ing, which results in disorder and inhomogeneity [20]. Inthe fabrication of semiconductor nanowires band bendingnear the crystal-vacuum interface may result in quasi-Majorana bound states that complicate the interpreta-tion of the zero bias peak [33, 34]. Moreover, in manyplatforms the appearance of Majorana bound states re-lies on strong external magnetic fields above 1 T (whichis detrimental to superconductivity) or fine-tuning thechemical potential into the Zeeman gap.

In this Letter we propose a new approach to creatingMajorana bound states in 2D systems that can resolvethese issues. Our proposal is based on a gapless super-conducting state of spin-helical electrons placed underan in-plane Zeeman field. Examples of such systems arepresented in Fig. 1 and include 3D topological insula-tors (TI) in proximity to conventional superconductors[35–42] or superconductors with Rashba spin-orbit cou-pling. In such systems the interplay of superconductivityand Zeeman field leads to a ”segmented Fermi surface”:

FIG. 1. Platforms for creating Majorana bound states fromsegmented Fermi surface, based on proximitized 3D topolog-ical insulator (such as Bi2Se3). Quasi-1D channel is formeddue to: (a) narrow strip of magnetic insulator (such as EuS)with magnetization parallel to the strip axis (b) narrow region(yellow) with a smaller superconducting gap under externalin-plane magnetic field.

while a large part of the normal state Fermi surface isstill gapped, its remainder is reconstructed into contoursconsisting of electron- and hole-dominated arcs [43]. Atopological gap can be opened in a quantum confinedquasi-1D channel of such a gapless superconducting statesurrounded by regions with a full superconducting gap.Majorana bound states will emerge at the boundaries ofthus formed 1D channel.

As a concrete example, we focus on the setup ofFig. 1(a), where we propose to use narrow strips of mag-netic insulator such as EuS on top of a proximitized thinfilm of 3D TI such as Bi2Se3. Crucially, since the narrowstrip region is surrounded by a superconducting (and notjust insulating) gap, the number of low energy modes de-pends on the strip width W and the superconducting co-herence length ξ, independent of the chemical potential.As the Zeeman field is induced by exchange interactionwith a magnetic insulator strip, it doesn’t impact theparent superconductor or destroy the proximity effect inthe rest of the surface. Since our proposal relies on seg-mented (rather than full) Fermi surface, it results in theremoval of the low energy states. This translates to largetopological gaps that are crucial for topological quantumcomputing applications. Combination of all these fea-tures makes our proposal an attractive alternative to theexisting systems.

Model.— While our proposal based on segmented

arX

iv:2

006.

0665

1v1

[co

nd-m

at.m

es-h

all]

11

Jun

2020

2

Fermi surfaces is rather versatile, for concreteness westart our analysis with a thin film of a 3D topologicalinsulator (TI) in proximity to an ordinary s-wave super-conductor [5], with a narrow strip of magnetic insulatordeposited on top as shown in Fig. 1(a). The TI surfacestates acquire the superconducting gap ∆ everywhere onthe surface and are subject to an exchange field onlyin the region beneath the magnetic insulator. This sys-tem is described by the following Bogoliubov-de GennesHamiltonian:

H = v(kxτzsy − kyτzsx)− µτz +Bx(x, y)sx + ∆τx, (1)

where τi and si are Pauli matrices describing the particle-hole and spin degrees of freedom, respectively, v is theFermi velocity of the surface state and µ is the chemi-cal potential. In our discussion we focus on the case ofexchange field parallel to the strip, which results in theZeeman energy Bx.

We first want to analyze the eigenstates of this Hamil-tonian for translationally invariant cases and then con-struct the solutions that are quantum confined within thefinite width strip. To simplify the analysis of the problemwe note that in the experimentally relevant scenarios µ isthe largest energy scale of the problem (µ� Bx,∆) andso we can concentrate only on the upper Dirac cone nearthe Fermi level and disregard the other band (assumingµ > 0). After the projection we obtain the following lowenergy Hamiltonian:

Hp =

(kv − µ−Bxky/k ∆

∆ −kv + µ−Bxky/k

)(2)

where k =√k2x + k2

y and we have neglected the off-

diagonal momentum dependent terms as they are sup-pressed by the factors of ∆/µ, Bx/µ.

This effective Hamiltonian clearly demonstrates thatdue to the spin-momentum locking of Dirac surface statesthe Zeeman field Bx causes a Fermi surface shift in the kydirection. This results in a direction-dependent depairingeffect on the superconducting state, which is maximumat ky = ±kF and zero for kx = ±kF . The eigenvaluesof Hamiltonian (2) at kx = 0 are presented in Fig. 2(a)for two values of magnetic field, Bx = 0 and Bx = 1.5∆.While in the absence of Zeeman energy there are no statesinside of the gap, when Bx > ∆ the gap closes and zeroenergy states of Bogoliubov quasiparticles are located intwo closed contours near ky = ±kF . Each contour iscomposed of electron- and hole-dominated arcs, whilethe remainder of the initial normal state Fermi surfaceis still gapped around kx = ±kF [43]. Such a segmentedFermi surface is presented in Fig. 2(b). The contour atε = 0 can be described by the ky wavevector componentsparametrized by kx, which are approximately given by:

ke/h,± = ±k0 + s

√B2x/v

2 − ∆2

v2

k2F

k20

(3)

FIG. 2. (a) Bands of the projected Hamiltonian (2) at kx = 0for Bx = 0 (dashed line) and Bx = 1.5∆ (solid line). (b)Segmented Fermi surface at ε = 0 for Bx = 2∆ with electron(solid line) and hole (dashed line) arcs. The arrow indicatesthe Andreev reflection process at kx = 0.

with k0 = kF√

1− k2x/k

2F and s = ±1 for electron- and

hole-dominated states. Such gapless superconductingstate has been recently observed in quasiparticle inter-ference experiments on proximitized Bi2Te3 in in-planemagnetic field [44], with the magnitude of few tens ofmT due to a large enhancement of the g-factor of theDirac surface state.

To open up the topological gap in a region with a seg-mented Fermi surface, we use the quantum confinementeffects by limiting the transverse size of the region withthe non-zero Zeeman energy. In such a case, by surround-ing the magnetic strip by a fully gapped superconduct-ing region with zero Zeeman energy we effectively forma 1D channel with the number of quasiparticle modesdetermined by the width of the strip. The use of thesuperconducting gap to confine Bogoliubiov quasiparti-cles distinguishes our proposal from the previous schemesbased on nanowires, which use vacuum to confine elec-trons [7, 8]. The idea of creating 1D topological channelsby confinement using superconducting gap has been ex-plored in other setups [5, 12, 31].

Topological phase diagram from scattering approach.—To investigate the topological properties of this system,we first focus on the quasiparticle spectrum in a stripof finite width W in y direction and infinite length in xdirection with no disorder inside of the wire. In such acase, kx remains a good quantum number and we canobtain the in-gap states spectrum by solving a scatter-ing problem along y axis with states under the magneticstrip normalized to carry a unit quasiparticle current.For energies |ε| < ∆ there are no propagating states out-side of the magnetic strip region and so at the interfacesat y = ±W/2 normal and Andreev reflection can occurwith no transmission into the surrounding superconduc-tor. Andreev reflection in proximitized surface states of3D TI has been intensively studied experimentally [45–47]. Inside the energy window where bands for oppositeky overlap we have electron- and hole-dominated states

3

moving in positive and negative y direction with ky givenby Eq. (3).

As the gap in the continuum model first closes atkx = 0 as Zeeman energy is increased, to determine thecondition for subband inversion and thus establish theboundaries of the topological phases in the Bx −W pa-rameter space it is enough to consider the states withno longitudinal momentum. Since kx is conserved in thescattering processes, this means that for such states nor-mal reflection at the strip boundaries is forbidden dueto the spin-momentum locking of Dirac surface states(states on the opposite sides of the Fermi surface are or-thogonal). Together with the unitarity condition for thescattering matrix this means that Andreev reflection atkx = 0 and ε = 0 can be characterized by a single phaseφA acquired by the particles during the reflection. There-fore, the scattering matrix S±W/2 at the two interfacesgiven by:

SW/2 =

(0 eiφA

e−iφA 0

), S−W/2 = S∗W/2 (4)

Since the particles propagate freely between An-dreev reflections at opposite interfaces, they acquirethe phase determined by their wavevectors. Thistranslates to transmission matrices for movementalong the positive and negative y directions T± =diag(exp(ike,±W ), exp(ikh,∓W )). To find the boundstate spectrum we use the condition [48]:

det(1− T−SW/2T+S−W/2

)= 0 (5)

Since we evaluate this condition for ε = 0 and kx = 0, itgreatly simplifies to the form:

1− cos(2∆kW − 2φA) = 0 (6)

where ∆k = (ke,+ − kh,+)/2 =√B2x −∆2/v is the

wavevector difference between the electron and hole-likestates propagating in the same direction. This allows usto derive the topological phase boundaries in the Bx−Wspace to be:

W/ξ =φA + πn√B2x − 1

, n ∈ N (7)

with superconducting coherence length ξ = v/∆, Bx =Bx/∆ and Andreev reflection phase determined from themicroscopic considerations:

φA = Arg

(i− exp(−arcoshBx)

−i+ exp(arcoshBx)

)(8)

We can also determine the quasiparticle spectrum atkx = 0 in the vicinity of the phase boundaries. To dothis, we solve Eq. (5) for ε. We expand the solution close

FIG. 3. (a) Subgap states in a narrow strip at Bx = 2.6∆and W/ξ = 1.16 without (blue line) and with (orange line)superconductivity under the magnetic strip. (b) Bound stateenergies at kx = 0 for increasing Zeeman energy. Solid orangelines indicate subsequent analytical solutions of Eq. (9). (c)The topological phase diagram of the system. Dashed linesindicate the phase boundaries described by Eq. (7).

to Bc,n, which is the Zeeman energy at which the gap atkx = 0 closes at the nth boundary. In doing so we get:

ε± = ±∆B2c,nW + v∆

B2c,n(v +W∆)

(Bx −Bc,n) (9)

where ± gives the two particle-hole symmetric eigenval-ues.

As pointed out by Kitaev [1], the topological phase ofa 1D superconductor is determined by the gap closing atkx = 0. This is exactly described by Eq. (7), as it de-termines the conditions for the subsequent quasiparticlebranches to cross ε = 0 and become inverted. Therefore,crossing the boundaries with even n marks the transitionfrom the trivial to the topological regime, and crossingodd n curves marks the opposite transition. We will befocusing on the first topological region between the n = 0and n = 1 boundaries.

It is worth highlighting the two key advantages of ourproposed setup. First, the phase boundaries and thenumber of subgap quasiparticle modes are independentof the chemical potential and instead rely only on theW/ξ and Bx/∆ ratios. Remarkably, this remains true

4

even in the presence of Fermi wavelength mismatch in-side and outside of the strip. This is a result of the spin-helical nature of the TI surface state that forbids normalreflection at kx = 0, leaving Andreev reflection as theonly confinement mechanism at kx = 0. As a conse-quence, even for a large chemical potential there is onlyone subgap mode as long as W ∼ ξ. This is in con-trast to semiconductor nanowires, which can have manylow energy subbands that greatly complicate the phasediagram, introducing numerous topological phase transi-tions. Secondly, since a large portion of the original Fermisurface remains gapped, there will be no detrimental lowenergy electrons moving parallel to the channel. Thesestates, unaffected by the normal and Andreev reflectionprocesses, would be present if the narrow strip region wasin normal state. With such states out of the picture, themaximum topological gap at given Bx will be determinedby the energy of states at kx = 0. Therefore, the upperbound on the topological gap is given by the crossingpoints of the subsequent branches of Eq.(9). This consti-tutes a strong enhancement of the achievable topologicalgap when compared to the platforms based on Joseph-son junctions with normal state weak link. Both of thedescribed features result from confining the quasiparti-cles with segmented Fermi surface by Andreev reflection,rather than electrons with full Fermi surface by normalreflection, to a 1D channel.

Comparison with numerical simulation.— We can nowcompare the approximate analytical solutions to thenumerical calculations based on a tight-binding modelwith translational invariance in x direction and periodicboundary conditions in y direction. The simulations wereperformed using the Kwant code [49] (details in the Sup-plementary Materials). First, we illustrate the differencebetween our proposal and a scenario in which there isno superconducting pairing under the magnetic strip. InFig. 3(a) we present the numerical spectrum of 1D sub-bands without and with superconductivity, with the samepotential barrier placed along the strip at its center tointroduce normal reflection. In both cases we observeinverted quasiparticle branches inside of the supercon-ducting gap. However, when superconductivity is absentin the weak link, there are low energy states with largekx (moving parallel to the strip) that will interfere withthe observation of Majorana modes. On the contrary,with superconductivity present under the magnetic strip,there are no low energy states at large kx and the topo-logical gap is several times larger with the same barrierstrength. We can thus further investigate the kx = 0eigenvalues numerically to characterize the upper boundon the topological gap. We present these eigenvaluesin Fig. 3(b), which shows the subsequent branch cross-ings. We note that the formula of Eq. (9) provides a verygood approximation of the energies for given n up untilit crosses with the n+ 1 branch. We therefore highlightthat the maximum upper bound is given by such crossing

FIG. 4. (a) Density of states of the system with W/ξ = 1.16(b) Local density of states at E = 0 for Bx = 2.6∆.

energies and this value can be larger than 60% of the orig-inal superconducting gap, which constitutes a significantadvantage of our proposal. Finally, we investigate thefull phase diagram of the system with normal reflectionin Fig. 3(c). The phase boundaries given by Eq. (7) are invery good agreement with the tight-binding calculationand the first topological region between curves n = 0 andn = 1 covers a wide area both in terms of the Zeeman en-ergy Bx and the strip width W , avoiding the necessity ofparameter fine-tuning. We also note that the topologicalgap is a significant fraction the original superconductinggap in majority of the first topological region. In eachfollowing region the gap becomes smaller, but as we wantto minimize the Zeeman energy necessary to obtain thetopological phase this is not a concern. In general, thewider the strip, the smaller the Zeeman energy requiredto invert the branches. However, at the same time thepossible size of the topological gap at the same normalreflection barrier strength decreases. Therefore, for thepurpose of the experiment it will be necessary to find asweet spot for a particular realization that maximizes thegap, based on a more precise modeling.

To further verify the topological character of the sys-tem, we perform numerical simulations of a finite lengthstrip to demonstrate that it hosts Majorana bound statesat its ends. First, we perform exact diagonalization ofthe 2D tight-binding Hamiltonian and plot the obtainedeigenvalues in the form of density of states of the sys-tem in Fig. 4(a). We observe, in accordance with theanalysis of the phase diagram, that for Zeeman energiesbelow ∼ 2.1∆ the system is gapped and there are no zeroenergy states. However, at Bc,0 ≈ 2.1∆ the gap closesand then reopens due to normal reflection processes with

5

a pair of zero energy states present in the system. Thetopological gap increases to above 0.25∆ for this partic-ular scattering strength (not saturating the upper boundof kx = 0) and then closes again at Bc,1 when the secondpair of subbands is inverted. The gap reopens again, butthis time we enter the trivial regime with no zero energybound state as anticipated from the phase diagram. Fi-nally, we plot the pair of zero energy states at Bx = 2.6∆,which are strongly localized at the ends of the strip asexpected for the Majorana bound states.

Additional discussion and summary.— In the preced-ing sections we have shown that using a narrow strip of amagnetic insulator, such as EuS, on top of a 3D topologi-cal insulator in proximity to a conventional superconduc-tor can yield Majorana bound states with a large topolog-ical gap. However, our approach is not limited to Zeemanfield induced by an adjacent magnetic insulator. Alterna-tively, if one applies external in-plane magnetic field, thesame type of gap inversion will occur. If the proximity-induced superconducting gap is non-uniform, e.g. witha narrow region of diminished magnitude ∆′ < ∆, asshown in Fig. 1(b), then the gap inversion will occurin that area, but not in the surrounding superconduc-tor, effectively realizing the scenario we discussed in ourwork. Another possible system are the superconductorswith Rashba spin-orbit coupling that have two indepen-dent pockets of spin-helical electrons (such as proximi-tized InAs 2D electron gas), resulting in two superim-posed topological phase diagrams. However, due to thedifferent velocities and superconducting gaps in each ofthe pockets, the superconducting coherence lengths willbe different in each case, thus enabling the existence ofa region with a single Majorana at each end. Such ascenario may help to understand the topological phasediagram of proximitized Rashba states in gold nanowiresfound in a recent study [28, 31]. This largely expandsthe spectrum of material platforms for realization of ourproposal, which is a testimony to its great versatility.

To sum up, in this work we proposed a new platform forcreating Majorana bound states using proximitized Diracsurface states with in-plane Zeeman energy. Among theadvantages of our proposal are: very small magnitudeof required magnetic fields, removal of the spurious lowenergy states and large possible topological gaps. To-gether with the continuous progress in fabrication of thin3D TI films coupled to superconductors this makes ourproposal an attractive platform for studying Majoranabound states.

Acknowledgments.— This work was supported by DOEOffice of Basic Energy Sciences, Division of Materials Sci-ences and Engineering under Award de-sc0019275. LFwas supported in part by a Simons Investigator Awardfrom the Simons Foundation.

[1] A. Y. Kitaev, Phys.-Usp. 44, 131 (2001).[2] N. Read and D. Green, Phys. Rev. B 61, 10267 (2000).[3] C. Beenakker, Annu. Rev. Condens. Matter Phys. 4, 113

(2013).[4] S. D. Sarma, M. Freedman, and C. Nayak, npj Quantum

Information 1, 1 (2015).[5] L. Fu and C. L. Kane, Phys. Rev. Lett. 100, 096407

(2008).[6] A. C. Potter and P. A. Lee, Phys. Rev. B 85, 094516

(2012).[7] R. M. Lutchyn, J. D. Sau, and S. Das Sarma, Phys. Rev.

Lett. 105, 077001 (2010).[8] Y. Oreg, G. Refael, and F. von Oppen, Phys. Rev. Lett.

105, 177002 (2010).[9] R. M. Lutchyn, E. P. a. M. Bakkers, L. P. Kouwenhoven,

P. Krogstrup, C. M. Marcus, and Y. Oreg, Nature Re-views Materials 3, 52 (2018).

[10] S. Nadj-Perge, I. K. Drozdov, B. A. Bernevig, andA. Yazdani, Phys. Rev. B 88, 020407 (2013).

[11] F. Pientka, L. I. Glazman, and F. von Oppen, Phys.Rev. B 88, 155420 (2013).

[12] F. Pientka, A. Keselman, E. Berg, A. Yacoby, A. Stern,and B. I. Halperin, Phys. Rev. X 7, 021032 (2017).

[13] Z. Wang, P. Zhang, G. Xu, L. K. Zeng, H. Miao, X. Xu,T. Qian, H. Weng, P. Richard, A. V. Fedorov, H. Ding,X. Dai, and Z. Fang, Phys. Rev. B 92, 115119 (2015).

[14] X. Wu, S. Qin, Y. Liang, H. Fan, and J. Hu, Phys. Rev.B 93, 115129 (2016).

[15] G. Xu, B. Lian, P. Tang, X.-L. Qi, and S.-C. Zhang,Phys. Rev. Lett. 117, 047001 (2016).

[16] S. Nadj-Perge, I. K. Drozdov, J. Li, H. Chen, S. Jeon,J. Seo, A. H. MacDonald, B. A. Bernevig, and A. Yaz-dani, Science 346, 602 (2014).

[17] D. Wang, L. Kong, P. Fan, H. Chen, S. Zhu, W. Liu,L. Cao, Y. Sun, S. Du, J. Schneeloch, R. Zhong, G. Gu,L. Fu, H. Ding, and H.-J. Gao, Science 362, 333 (2018).

[18] L. Kong, S. Zhu, M. Papaj, H. Chen, L. Cao, H. Isobe,Y. Xing, W. Liu, D. Wang, P. Fan, Y. Sun, S. Du,J. Schneeloch, R. Zhong, G. Gu, L. Fu, H.-J. Gao, andH. Ding, Nature Physics 15, 1181 (2019).

[19] S. Zhu, L. Kong, L. Cao, H. Chen, M. Papaj, S. Du,Y. Xing, W. Liu, D. Wang, C. Shen, F. Yang, J. Schnee-loch, R. Zhong, G. Gu, L. Fu, Y.-Y. Zhang, H. Ding, andH.-J. Gao, Science 367, 189 (2020).

[20] T. Machida, Y. Sun, S. Pyon, S. Takeda, Y. Kohsaka,T. Hanaguri, T. Sasagawa, and T. Tamegai, Nature Ma-terials 18, 811 (2019).

[21] V. Mourik, K. Zuo, S. M. Frolov, S. R. Plissard, E. P.a. M. Bakkers, and L. P. Kouwenhoven, Science 336,1003 (2012).

[22] A. Das, Y. Ronen, Y. Most, Y. Oreg, M. Heiblum, andH. Shtrikman, Nature Physics 8, 887 (2012).

[23] S. M. Albrecht, A. P. Higginbotham, M. Madsen,F. Kuemmeth, T. S. Jespersen, J. Nyga rd, P. Krogstrup,and C. M. Marcus, Nature 531, 206 (2016).

[24] H. Ren, F. Pientka, S. Hart, A. T. Pierce, M. Kosowsky,L. Lunczer, R. Schlereth, B. Scharf, E. M. Hankiewicz,L. W. Molenkamp, B. I. Halperin, and A. Yacoby, Nature569, 93 (2019).

[25] A. Fornieri, A. M. Whiticar, F. Setiawan, E. Por-toles, A. r. C. C. Drachmann, A. Keselman, S. Gronin,

6

C. Thomas, T. Wang, R. Kallaher, G. C. Gardner,E. Berg, M. J. Manfra, A. Stern, C. M. Marcus, andF. Nichele, Nature 569, 89 (2019).

[26] W. Liu, L. Cao, S. Zhu, L. Kong, G. Wang, M. Pa-paj, P. Zhang, Y. Liu, H. Chen, G. Li, F. Yang,T. Kondo, S. Du, G. Cao, S. Shin, L. Fu, Z. Yin, H.-J.Gao, and H. Ding, arXiv:1907.00904 [cond-mat] (2019),arXiv:1907.00904 [cond-mat].

[27] Z. Wang, J. O. Rodriguez, L. Jiao, S. Howard, M. Gra-ham, G. D. Gu, T. L. Hughes, D. K. Morr, and V. Mad-havan, Science 367, 104 (2020).

[28] S. Manna, P. Wei, Y. Xie, K. T. Law, P. A. Lee, andJ. S. Moodera, PNAS 117, 8775 (2020).

[29] S. Vaitiekenas, Y. Liu, P. Krogstrup, and C. M. Marcus,arXiv:2004.02226 [cond-mat] (2020), arXiv:2004.02226[cond-mat].

[30] T. Laeven, B. Nijholt, M. Wimmer, and A. R.Akhmerov, arXiv:1903.06168 [cond-mat] (2019),arXiv:1903.06168 [cond-mat].

[31] Y.-M. Xie, K. T. Law, and P. A. Lee, arXiv:2003.07052[cond-mat] (2020), arXiv:2003.07052 [cond-mat].

[32] F. Schulz, K. Plekhanov, D. Loss, and J. Klinovaja,arXiv:2004.10623 [cond-mat, physics:quant-ph] (2020),arXiv:2004.10623 [cond-mat, physics:quant-ph].

[33] A. Vuik, B. Nijholt, A. Akhmerov, and M. Wimmer,SciPost Physics 7, 061 (2019).

[34] E. Prada, P. San-Jose, M. W. A. de Moor, A. Geresdi,E. J. H. Lee, J. Klinovaja, D. Loss, J. Nyga rd,R. Aguado, and L. P. Kouwenhoven, arXiv:1911.04512[cond-mat] (2020), arXiv:1911.04512 [cond-mat].

[35] M.-X. Wang, C. Liu, J.-P. Xu, F. Yang, L. Miao, M.-Y.Yao, C. L. Gao, C. Shen, X. Ma, X. Chen, Z.-A. Xu,Y. Liu, S.-C. Zhang, D. Qian, J.-F. Jia, and Q.-K. Xue,Science 336, 52 (2012).

[36] F. Yang, F. Qu, J. Shen, Y. Ding, J. Chen, Z. Ji, G. Liu,J. Fan, C. Yang, L. Fu, and L. Lu, Phys. Rev. B 86,134504 (2012).

[37] J. Wiedenmann, E. Liebhaber, J. Kubert, E. Bocquillon,P. Burset, C. Ames, H. Buhmann, T. M. Klapwijk, andL. W. Molenkamp, Phys. Rev. B 96, 165302 (2017).

[38] A. Q. Chen, M. J. Park, S. T. Gill, Y. Xiao, D. Reig-i-Plessis, G. J. MacDougall, M. J. Gilbert, and N. Mason,Nature Communications 9, 3478 (2018).

[39] S. Ghatak, O. Breunig, F. Yang, Z. Wang, A. A. Taskin,and Y. Ando, Nano Lett. 18, 5124 (2018).

[40] C. X. Trang, N. Shimamura, K. Nakayama, S. Souma,K. Sugawara, I. Watanabe, K. Yamauchi, T. Oguchi,K. Segawa, T. Takahashi, Y. Ando, and T. Sato, Na-ture Communications 11, 159 (2020).

[41] Q. L. He, L. Pan, A. L. Stern, E. C. Burks, X. Che,G. Yin, J. Wang, B. Lian, Q. Zhou, E. S. Choi, K. Mu-rata, X. Kou, Z. Chen, T. Nie, Q. Shao, Y. Fan, S.-C.Zhang, K. Liu, J. Xia, and K. L. Wang, Science 357,294 (2017).

[42] M. Kayyalha, D. Xiao, R. Zhang, J. Shin, J. Jiang,F. Wang, Y.-F. Zhao, R. Xiao, L. Zhang, K. M. Fi-jalkowski, P. Mandal, M. Winnerlein, C. Gould, Q. Li,L. W. Molenkamp, M. H. W. Chan, N. Samarth, andC.-Z. Chang, Science 367, 64 (2020).

[43] N. F. Q. Yuan and L. Fu, Phys. Rev. B 97, 115139 (2018).[44] Z. Zhu, M. Papaj, X.-A. Nie, H.-K. Xu, Y.-S. Gu,

X. Yang, D. Guan, S. Wang, Y. Li, C. Liu, Z.-A. Xu,H. Zheng, L. Fu, and J.-F. Jia, In preparation.

[45] A. D. K. Finck, C. Kurter, Y. S. Hor, and D. J. Van Har-

lingen, Phys. Rev. X 4, 041022 (2014).[46] E. Tikhonov, D. Shovkun, M. Snelder, M. Stehno,

Y. Huang, M. Golden, A. Golubov, A. Brinkman, andV. Khrapai, Phys. Rev. Lett. 117, 147001 (2016).

[47] L. A. Jauregui, M. Kayyalha, A. Kazakov, I. Miotkowski,L. P. Rokhinson, and Y. P. Chen, Appl. Phys. Lett. 112,093105 (2018).

[48] C. W. J. Beenakker, Phys. Rev. Lett. 67, 3836 (1991).[49] C. W. Groth, M. Wimmer, A. R. Akhmerov, and

X. Waintal, New J. Phys. 16, 063065 (2014).

1

Supplementary Material for ”Creating Majorana modes from segmented Fermisurface”

DERIVATION OF THE PROJECTED HAMILTONIAN

To simplify the analysis of the problem of a proximitized Dirac surface state in in-plane magnetic field, we firstnotice that the chemical potential µ� ∆, Bx is the largest energy scale of the problem. This allows us to project theoriginal Hamiltonian (1) of the main text to just the electron and hole component of the band closest to Fermi energy.To do so, we first diagonalize the problem in the absence of superconductivity (∆ = 0), which yields the followingeigenenergies and spinors:

E1,± = ±|kxv + i(Bx − kyv)| − µ, ψ1,± =1√2

(∓ieiα− , 1, 0, 0)T (S1)

E2,± = ±|kxv + i(Bx + kyv)|+ µ, ψ2,± =1√2

(±ie−iα+ , 1, 0, 0)T (S2)

where we define α± = Arg (kxv + i(Bx ± kyv)). We can now express the full Hamiltonian with superconductivityfrom Eq. (1) of the main text in the basis of these four states. However, for large µ in each of the pairs of statesabove there is only one (ψ1,+ and ψ2,−) whose energy is close to µ and thus is relevant to the low energy properties ofthe system. Moreover, these states are only weakly coupled to the remaining pair by a term of magnitude ∼ ∆Bx/µ.Therefore, we can neglect the remaining states and focus only on the two low energy states, obtaining in this way theprojected Hamiltonian (2) of the main text:

Hp =

(kv − µ−Bxky/k ∆

∆ −kv + µ−Bxky/k

)(S3)

with the following eigenvalues:

E± = ±√

(kv − µ)2 + ∆2 −Bxkyk

(S4)

These eigenvalues are presented in Fig. 2(a) of the main text for two values of magnetic field. When Bx > ∆, thezero energy contours for these bands form the segmented Fermi surface as shown in Fig. 2(b) of the main text.

SCATTERING MATRIX APPROACH CALCULATION

Using the low energy Hamiltonian we can now obtain the subgap quasiparticle modes underneath the magneticstrip. We consider a strip infinite in the x direction with translational invariance, which allows us to parametrizethe states by their energy and longitudinal momentum kx. Since the projected basis is momentum-dependent, it isuseful to transform it back to the original basis to solve the scattering problem in the y direction. When Bx > ∆,there are four states on the segmented Fermi surface for given kx, two electron-dominated and two hole-dominated,that are moving in the opposite directions along the y axis. These states, expressed in the initial basis of the originalHamiltonian, are approximately given by:

ψe,+ =1

2

(−ieiα−−iβ+ , e−iβ+ ,−ie−iα+ , 1

)Teikxx+ike,+y (S5a)

ψe,− =1

2

(−ieiα++iβ− , eiβ− ,−ie−iα− , 1

)Teikxx+ike,−y (S5b)

ψh,+ =1

2

(−ieiα−+iβ+ , eiβ+ ,−ie−iα+ , 1

)Teikxx+ikh,+y (S5c)

ψh,− =1

2

(−ieiα+−iβ− , e−iβ− ,−ie−iα− , 1

)Teikxx+ikh,−y (S5d)

with β± = −arccos ((ε±Bxk0v/µ)/∆) and ke/h,± given in the main text. Each of these wavefunctions carries aquasiparticle current, which for the Hamiltonian under consideration is given by:

jqp = 2Imψ†i(τzsy,−τzsx)ψ (S6)

2

As we want to use the wavefunctions of Eq. (S5) as a basis for the scattering matrices, we have to normalize them sothat they carry the same current in the direction perpendicular to the interface at the boundary of the narrow strip:

ψe/h,± =ψe/h,±

Ne/h,±, Ne/h,± =

√|jqp,y(ψe/h,±)| (S7)

Inside of the surrounding superconductor there will be also four possible solutions, but since Bx = 0 in that region,for ε within the gap ∆ the solutions will be exponentially decaying in either +y or −y direction. These wavefunctionsare:

ψSC1 =1

2

(−ieiα0+iβ0 , eiβ0 ,−ieiα0 , 1

)Teikxx−κy (S8a)

ψSC2 =1

2

(−ie−iα0−iβ0 , e−iβ0 ,−ie−iα0 , 1

)Teikxx−κy (S8b)

ψSC3 =1

2

(−ie−iα0+iβ0 , eiβ0 ,−ie−iα0 , 1

)Teikxx+κy (S8c)

ψSC4 =1

2

(−ieiα0−iβ0 , e−iβ0 ,−ieiα0 , 1

)Teikxx+κy (S8d)

where β0 = arccos ε/∆, α0 = Arg(kx + iky) and κ = kFk0

√ε2/v2 −∆2/v2.

Equipped with the wavefunctions in both regions we can now derive the normal and Andreev reflection coefficientsat the two interfaces at y = ±W/2. To this we solve the set of equations:

ψe,+ + rN1ψe,− + rA1ψh,+ = a1ψSC1 + b1ψSC2 (S9a)

ψh,− + r′N1ψh,+ + r′A1ψe,− = a′1ψSC1 + b′1ψSC2 (S9b)

ψe,− + rN2ψe,+ + rA2ψh,− = a2ψSC3 + b2ψSC4 (S9c)

ψh,+ + r′N2ψh,− + r′A2ψe,+ = a′2ψSC3 + b′2ψSC4 (S9d)

We solve these equations for the complex reflection coefficients rN (normal reflection) and rA (Andreev reflection) andfor clarity extract the phases acquired during the propagation across the strip region into the transmission matricesas indicated in the main text. The reflection coefficients at both interfaces can then be arranged into two scatteringmatrices S±W/2 at both boundaries of the narrow strip region:

SW/2 =

(rN1 r′A1

rA1 r′N1

)S−W/2 =

(rN2 r′A2

rA2 r′N2

)(S10)

As the scattering matrices are unitary, in general case they can be parametrized by four parameters each. In thesituation under consideration we therefore have:

SW/2 =

(reiφN1

√1− r2eiφ

′A1√

1− r2eiφA1 −rei(φA1+φ′A1−φN1)

)S−W/2 =

(reiφN2

√1− r2eiφ

′A2√

1− r2eiφA2 −rei(φA2+φ′A2−φN2)

)(S11)

For ε = 0 and kx = 0 the scattering matrices simplify greatly as r = 0 (no normal reflection due to spin-momentumlocking of Dirac surface states) and φA2 = −φ′A2 = −φA1 = φ′A1 = φA and we obtain Eq.(4) of the main text. In amore general scenario of ε 6= 0, while we still have r = 0, there will be two Andreev reflection phases describing thescattering. The bound state equation (5) of the main text will then reduce to:

r2A = e2i

kFk0

√(ε−Bx)2−∆2W , r

′2A = e−2i

kFk0

√(ε+Bx)2−∆2W (S12)

To obtain an approximate analytical solution to these equations, we expand to linear order in ε both the reflectioncoefficients rA and r′A obtained from Eq. (S9) and the expression in the exponent. Then we can make use of theLambert W function definition, arriving at:

E± = ±∆ Im

√B2x − 1

2Bx

1− ξ

WW0

2− B2x + 2i

√B2x − 1

B2x

eWξ (1+2i

√B2x−1)

(S13)

where W0(x) is the principal branch of the Lambert W function, Bx = Bx/∆ and ξ = v/∆. The comparison ofthe analytical solution with the numerical calculation based on the tight-binding model is shown in Fig. S1. Thesmall difference in the quasiparticle branch crossing points can be attributed to the imperfect approximation ofthe rotationally symmetric Dirac cone in the numerical tight-binding model and the use of projected low-energyHamiltonian in the analytical derivation.

3

2 3 4 5 6 7Bx/∆

1.0

0.5

0.0

0.5

1.0

E/∆

FIG. S1. Bound state energies at kx = 0 for increasing Zeeman energy. Solid red lines indicates the analytical solution ofEq. (S13).

TIGHT-BINDING MODEL FOR NUMERICAL CALCULATIONS

The numerical calculations were performed using a tight-binding approximation to the Hamiltonian (1) of the maintext, with terms included to ensure the presence of only a single Dirac cone in the Brillouin zone. The model isdiscretized on a square lattice with lattice constant a = 1, with the final Hamiltonian of the form:

HTB =∑j

c†j(2tτzsz − µτz +Bx(jx, jy)sx + ∆τx)cj +

(c†j+x(−i t

2τzsy −

t

2τzsz)cj + c†j+y(i

t

2τzsx −

t

2τzsz)cj + H.c.

)(S14)

where j = (jx, jy) is the index labeling each site of the tight-binding lattice, cj is the vector of annihilation operators inBogoliubov-de Gennes formalism, t is the hopping strength between the neighboring sites, µ is the chemical potential,∆ is the superconducting gap parameter and Bx(jx, jy) is the Zeeman energy, which is Bx underneath the magneticstrip and 0 otherwise. In some of the calculations we also include a potential barrier along the magnetic strip at itscenter to introduce normal reflection into the system. The barrier Hamiltonian is given by:

Hbarrier =∑

j:jy=0,

−LS2 <jx<LS2

Vbc†jτzcj (S15)

In the calculations we use the following values of the parameters: t = 1, µ = 0.8, ∆ = 0.02, Vb = 0.8. We change Bxand the width of the strip W as indicated for each simulation results figure.

We perform the simulations in two different configurations: (a) infinite strip with translational invariance in xdirection and (b) a finite strip of length LS that is fully surrounded by a gapped superconducting region of proximitizedDirac surface state. In both situations we apply periodic boundary conditions in y direction. In the first case we keepthe total system width WT = 300 lattice sites. We can then calculate the spectrum as a function of the longitudinalmomentum kx and in this way obtain the subgap quasiparticle modes as shown in Fig. 3(a) of the main text.This approach is also used to determine the phase diagram numerically, where the gap size plotted in Fig. 3(b) isdetermined by finding the smallest positive eigenvalue over all of kx. In the second case, we exactly diagonalize thefull tight-binding Hamiltonian matrix with the total system size of width WT = 250 and length LT = 500 latticesites. We then plot its eigenvalues as a function of the Zeeman energy in Fig. 4(a). Fig. 4(b) shows the local densityof states at E = 0 obtained as |ψ1|2 + |ψ2|2, where ψi are the two electron components of the wavefunctions obtainedfrom the diagonalization procedure.