Paper The following article is Open access

Chiral Majorana edge states in HgTe quantum wells

and

Published 9 August 2013 © IOP Publishing and Deutsche Physikalische Gesellschaft
, , Citation L Weithofer and P Recher 2013 New J. Phys. 15 085008 DOI 10.1088/1367-2630/15/8/085008

1367-2630/15/8/085008

Abstract

HgTe-based quantum wells (QWs) recently attracted a lot of attention for the realization of a two-dimensional topological insulator with protected helical edge states. Another class of topological systems is topological superconductors (TSCs) with Majorana edge states. In this paper, we show how proximity induced s-wave superconductivity in the bulk of HgTe-QWs and in the presence of a Zeeman field can exhibit a TSC with chiral Majorana edge states. We calculate the topological invariants and the corresponding Majorana edge states explicitly within a four-band model accounting for inversion symmetry breaking terms due to the Rashba spin–orbit coupling and bulk inversion asymmetry present in these QWs.

Export citation and abstract BibTeX RIS

Content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

The search for Majorana quasiparticles in condensed matter systems is a subject of great fundamental and practical interest. Majorana quasiparticles have recently been proposed to exist in various systems which typically combine strong spin–orbit coupling (SOC), proximity to an s-wave superconductor and magnetic elements [15].

In particular, Majorana bound states (MBS) have been proposed to emerge in vortex cores on the surface of strong three-dimensional (3D) topological insulators (TIs) [6] or at the edges of two-dimensional (2D) TIs when brought to the proximity of an s-wave superconductor and a magnet [7, 8]. Characteristic features of Majorana fermions have also been predicted in spin–orbit coupled 2D semiconductors [911] and one-dimensional (1D) spin–orbit coupled semiconductor nanowires [10, 12, 13]; very recent experiments show signatures of MBS which are currently being scrutinized [1417]. In these semiconductor systems, time-reversal symmetry is broken by a Zeeman splitting which, in addition to the proximity of an s-wave superconductor, is necessary to induce a topological superconductor (TSC). In such systems, a large SOC is desirable as the vulnerability of the energy gap to disorder typically depends on the size of the SOC [18, 19]. If the Zeeman splitting is induced by external magnetic fields, low field strengths and therefore small orbital effects are beneficial for the superconducting correlations, which favors semiconducting materials with large effective g-factors.

In this work we show that a doped HgTe-quantum well (QW) can implement a TSC, when s-wave superconductivity and a Zeeman-field are induced in the QW. In order to describe the HgTe-QW, we use the four-band Bernevig–Hughes–Zhang (BHZ) model [2022] valid near the transition between the inverted and non-inverted regime. In the inverted regime and in the absence of superconductivity and a Zeeman gap, the well-known helical edge states appear in the spectral bulk gap [20, 23, 24]. Importantly, HgTe-QWs are intrinsically subject to inversion symmetry breaking SOC due to bulk inversion asymmetry (BIA) [22, 25], which breaks the Kramer spin-degeneracy. We show that BIA-SOC alone is sufficient to induce a TSC when superconductivity and a Zeeman gap are added. However, tunable (extrinsic) inversion symmetry breaking terms provided by Rashba-SOC can be added as well.

It is known that HgTe-QWs are characterized by a very large Rashba-SOC [26], which can be one order of magnitude larger compared to other III–V compound semiconductors [2627]. In [26], Rashba-splittings in inverted HgTe-QWs as large as 30 meV were reported. In addition, the effective Zeeman gap in a perpendicular magnetic field was recently calculated and measured in zero gap samples to be consistent with an effective g-factor g* ∼ 55.5 [28]. This shows that HgTe-based QWs could be a feasible candidate for a proximity induced TSC.

In addition to a potentially feasible implementation of already existing ideas, we show that the combination of BIA and Rashba-SOC provides an interesting handle on the topological properties of the TSC. In particular, we show that (consistent with topological invariants) the chirality of Majorana edge states can be reversed by changing the relative strength of the two SOC terms.

In contrast to the proposal of using 1D helical edge states as a platform for MBS [6], we would like to point out in this work the possibility of switching between chiral (Majorana) and the well-known helical edge states on the same 1D boundary of a HgTe-QW.

The paper is organized as follows. In section 2 we introduce the setup and the relevant model Hamiltonian. Subsequently, we describe the bulk properties of the system, emphasizing the topologically non-trivial and trivial phases of the system in section 3. In section 4, we solve the chiral Majorana edge states at the domain wall between topologically non-trivial and trivial phases and calculate the group velocity along the edge. Finally, we compare the chiral Majorana edge states that emerge when the system is in a TSC phase to the helical edge states that emerge in a TI phase in section 5.

2. The model

We consider a doped HgTe-QW in the xy-plane in contact with an s-wave superconductor, as sketched in figure 1. There are two inversion2-asymmetry induced SOC effects: firstly, BIA-induced SOC is intrinsically present as a consequence of the zincblende structure of the underlying crystal. Secondly, so-called Rashba-SOC arises when an external electric field is applied in the z-direction. In addition, a Zeeman field is present in the z-direction, which can be induced by the proximity to a ferromagnetic insulator (via spin-exchange interaction), or intrinsically by using Mn-doped (Hg1−yMnyTe) QWs and a small external magnetic field polarizing the Mn-spins [29]. Although a magnetic field also induces an orbital effect not taken into account here, we note that the vector potential of a perpendicular magnetic field induces an additional Zeeman splitting between the Kramer spins with an effective giant g-factor of g* ∼ 55.5 [28] so that only small external magnetic fields below 1 T would be needed.

Figure 1.

Figure 1. HgTe-QW in proximity to an s-wave bulk superconductor.

Standard image High-resolution image

The resulting QW system is described in terms of the BHZ model [20], complemented with extra terms accounting for BIA [22], Rashba-SOC [21], Zeeman field [22, 29] and the proximity to an s-wave superconductor [30]. The Hamiltonian can be written in the basis order ψE1+,ψH1+,ψE1−,ψH1−,ψE1−,ψH1−,−ψE1+,−ψH1+ in terms of the products of the Pauli matrices $\underline {s}, \underline {\sigma },\underline {\tau }$ and the identity matrix, where $\underline {s}$ acts in the pseudospin (E1,H1)-space, $\underline {\sigma }$ acts in the Kramer spin space3 and $\underline {\tau }$ acts in the electron–hole space. The extended BHZ-model Hamiltonian reads

Equation (1)

with

Equation (2)

Equation (3)

Here, $\hat {k}_x$ and $\hat {k}_y$ denote the differential operators −i∂x and −i∂y.

The material parameters A,B,D,M depend on the geometry of the QW. In particular, the mass parameter M is tunable by the QW thickness. The BHZ-model was successfully used to explain experimental transport data for mass-parameters in the range −24 ⩽M ⩽ 10 meV [31]. The parameter C defines the doping of the QW (tunable by top and/or bottom gates).

The two additional SOC terms have coefficients denoted by h for the BIA term and R0 for the linear Rashba-SOC term4. As the BIA term is not invariant under rotation in the xy plane, the angle θ is introduced to describe the angle between the x-axis and the [100] crystal direction.

The pairing terms ΔE and ΔH describe the effect of the s-wave superconductor on the E1 and H1 bands. We neglect a possible off-diagonal proximity coupling between E1 and H1 bands and, without loss of generality, assume that the pairing terms ΔE and ΔH in (1) are real [30].

Typical values of the above mentioned parameters are presented in table 1. As Zeeman energy of the E1 and H1 bands, we use BE = 1.3 meV and BH = −0.07 meV, respectively. In addition, ${\cal E}=3\,\mathrm {mV}\,\mathrm {nm}^{-1}$ is assumed to be an effective electric field invoking the Rashba-SOC. Unless stated otherwise, we also use a doping parameter of C = 7.8 meV and θ = 0, as well as ΔE = 0.5 meV and ΔH = 0.5 meV for the superconducting proximity terms5.

Table 1. Typical values for the parameters of the inverted HgTe-QW, as taken from [21, 22].

A (eV Å) B (eV Å2) D (eV Å2) M(eV) hBIA(eV) $R_0/(e{\cal E})\,({\rm nm}^{2})$
3.645 −68.6 −51.2 −0.008 0.0016 −15.6

3. Bulk physics

3.1. Bulk inversion asymmetry spin–orbit coupling is sufficient

Firstly, we show that BIA-SOC is sufficient in order to lift the (Kramer) spin-degeneracy [3] by freezing out half the degrees of freedom (unless k = 0), which otherwise would make the spectrum doubly degenerate. Under periodic boundary conditions, kx and ky are good quantum numbers and the energy dispersion for the standard BHZ Hamiltonian supplemented with the BIA term can be calculated to

Equation (4)

Clearly, the spectrum is non-degenerate for |k| ≠ 0. By additionally introducing the Zeeman terms, the time-reversal symmetry is broken and thus the Kramer degeneracy at |k| = 0 is lifted.

The combined effect of BIA and Zeeman terms on the bulk energy dispersion is visualized in figure 2 (left). In the following, we choose the Fermi energy EF to be zero so that the Fermi level is situated within the gap between the E1-like bands opened by the Zeeman splitting. The doping parameter is tuned in such a way that C ≈ −M. As a result, only a single pair of Fermi points exists and thus the system has effectively spinless dynamics.

Figure 2.

Figure 2. Left: bulk energy dispersion E(kx,ky = 0) of a HgTe-QW with doping parameter C = 7.8 meV in the presence of Zeeman terms and BIA-SOC; Rashba-SOC is absent and only electron states are shown. Right: the same parameters are used but the s-wave superconducting proximity terms are included. The colors indicate the charge expectation value that ranges between e (red) and −e (blue).

Standard image High-resolution image

When, in addition, the proximity of the s-wave superconductor is taken into account, electrons with energy around the Fermi level and momenta k and −k couple, opening a gap around the Fermi level as shown in figure 2 (right). The special conditions under which this gap closes will be discussed in the following. The Fermi points at which this happens are the key to the emergence of the chiral Majorana edge states [32].

3.2. Symmetry properties

Before calculating the Fermi points, we will briefly review the symmetry properties of the two E1-like bands closest to the Fermi level. Their bulk energy spectrum is presented in figure 3 in the presence of different SOC terms: BIA or Rashba-SOC or the cooperation of BIA and Rashba-SOC. Interestingly, although the spectrum is rotationally symmetric when either Rashba-SOC or BIA are present, this rotational symmetry is broken when Rashba-SOC and BIA are simultaneously present.

Figure 3.

Figure 3. Bulk energy dispersion including superconductivity and Zeeman terms when (a) only the Rashba-SOC term is included, (b) only the BIA term is included and (c) both terms are included. Top: energy dispersion E(kx) where ky = 0 (dashed line) compared to energy dispersion Emin(kx) where ky is such that the energy becomes minimal (solid line). Bottom: energy dispersion E(kx,ky) for the minimal positive energy band. Vectors illustrate in-plane spin expectation values projected onto the E1-band 〈σx(1 + sz)/2〉 and 〈σy(1 + sz)/2〉.

Standard image High-resolution image

This can be understood as follows [33]: while the effective magnetic fields due to Rashba-SOC are oriented clockwise and perpendicular to the wave vector k, the effective magnetic fields due to the BIA spin splitting have tetrahedral symmetry, as shown in figure 3. When both BIA and Rashba-SOC are present, the effective magnetic fields add vectorially, decreasing or enhancing each other. The energy of the lowest-energy band is very sensitive to the effective k-dependent in-plane magnetic field. Therefore, it exhibits tetrahedral symmetry when BIA- and Rashba-SOC terms interfere with each other.

3.3. Fermi points

In the following, we will identify the Fermi points of the spectrum. To simplify analytical calculations, we project the full Hamiltonian (1) onto an effective Hamiltonian. When M defines the dominant energy scale and the doping is chosen as C ≈ −M, four of the eight bands become irrelevant. Using quasi-degenerate perturbation theory [33] ('Löwdin partitioning'), the eight-band Hamiltonian (1) can be reduced to an effective four-band Hamiltonian acting on the E1-like electron and hole states only (see appendix)

Equation (5)

where the new parameters $\mu ({\hat k}^2)=\epsilon ({\hat k}^2)+M({\hat k}^2) +(A^2{\hat k}^2+h^2)/(2M)$ and $\tilde {h}=Ah/M$ have been introduced.

In the absence of BIA-SOC, the reduced form (5) is analogous to the effective Hamiltonians of different semiconductor systems, which have already been shown to host chiral Majorana edge states [913]. Note that the Rashba-SOC energy UR ≡ 2m*R20 [18] is kept quite small here intentionally in order to study the combined effect of BIA and Rashba-SOC: for the values in table 1, it amounts to $U_{\mathrm { R}}=R_{0}^{2}/(\frac {A^2}{2M}-D-B) \approx 0.3\,\mathrm {meV}$ . On the other hand, the BIA-SOC leads to an effective Dresselhaus-SOC in the reduced model (5), which amounts to $U_{\mathrm {D}}=(Ah/M)^{2}/(\frac {A^2}{2M}-D-B) \approx 0.7\,\mathrm {meV}$ . In [34], electric fields up to $\mathcal {E}{\sim }100\,\mathrm {mV}\,\mathrm {nm}^{-1}$ were estimated to be possible, which could enhance the Rashba-SOC energy in principle by a factor of 30, without affecting the effective Dresselhaus energy.

The effective Hamiltonian (5) yields (for θ = π/4 and ΔE ≠ 0) zero energy level crossings at

Equation (6)

This implies that for non-vanishing SOC, Fermi points exist (for B2E − μ(k2)2 > 0) at

Equation (7)

Equation (8)

Equation (9)

where μ0 = μ(k2 = 0) = C + M + h2/(2M). While the Fermi point (7) exists also in other semiconductor systems that have been proposed to host chiral Majorana edge states [913], the Fermi points (8) and (9) are characteristic of a system with two different SOC terms.

3.4. Bulk topological invariant

In this section, we identify the topological phases of the non-trivial momentum space topology of Hamiltonian (1). The fundamental Hamiltonian (1) belongs to the Cartan–Altland–Zirnbauer symmetry class D, which in two spatial dimensions is characterized by a topological invariant $\in \mathbb {Z}$  [35]. The first Chern number which is commonly used as the topological invariant can be obtained for a non-interacting Hamiltonian with non-degenerate spectrum as the integral of the Berry curvature over the first Brillouin zone

Equation (10)

where the Berry curvature $\mathcal {F}=\partial _{k_x}\mathcal {A}_{k_y}-\partial _{k_y}\mathcal {A}_{k_x}$ is obtained from the the Bloch functions via $\mathcal {A}_{k_\alpha }=-\mathrm {i}\sum _j\langle u _j(k)|\partial _{k_\alpha }|u_j(k)\rangle $ , α = x,y, the index j running over the occupied states, as reviewed in [36].

We evaluate the general formula of the first Chern number (10) for the parameterized Hamiltonian (1) both numerically and analytically in the parameter regime outlined previously6.

Numerically, the first Chern number is computed from a discretized version [37] of (10) for a regularized version7 of Hamiltonian (1) using MATLAB at several illustrative values of the parameters on each side of the topological transitions. The topological transitions have been obtained by identifying the zero energy level crossings of the effective four-band Hamiltonian, see (7)–(9). This provides a good approximation for the topological transitions of the full model (1) in the considered parameter regime.

The analytical calculation [38] of the first Chern number is performed in the limit where an effective two-band description is possible [39] (see appendix).

We find that the first Chern number characterizing the topological phases of our physical system assumes the values 0,1,−1 in the considered parameter regime, depending on the specific values of the parameters in the following way:

Equation (11)

where Θ(x) denotes the Heaviside step function.

The Chern numbers 1 and −1 indicate non-trivial topology, whereas the Chern number 0 indicates trivial topology which is adiabatically connected to a vacuum.

While the general form of (11) is analogous to known topological invariants of similar systems, such as spin–orbit coupled quantum wires [10, 12, 13] or 2D spin–orbit coupled semiconductors [911], this constitutes, to our knowledge, the first such calculation for a HgTe-QW in proximity to a superconductor. Importantly, this is also the first time that the BIA term h is included. When both Rashba-SOC and BIA are present, the sign of the topological invariant $\mathcal {C}_1$ can be reversed due to the interplay between both SOCs. This is similar to the combined effect of Rashba and Dresselhaus SOCs that was described for non-centrosymmetric superconductors [40].

In the absence of BIA, the relevant parameters for a topological transition to occur are the proximity coupling ΔE, the Zeeman energy BE and the sum of the mass gap and doping, μ0 = C + M. When BIA is taken into account, the topological transition is shifted toward higher doping, as described by the term $\mu _0=C+M+\frac {h^2}{2M}$ (in the limit that M is the dominant energy scale and C ≈ −M). The topological transition is presented as a function of Zeeman energy BE and doping C in figure 4.

Figure 4.

Figure 4. Topological phase diagram in the parameter region Zeeman energy BE and doping C. The topological transition described approximatively by (11) (blue solid line when BIA is present, red dashed line when BIA is absent) is compared to numerical calculations based on (1) (blue triangle when BIA is present, red square when BIA is absent). When BIA asymmetry is taken into account, the topological transition is shifted toward higher doping. The other parameters are given in table 1.

Standard image High-resolution image

4. Majorana edge states

In this section, we discuss the appearance of Majorana edge states in this system.

By an edge we mean that there is a domain wall in the shape of a straight edge where one side is topologically non-trivial ( $\mathcal {C}_1=\pm 1$ ) and the other side is topologically trivial ($\mathcal {C}_1=0$ ). We distinguish between two cases: either (i) the topologically non-trivial region is a vacuum or (ii) a topologically trivial QW. In the following, these cases will be referred to as (i) hard-wall boundary conditions and (ii) soft-wall boundary conditions. Soft-wall boundary conditions (two different topological phases in a QW) can be achieved by varying the doping parameter C accordingly across the sample, as shown in figure 5(b).

Figure 5.

Figure 5. Normalized probability density at kx = 0 as a function of y for hard-wall boundary conditions separating a topologically non-trivial QW from a vacuum (trivial) (a) and soft-wall boundary conditions separating a topologically non-trivial QW region from a topologically non-trivial QW region (b).

Standard image High-resolution image

Without loss of generality, we assume that the domain wall lies parallel to the x-axis and that the region y < 0 is topologically non-trivial while y > 0 is topologically trivial. Due to translational invariance along the x-axis, we can consider solutions that propagate along the edge with definite kx.

The general strategy when solving such a boundary problem is as follows: in a first step, we use the ansatz

Equation (12)

to calculate the wave functions on both sides of the boundary. Hereby, ky, which is not a good quantum number, has to be determined as a function of kx and the energy E separately for both sides of the boundary. In addition, the eigenvectors are determined as a function of kx and E.

In the second step, the boundary conditions are applied, finally yielding an energy dispersion E(kx). For hard-wall boundary conditions, the wave function is required to vanish at the boundary between the HgTe sample and the vacuum. For soft-wall boundary conditions, both the wave function and its first derivative are required to be continuous at the boundary between the topologically trivial and the topologically non-trivial HgTe samples. In both cases, we require that the solutions are bound to the edge, i.e. that the probability densities vanish far away from the boundary (see figure 5).

4.1. Calculation of edge state dispersion for small kx

Under certain simplifying assumptions, the edge dispersion of the boundary problem can be calculated to first order in k using perturbation theory. We assume that it is sufficient to calculate the edge dispersion in the reduced model (5) and require that quadratic terms ∝k2x,k2y may be neglected (assumption 1). Consistently, we also match only the wave functions and not the derivatives at the boundary.

We first prove that the boundary conditions are fulfilled for E = 0, kx = 0 and calculate the wave function of the corresponding bound state.

Imposing the condition E = kx = 0 in (5) yields four possible values for λ = iky:

Equation (13)

Depending on the parameter values, the signs of λ can be determined as follows under the assumption that |BE| > |μ0|:

The respective eigenvectors are

Equation (14)

with $\mathrm {e}^{\mathrm {i}\chi }=(\mathrm {i}R_0+\tilde {h}\mathrm {e}^{2\theta \mathrm {i}})/\sqrt {R_0^2+\tilde {h}^2+2R_0\tilde {h}\sin {2\theta }}$ and where |λ〉, | − λ+〉 and | − λ〉 can be obtained from |λ+〉 by a clockwise translation of the minus sign in front of the second entry.

If B2E > μ2l + Δ2E on the left side of the boundary (topologically non-trivial) and B2E < μ2r + Δ2E on the right side of the boundary (topologically trivial), the matching condition requires that the eigenvectors on the left- and right-hand sides of the boundary are linearly dependent

Equation (15)

yielding the solution

Equation (16)

Here $\mathcal {N}$ is a normalization factor such that $\int \rm d y|\psi (y)|^2=1$ . The coefficients c1 and c2 are determined by the matching condition.

This proves that there is a bound state for kx = E = 0. The antisymmetry of the Hamiltonian (1) under particle–hole transformation ΞHΞ = −H, where Ξ = iKσyτy, has the important consequence that ψ(y) for kx = E = 0 constitutes a Majorana excitation, as can be easily confirmed by explicit calculation.

Now we consider the case of non-zero kx. When kx is small, the kx-dependent BIA and Rashba-SOC terms in equation (5) can be treated as a perturbation V . The expectation value of this perturbation with respect to the kx = E = 0 state gives the energy in terms of kx to first order in perturbation theory

Equation (17)

4.2. Edge state dispersions

The resulting edge energy dispersions that were calculated numerically from (1) and using (17) are presented and compared to the bulk energy dispersions in figure 6.

Figure 6.

Figure 6. Energy dispersion with a boundary in the x-direction. On the left-hand side of the boundary there is a topologically non-trivial QW with Cl = 7.8 meV (black bulk energy minimum), while on the right-hand side of the boundary there is either a vacuum (a) or a topologically trivial QW with Cr = 6.8 meV, indicated by the green bulk energy minimum (b). The chiral Majorana edge states obtained numerically as boundary solutions to (1) are shown in red. The red dotted lines are estimations derived from (17). The parameter values are taken from table 1.

Standard image High-resolution image

The boundary between topologically trivial and non-trivial regions results in a single chiral MBS of approximately linear energy dispersion that smoothly connects to the occupied and non-occupied bulk bands. The existence of this edge state is independent of which boundary condition (soft or hard wall) or which inversion asymmetry term (BIA, Rashba-SOC or both) is considered. As expected, its existence depends on the topological bulk properties on both sides of the boundary—the edge state disappears when the Chern numbers on both sides of the boundary are identical.

The formula (17) describes the numerically obtained energy dispersions of figure 6 well, regardless of whether hard-wall or soft-wall boundary conditions have been employed. A small discrepancy between the numerical calculation on the basis of the Hamiltonian (1) and (17) is mostly due to the limited applicability of assumption 1. Yet, (17) describes the salient features of the edge dispersion well.

As predicted by (17), the edge state velocity depends most strongly on the parameters θ,A,h,R0,BEE. We have checked that the parameters D,B,CH,BH have comparably little effect on the energy dispersion.

Interestingly, we find that the group velocity $v=\frac {1}{\hbar }\frac {\partial E}{\partial k_x}$ of the chiral edge state is tunable by the interplay between the Rashba-SOC and BIA spin–orbit interactions. The edge state is right-moving when the Rashba-SOC term dominates and left-moving when the BIA term dominates (see figure 7). This is consistent with the change of the first Chern number in (11) between $\mathcal {C}_1=\pm 1$ by means of the index theorem [32]. When applied to the hard-wall boundary condition, the index theorem equates the topological charge $\mathcal {C}_1$ of the bulk with the sign ν of the slope of the chiral edge mode. It is therefore easy to see that as the Rashba-SOC parameter R0 changes from $R_0^2{<}\tilde {h}^2$ to $R_0^2{>}\tilde {h}^2$ , the first Chern number changes from −1 to +1 and the edge state from left to right moving. This could be experimentally verified by changing the Rashba-SOC parameter, while the intrinsic BIA-SOC remains fixed. Another interesting observation is that the group velocity of the chiral Majorana edge state depends on the orientation of the edge with respect to the crystal axis when both Rashba-SOC and BIA-SOC terms are present. The absolute values of the velocity are smallest along the [110] (θ = π/4) and largest along the $[\bar {1}10]$ (θ = 3π/4) direction. In HgTe-QWs, the edge structures are, for experimental reasons, prepared in parallel to the cleavage planes; therefore, 95% of the current carrying edges are orientated either in the [110] or $[\bar {1}10]$ direction [41]. Due to the simultaneous presence of Rashba-SOC and BIA-SOC terms, these two experimentally accessible edge orientations are expected to have distinct chiral Majorana edge state properties.

Figure 7.

Figure 7. The edge group velocities $v=\frac {1}{\hbar }\frac {\partial E}{\partial k_x}$ as a function of different parameters. Analytic approximations (solid black lines) obtained from (17) are compared with numerical values for hard-wall (red triangles) and soft-wall boundary conditions (blue crosses). This confirms that the edge state only exists when the boundary connects topologically trivial and non-trivial regions, compare (11). The Rashba-SOC is set to zero except in the last line; all other parameters are taken from table 1.

Standard image High-resolution image

5. Comparison of chiral Majorana edge states and helical edge states

As described in the previous section, a HgTe-QW in proximity to a superconductor with BIA and a Zeeman field hosts chiral Majorana edge states at its edges when the parameters are in the topological non-trivial TSC phase ($\mathcal {C}_1=\pm 1$ ).

On the other hand, it is well known that a time-reversal invariant HgTe-QW hosts helical edge states at its boundary [20, 23, 42] when the mass parameter M is negative (B/M > 0).

In the following, we will compare the chiral Majorana edge states and the helical (Dirac) edge states, assuming that only the BIA-SOC term is present.

Chiral Majorana edge states and helical edge states at a single boundary are mutually exclusive. Firstly, helical edge states are not topologically protected when time-reversal symmetry is broken by a Zeeman field, which is needed for the emergence of chiral Majorana edge states [42]. Secondly, it has recently been shown that even in the presence of time-reversal symmetry, the introduction of the superconducting term Δτx adiabatically interconnects the regions that in its absence would differ by their topological invariants [43].

In the parameter range where there is a chiral Majorana edge state, there are therefore no additional helical edge states within the bulk energy gap, see figures 8(a) and (b). Even if the mass gap M is negative, potential helical edge states are gapped out by the proximity of the superconductor.

Figure 8.

Figure 8. Energy dispersion with a hard-wall boundary in the x-direction for a HgTe-QW with topologically non-trivial parameters (a) C = −7.8, M = 8 meV and (b) C = 7.8, M = −8 meV. This allows chiral Majorana edge states to form. All other parameters except the Rashba-SOC, which is set to zero, are taken from table 1. In (c), and (d), the energy dispersions of a time-reversal invariant HgTe-QW in the inverted regime without the presence of a superconductor are shown (C = 6.8, M = −8 meV). For better comparison with (b), the degrees of freedom have been doubled in (d) by reflection at the Fermi level.

Standard image High-resolution image

For comparison to the chiral Majorana edge states, the helical edge states are shown in the absence of both the Zeeman field and the superconductor in figures 8(c) and (d), where in contrast to the standard HgTe, the additional BIA term is included. As reported in [44], it has a negligible effect and the velocity of the helical edge states is still well described by [45]

Equation (18)

for small |kx|. The chiral and helical edge states differ greatly in their velocities, which are given by (17) and (18), respectively. The offset is also different: while chiral Majorana edge states are confined to the crossing kx = E = 0 by particle–hole symmetry, the helical edge states are only restricted to crossing each other somewhere on the E axis by time-reversal symmetry.

Helical and chiral edge states have similar widths. While recent calculations for the width of helical edge states predict an edge width under 100 nm in a HgTe-QW near the Dirac point [45], experimentally confirmed up to a factor of 2 in [46], the chiral Majorana edge states typically extend to about 200–300 nm in a HgTe-QW for the parameters in table 1.

6. Summary

In summary, we have shown that a HgTe/CdTe QW in the proximity of an s-wave superconductor and the presence of a Zeeman-splitting can represent a TSC with chiral 1D edge states. At kx = E = 0, these quasiparticles have the Majorana property of being their own antiparticle.

Importantly, an external electric field that leads to an additional Rashba-SOC term is not required, as long as the intrinsic BIA is large enough. While in most publications, BIA is discussed as an unwanted but small perturbation, in this case a high BIA would be of interest. Interestingly, if Rashba-SOC and BIA-SOC are both present, the group velocity $v=\frac {1}{\hbar }\frac {\partial E}{\partial k_x}$ of the chiral edge state depends crucially on the interplay between both spin–orbit interactions and can be reversed if the Rashba-SOC is tuned from zero to a value that exceeds the BIA term. We discussed in analytical and numerical terms in which parameter regime the TSC exists using the extended four-band model of BHZ, valid near the topological phase transition between a trivial and a non-trivial TI that would exist without the superconducting and Zeeman terms.

Although many systems hosting chiral Majorana fermions have so far been presented [6, 12], this is the first proposal to switch between 1D Majorana and Dirac edge states in a (HgTe) QW, rather than the surface of a 3D TI, where different scenarios to convert between 1D Majorana and Dirac edge states have previously been proposed [4749] and the first time chiral Majorana and helical Dirac edge states appear in the same system.

The possibility of switching between chiral Majorana and helical Dirac edge states on the same 1D boundary might not only enhance experimental control in the detection of Majorana quasiparticles, but could also provide an interesting framework that combines dissipationless transport and the non-local storage of quantum information.

Importantly, we would like to point out that our analysis should equally hold well for other 2D TIs, like type-II QWs made of InAs/GaSb/AlSb, which have intrinsically inverted bandstructures and similar inversion symmetry breaking SOC terms [50]. Experimental evidence of the quantum spin Hall effect in these systems has recently been given [51], also in connection with superconducting contacts [52]. Further, it has been proposed that bilayer HgTe-based QWs in the presence of an interlayer voltage allows for an all electrically tunable TI [53, 54].

Acknowledgments

We thank Hartmut Buhmann for enlightening discussions, Jan Budich and Björn Trauzettel for their useful comments on the paper and the financial support from the 'NTH school for contacts in nanosystems', the DFG grant RE 2978/1-1 and the EU-FP7 project SE2ND.

Appendix

A.1. Derivation of effective four-band Hamiltonian

If the mass gap M is the dominant energy scale of the problem and C ≈ −M, leading to only four relevant E1-like energy bands, this allows us to reduce the eight-band Hamiltonian (1) to an effective four-band Hamiltonian.

To this end, we divide the Hamiltonian into three parts:

Equation (A.1)

Equation (A.2)

Equation (A.3)

Equation (A.4)

where for clarity, in contrast to the more general Hamiltonian (1), we set θ = 0.

In the following, H0 (containing only sz) is considered as the unperturbed Hamiltonian and H1 (containing only sz) and H2 (containing sy) as the block-diagonal and block-offdiagonal perturbations, respectively. The perturbations are small in the sense that the mass gap M is the dominant energy scale, i.e.

Equation (A.5)

In order to approximately diagonalize the Hamiltonian H

Equation (A.6)

we apply quasi-degenerate perturbation theory ('Löwdin partitioning') [33].

The important matrix for the transformation, S, consists of different orders of the perturbation parameter, $S=S^{(1)}+\mathcal {O}(2)$ . It can be shown that

Equation (A.7)

is block-diagonal up to the second order if S(1) is determined in such a way that

Equation (A.8)

For the Hamiltonian (A.1) it is straightforward to show8 that

Equation (A.9)

Equation (A.10)

where the new parameters

Equation (A.11)

Equation (A.12)

Equation (A.13)

Equation (A.14)

Equation (A.15)

have been introduced. The advantage of the transformed Hamiltonian (A.10) is that sz is (approximately) a good quantum number. For further calculations, we neglect the sz = −1 solutions in the main text (5), and retrieve the angular dependence on θ by the simple replacement rules

Equation (A.16)

Equation (A.17)

which amount to a rotation in the xy-plane.

A.2. Analytical calculation of the first Chern number

In order to verify (11) of the main text in an analytical way, we project the four-band Hamiltonian

Equation (A.18)

or, equivalently,

Equation (A.19)

onto a two-band Hamiltonian. Since the precise value of the angle θ does not affect the topological properties of the bulk, we have chosen $\theta =\frac {\pi }{4}$ here for convenience. The projection procedure is similar to [39] and holds in the limit $B_E\gg m(R^2+\tilde {h}^2),|\Delta _E|$ where $m\!=\!\frac {A^2}{2M}\!-\!D\!-\!B$ .

This projection relies on the approximations ψ ≈ ψ and $\psi _\uparrow \approx \frac {(R_0+\tilde {h})k_y+\mathrm {i}(R_0-\tilde {h})k_x}{2B_E}\psi _-$ and yields

Equation (A.20)

where μeff = μ − BE, $\Delta ^x_{\mathrm {eff}}=\frac {|\Delta _E|}{2|B_E|}(R_0-\tilde {h})$ and $\Delta ^y_{\mathrm {eff}}=\frac {|\Delta _E|}{2|B_E|}(R_0+\tilde {h})$ .

After regularization, this can be written as

Equation (A.21)

Equation (A.22)

Equation (A.23)

Equation (A.24)

For this two-band model, the first Chern number is defined as [1]

Equation (A.25)

where $\hat {{\bf{ h}}}$ is defined as $\hat {{\bf{ h}}}={\bf{ h}}/|{\bf{ h}}|$ .

The explicit integration is cumbersome and can be greatly simplified using the equivalent formula [38] which consists of a sum over the 'Dirac points' $D_3=\{(0,0),(0,\frac {\pi }{a}),(\frac {\pi }{a},0),(\frac {\pi }{a},\frac {\pi }{a})\}$ of the two-Pauli-matrix Hamiltonian h1s1 + h2s2,

Equation (A.26)

Equation (A.27)

If we assume that the regularization is such that $\frac {1}{a^2m}\gg |\mu _{\mathrm {eff}}|$ , then the Chern number simplifies to

Equation (A.28)

which is for BE > 0 consistent with formula (11) in the main text.

In the case BE < 0, the projection relies on the approximations ψ ≈ ψ and $\psi _\downarrow \approx \frac {-(R_0+\tilde {h})k_y-\mathrm {i}(R_0-\tilde {h})k_x}{2B_E}\psi _-$ and yields

Equation (A.29)

which is for BE < 0 consistent with formula (11) in the main text.

In summary, we have verified

Equation (A.30)

and therefore equation (11) in the main text.

Footnotes

  • That is (x,y,z) → − (x,y,z).

  • The quantum number that transforms under time reversal. The electron spin is not a good quantum number in the presence of spin–orbit interactions.

  • Rashba-SOC terms of higher order in k are neglected.

  • In reality, ΔE and ΔH will not be exactly equal, however, in our parameter regime C ≈ −M (see below), the dominant contribution will come from ΔE.

  • That is BIA and Rashba-SOC should not both be zero. The parameters A,B,D and M describing the HgTe-QW are considered as constants, except for the doping parameter C which may vary around C ≈ −M. All other parameters may vary, as long as M remains the dominant energy scale.

  • That is replacing $k_x^2+k_y^2\rightarrow \frac {2}{a^2}\left (2-\cos (ak_x)-\cos (ak_y)\right )$ , $k_x\rightarrow \frac {1}{a}\sin (ak_x)$ , $k_y\rightarrow \frac {1}{a}\sin (ak_y)$ where a is a lattice constant, which is chosen in such a way that no artificial level crossings occur.

  • Under the requirement that the parameters do not depend on x or y.

Please wait… references are loading.