This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Paper The following article is Open access

Exact eigenvectors and eigenvalues of the finite Kitaev chain and its topological properties

, , and

Published 11 August 2020 © 2020 The Author(s). Published by IOP Publishing Ltd
, , Citation Nico Leumer et al 2020 J. Phys.: Condens. Matter 32 445502 DOI 10.1088/1361-648X/ab8bf9

0953-8984/32/44/445502

Abstract

We present a comprehensive, analytical treatment of the finite Kitaev chain for arbitrary chemical potential and chain length. By means of an exact analytical diagonalization in the real space, we derive the momentum quantization conditions and present exact analytical formulas for the resulting energy spectrum and eigenstate wave functions, encompassing boundary and bulk states. In accordance with an analysis based on the winding number topological invariant, and as expected from the bulk-edge correspondence, the boundary states are topological in nature. They can have zero, exponentially small or even finite energy. Further, for a fixed value of the chemical potential, their properties are ruled by the ratio of the decay length to the chain length. A numerical analysis confirms the robustness of the topological states against disorder.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 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 quest for topological quantum computation has drawn a lot of attention to Majorana zero energy modes (MZM), quasi-particles obeying non-Abelian statistics hosted by topological superconductors [1]. The archetypal model of a topological superconductor in one dimension was proposed by Kitaev [2]. It consists of a chain of spinless electrons with nearest neighbour superconducting pairing, a prototype for p-wave superconductivity. As shown by Kitaev in the limit of an infinite chain, for a specific choice of parameters, the superconductor enters a topological phase where the chain can host a couple of unpaired zero energy Majorana modes at the end of the chain [2]. This model has thus become very popular due to its apparent simplicity and it is often used to introduce topological superconductivity in one dimension [1, 3]. Also more sophisticated realizations of effective p-wave superconductors, based on semiconducting nanowire-superconductor nanostructures [411], ferromagnetic chains on superconductors [1216] or s-wave proximitized carbon nanotubes [1719], all rely on these fundamental predictions of the Kitaev model. While the theoretical models are usually solved in analytic form for infinite or semi-infinite chains, the experiments are naturally done on finite-length systems. For example, for the iron chain on a superconductor investigated in [13] it is expected that the chain length is shorter than the superconducting coherence length [20]. Spectral properties of a finite-length Kitaev chain have been addressed in more recent papers [2124], and have confirmed the presence of bound states of exponentially small energy in sufficiently long finite Kitaev chains.

As noticed by Kao using a chiral decomposition [21], a finite-length Kitaev chain also supports modes with exact zero-energy. However, they are only found for discrete values of the chemical potential. As shown by Hegde et al, in [22], these exact zeros can be associated to a fermionic parity crossing in the open Kitaev chain. Investigations of the finite chain have also been performed by Zvyagin [23] using the mapping of a Kitaev chain onto an X–Y model for N spin 1/2 particles in transverse magnetic field, for which convenient diagonalization procedures are known [25, 26]. Kawabata et al in [27] demonstrated that exact zero modes persist also in a Kitaev chain with twisted boundary conditions.

Despite much being known by now about the finite length Kitaev chain, and in particular its low energy properties, the intricacy of the eigenvector equation still constitutes a challenge. In this work we address this longstanding problem. Exact results for the full energy spectrum and the associated bound states are provided for arbitrary chemical potential by analytical diagonalization in the real space. Our results are not restricted to long chains or to long wavelengths, and thus advance some of the findings in [22, 23] and respectively [24]. They also complete the analysis of the eigenstates of an open Kitaev chain performed by Kawabata et al [27] which was restricted to the exact zero modes. This knowledge allows one a deeper understanding of the topological properties of the Kitaev chain and, we believe, also of other one-dimensional p-wave superconductors.

Before summarizing our results, we clarify the notions used further in our paper. (i) Since 'phase' properly applies only to systems in the thermodynamic limit, we shall use 'topological regime' to denote the set of parameters for which a topological phase would develop in an infinite system. (ii) We shall call 'topological states' all boundary states of the finite system whose existence in the topological regime is enforced by the bulk-boundary correspondence [1]. Hence, the existence of the topological states is associated to a topological invariant of the bulk system being non trivial. Quite generally, the topological states can have zero, exponentially small, or even finite energy. (iii) When the gap between the topological states and the higher energy states is of the same order or larger than the superconducting gap Δ, we consider them to be 'robust' or 'topologically protected'. Their energy may be affected by perturbations, but not sufficiently to make them hybridize with the extended (bulk) states. (iv) All Hamiltonian eigenstates can be written as superpositions of Majorana (self-conjugate) components. When the energy of the topological state is strictly zero, the whole eigenstate has the Majorana nature and becomes a true 'Majorana zero mode' (MZM).

Using our analytical expressions for the eigenstates of a finite Kitaev chain, we recover in the limit of an infinite chain the region for the existence of the MZM given by the bulk topological phase diagram; the latter can be obtained using the Pfaffian [2] or the winding number [28] topological invariant. For a finite-length chain MZM only exist for a set of discrete values of the chain parameters, see equation (87) below, in line with [21, 27]. These states come in pairs and, depending on the decay length, they can be localized each at one end of the chain but they can also be fully delocalized over the entire chain. Even in the latter case the states are orthogonal and do not 'hybridize' since they live in two distinct Majorana sublattices. Similar protection of topological zero energy modes living on different sublattices has recently been observed experimentally in molecular Kagome lattices [29].

The paper is organized as follows. Section 2 shortly reviews the model and its bulk properties. Section 3 covers the finite size effects on the energy spectrum and on the quantization of the wave vectors for some special cases, including the one of zero chemical potential. The spectral properties at zero chemical potential are fully understood in terms of those of two independent Su–Schrieffer–Heeger-like (SSH-like) chains [27, 30]. The eigenstates at zero chemical potential, the symmetries of the Kitaev chain in real space, as well as the Majorana character of the bound state wave functions are discussed in section 4. In sections 57 we turn to the general case of finite chemical potential which couples the two SSH-like chains. While section 5 deals with the energy eigenvalues and eigenvectors of the finite chain, section 6 provides exact analytical results for the MZM. In section 7 the influence of disorder on the energy of the lowest lying states is investigated numerically. In section 8 conclusions are drawn. Finally, appendices AG contain details of the factorisation of the characteristic polynomial in real space and the calculation of the associated eigenstates.

2. The Kitaev chain and its bulk properties

2.1. Model

The Kitaev chain is a one dimensional model based on a lattice of N spinless fermions. It is characterized by three parameters: the chemical potential μ, the hopping amplitude t, and the p-wave superconducting pairing constant Δ. The Kitaev Hamiltonian, written in a set of standard fermionic operators ${d}_{j},\;{d}_{j}^{{\dagger}}$, is [1, 2]

Equation (1)

where the p-wave character allows interactions between particles of the same spin. The spin is thus not explicitly included in the following. We consider Δ and t to be real parameters from now on.

The Hamiltonian in equation (1) has drawn particular attention in the context of topological superconductivity, due to the possibility of hosting MZM at its end in a particular parameter range [2]. This can be seen by expressing the Kitaev Hamiltonian in terms of so called Majorana operators γA,B,

Equation (2)

yielding the form

Equation (3)

Notice that, in virtue of equation (2) it holds $\left\{{\gamma }_{j}^{A},\;{\left({\gamma }_{j}^{A}\right)}^{{\dagger}}\right\}=2\;{\left({\gamma }_{j}^{A}\right)}^{2}=1$, and similarly for ${\gamma }_{j}^{B}$. For the particular parameter settings Δ = ±t and μ = 0, which we call the Kitaev points, equation (3) leads to a 'missing' fermionic quasiparticle q±:

Equation (4a)

Equation (4b)

This quasiparticle has zero energy and is composed of two isolated Majorana states localised at the ends of the chain. In general, the condition of hosting MZM does not restrict to the Kitaev points (μ = 0, Δ = ±t). Further information on the existence of boundary modes is evinced from the bulk spectrum and the associated topological phase diagram. If the boundary modes have exactly zero energy, their Majorana nature can be proven by showing that they are eigenstates of the particle–hole operator $\mathcal{P}$. Equivalently, if ${\gamma }_{M}^{{\dagger}}$ is an operator associated to such an MZM it satisfies ${\gamma }_{M}^{{\dagger}}={\gamma }_{M}$ and ${\gamma }_{M}^{2}=1/2$.

The topological phase diagram is shortly reviewed in section 2.3.

2.2. Bulk spectrum

The Hamiltonian from equation (1) in the limit of N reads in k space

Equation (5)

where we introduced the operators ${d}_{k}=\frac{1}{\sqrt{N}}\sum _{j}{\mathrm{e}}^{-\text{i}\;j\;kd}\;{d}_{j}$ and k lies inside the first Brillouin zone, i.e. $k\in \left[-\frac{\pi }{d},\frac{\pi }{d}\right]$ and d is the lattice constant. The 2 × 2 Bogoliubov–de Gennes (BdG) matrix

Equation (6)

is easily diagonalized thus yielding the excitation spectrum

Equation (7)

Note that for μ = 0 equation (7) predicts a gapped spectrum whose gap width is either 4Δ (|Δ| < |t|) or 4t (|t| < |Δ|).

2.3. Topological phase diagram

The BdG Hamiltonian (6) is highly symmetric. By construction it anticommutes with the particle–hole symmetry $\mathcal{P}={\sigma }_{x}\mathcal{K}$, where $\mathcal{K}$ accounts for complex conjugation. The particle–hole symmetry turns an eigenstate in the k space corresponding to an energy E and wavevector k into one associated with −E and −k. The time reversal symmetry is also present in the Kitaev chain and is given by $\mathcal{T}=\mathbb{1}\;\mathcal{K}$. Finally, the product of $\mathcal{TP}=C={\sigma }_{x}$ is the chiral symmetry, whose presence allows us to define the topological invariant in terms of the winding number [28]. Note that all symmetries square to +1, placing the Kitaev chain in the BDI class [31].

The winding number is given by [28, 32]

Equation (8)

where $w\left(k\right)=\mathrm{arg}\left[2{\Delta}\;\mathrm{sin}\left(kd\right)+i\;\left(\mu +2t\;\mathrm{cos}\left(kd\right)\right)\right]$ and ∂kw(k) is the winding number density. A trivial phase corresponds to ν = 0, a non trivial one to finite integer values of ν. The winding number relates bulk properties to the existence of boundary (not necessarily MZM) states in a finite chain. This property is known as bulk-edge correspondence. Due to their topological nature, their existence is robust against small perturbations, like disorder. This point is further discussed in section 7.

The phase diagram constructed using the winding number invariant is shown in figure 1. The meaning of two different values for the winding number is clearer when we recall the Kitaev Hamiltonian in the Majorana basis. In a finite chain the leftmost lattice site consists of the A Majorana operator ${\gamma }_{1}^{A}$ connected to the bulk by the i(Δ − t) hopping and the B Majorana operator ${\gamma }_{1}^{B}$ connected by the i(Δ + t) hopping. With Δ > 0 and t < 0 (the ν = +1 phase) the Majorana state at the left end of the chain will consist mostly of the weakly connected ${\gamma }_{1}^{B}$. If t > 0 (the ν = −1 phase), ${\gamma }_{1}^{A}$ is connected to the bulk more weakly and contributes most to the left end bound state.

Figure 1.

Figure 1. Topological phase diagram of the Kitaev chain for Δ > 0, constructed with the winding number invariant equation (8). Distinct topological phases are separated by the phase boundary at μ = ±2t, visualised by the red lines. The four insets illustrate the evolution of w(k) along the Brillouin zone at the four marked positions in the phase space, (t/Δ = ±2, μ/Δ = 0.2) in the topological and (t/Δ = ±2, μ/Δ = 4.2) in the trivial phase. In the phase diagram for Δ < 0 the ν = +1 and ν = −1 regions are swapped.

Standard image High-resolution image

The boundaries between different topological phases can be obtained from the condition of closing the bulk gap, i.e. E±(k) = 0 (cf equation (7)). That is only possible if both terms under the square root vanish. The condition of Δ ≠ 0 forces the gap closing to occur at kd = 0 or k = πd, and the remaining term vanishes at these momenta if μ = ±2t. The four insets in figure 1 show the behavior of w(k), leading to either a zero (for w(−π) = w(π)) or non zero winding number, see equation (8).

Physically speaking, the Kitaev chain is in the topological phase provided that Δ ≠ 0 and the chemical potential lies inside the 'normal' band (|μ| ⩽ 2|t|).

3. Spectral analysis of the finite Kitaev chain

One of the characteristics of finite systems is the possibility to host edge states at their ends. To account for the presence and the nature of such edge states, we consider a finite Kitaev chain with N sites and open boundary conditions, yielding N allowed k values. In this section we shall consider the situation in which one of the three parameters Δ, t and μ is zero. Already for the simple case μ = 0 and Δ ≠ 0, t ≠ 0 the quantization of the momentum turns out to be non trivial. The general case in which all parameters are finite is considered in sections 57.

We start with the BdG Hamiltonian of the open Kitaev chain in real space, expressed in the basis of standard fermionic operators $\hat{\psi }={\left({d}_{1},\dots ,\;{d}_{N},{d}_{1}^{{\dagger}},\dots ,\;{d}_{N}^{{\dagger}}\right)}^{\text{T}}$. Then

Equation (9)

where the BdG Hamiltonian ${\mathcal{H}}_{\text{KC}}$ is

Equation (10)

These matrices have the tridiagonal structure

Equation (11)

Equation (12)

The spectrum can be obtained by diagonalisation of ${\mathcal{H}}_{\text{KC}}$ in real space. We consider different situations.

3.1. Δ = 0

The BdG Hamiltonian is block diagonal and its characteristic polynomial ${P}_{\lambda }\left({\mathcal{H}}_{\text{KC}}\right)=\mathrm{det}\left[\lambda \;\mathbb{1}-{\mathcal{H}}_{\text{KC}}\right]$ factorises as

Equation (13)

The tridiagonal structure of C straightforwardly yields the spectrum of a normal conducting, linear chain [33]

Equation (14)

where n runs from 1 to N. Since ${k}_{n}\in \mathbb{R}$, only bulk states exist for Δ = 0.

3.2. t = 0

In the beginning we consider both t and μ to be zero and include μ ≠ 0 in a second step. The parameter setting leads to a vanishing matrix C, see equation (11), and the characteristic polynomial of the system reads:

Equation (15)

where we used the property S = −S. Due to the fact that the commutator $\left[\mathbb{1},\;S\right]=0$ vanishes4, one finds [34]

Equation (16)

The characteristic polynomial can still be simplified to the product

Equation (17)

The matrix iS is hermitian and describes a linear chain with hopping iΔ. As a consequence, we find the spectrum to be [33]

Equation (18)

where n runs from 1 to N and each eigenvalue is twice degenerated. Notice the phase shift by π/2 compared to the spectrum of an infinite chain equation (7). We discuss this phase shift in more detail in section 3.3.

Furthermore if, and only if, N is odd, we find two zero energy modes, namely for n = (N + 1)/2. Their existence for odd N and the degeneracy is due to the chiral symmetry.

The chemical potential μ can be included easily. Exploiting the properties of ${\mathcal{H}}_{\text{KC}}$, we find the characteristic polynomial to be

Equation (19)

with Λ2 := λ2μ2. The same treatment as in the previous μ = 0 case yields ${P}_{\lambda }\left({\mathcal{H}}_{\text{KC}}\right)={P}_{{\Lambda}}\left(iS\right)\;{P}_{{\Lambda}}\left(-iS\right)$. Consequently the spectrum is

Equation (20)

where n runs again from 1 to N. Again no boundary modes are found for t = 0.

3.3. μ = 0

The calculation of the spectrum for μ = 0 requires a more technical approach, since the structure of the BdG Hamiltonian equation (10) prohibits standard methods.

One important feature of the Kitaev chain can be appreciated inspecting equation (3). The entire model is equivalent to two coupled SSH-like chains [30, 35] containing both the hopping parameters $a\;{:=}\;i\left({\Delta}-t\right)$ and $b\;{:=}\;i\left({\Delta}+t\right)$, see figure 2. Explicitly,

Equation (21)

where N1,2 depend on N. If N is even we have N1 = N/2 and N2 = N1 − 1, while N1 = N2 = (N − 1)/2 for odd N. Independent of the number of atoms, the first and the second lines in equation (21) describe two SSH-like chains, coupled by the chemical potential μ. We define here the SSH-like basis of the Kitaev chain as:

Equation (22)

where '|' marks the boundary between both chains. We call the first one, starting always with ${\gamma }_{1}^{A}$, the α chain, and the second the β chain, such that ${{\Psi}}_{\text{SSH}}^{\text{even,odd}}={\left({\to {\gamma }}_{\alpha }\;\vert \;{\to {\gamma }}_{\beta }\right)}^{\text{T}}$. The BdG Hamiltonian in the SSH-like basis reads

Equation (23)

with ${\hat{H}}_{\text{KC}}=\frac{1}{2}{\hat{{\Psi}}}_{\text{SSH}}^{{\dagger}}{\mathcal{H}}_{\text{KC}}^{\text{SSH}}{\hat{{\Psi}}}_{\text{SSH}}$. The independent SSH-like chains are represented by the square matrices ${\mathcal{H}}_{\alpha }$ and ${\mathcal{H}}_{\beta }$ of size N. Both chains are coupled by the matrices τ and τ, which contain only the chemical potential μ, in a diagonal arrangement specified below.

Figure 2.

Figure 2. Kitaev chain viewed as two coupled SSH-like chains for (a) N = 4 and (b) N = 3 sites. These two chains α and β are coupled by ±. The hoppings a = i(Δ − t) in red and b = i(Δ + t) in blue alternate (dashed lines correspond to −a and −b) and connect neighbouring Majorana operators ${\gamma }_{j}^{A}$ (blue spheres) and ${\gamma }_{j{\pm}1}^{B}$ (orange spheres). The unit cell has size 2d.

Standard image High-resolution image

The pattern of these matrices is slightly different for even and odd number of sites. If N is even we find

Equation (24)

Equation (25)

and ${\tau }^{\text{even}}=-i\mu \;{\mathbb{1}}_{N/2}\otimes {\tau }_{z}$, where τz denotes the Pauli matrix. The odd N expressions are achieved by removing the last line and column in ${\mathcal{H}}_{\alpha }^{\text{even}}$, ${\mathcal{H}}_{\beta }^{\text{even}}$ and τeven.

As shown in more detail in appendix B, for μ = 0 the characteristic polynomial can be expressed as the product of two polynomials of order N

Equation (26)

where the product form reflects the fact that the Kitaev chain is given in terms of two uncoupled SSH-like chains, as illustrated in figure 2. Even though the polynomials ζN and epsilonN belong to different SSH-like chains, both obey a common recursion formula typical of Fibonacci polynomials [3638]

Equation (27)

and differ only in their initial values

Equation (28)

Fundamental properties of Fibonacci polynomials are summarized in appendix A. The common sublattice structure of both chains sets the stage for a relationship between ζj and epsilonj: the exchange of a's and b's enables us to pass from one to the other

Equation (29)

Moreover, equation (27) implies that a Kitaev chain with even number of sites N is fundamentally different from the one with an odd number of sites. This property is a known feature of SSH chains [39]. The difference emerges since, according to equations (B20), (B22), it holds

Equation (30)

because the number of a and b type bondings in both subchains is the same. This leads to twice degenerate eigenvalues. An equivalent relationship for even N does not exist. The closed form for ζj and epsilonj, as well as their factorization, is derived in appendix B.

The characteristic polynomial can be used to obtain the determinant of the Kitaev chain, here for μ = 0, because evaluating it at λ = 0 leads to:

According to equation (26) we need only to know ζN and epsilonN at λ = 0. The closed form expression for ζj at λ = 0 reduces to

Equation (31)

while epsilonj|λ=0. follows from equation (29). We find that there are always zero energy eigenvalues for odd N, but not in general for even N, as it follows from

Equation (32)

Additional features of the spectrum are discussed in the following.

3.3.1. Odd N

The spectrum for odd N is given by two contributions

Equation (33)

Equation (34)

where knd = /(N + 1) and n runs from 1 to N, except for n = (N + 1)/2. This constraint on n is a consequence of the equations (67), (69) below which show that the boundary condition equation (70) cannot be satisfied for kd = π/2. Hence, no standing wave can be formed.

Each zero eigenvalue belongs to one chain. As discussed below, two decaying states are associated to equation (33), whose wave functions are discussed in section 4.2. These states are MZM.

3.3.2. Even N

In the situation of even N we find for the Kitaev's bulk spectrum at zero μ

Equation (35)

where the momenta k are in general not equidistant in the first Brillouin zone. Rather, the bulk quantization condition follows from the interplay between Δ and t and is captured in form of the functions fβ,α(k) (cf appendix B.2),

Equation (36)

whose zeros

Equation (37)

define the allowed values of k. Note that kd = 0, π/2 are excluded as solutions, due to their trivial character. The functions fβ,α(k) follow from the factorisation of the polynomials epsilonN and ζN. The negative sign in equation (36) belongs to the α subchain, while the positive one to the β subchain. The spectrum following from equation (37) is illustrated in figure 3. We observe that equations (35) and (37) hold for all values of t and Δ, independent of whether |Δ| is larger or smaller than |t|. The two situations are connected by a phase shift of the momentum kdkd + π/2, which influences both the spectrum and the quantization condition. In the end all different ratios of Δ and t are captured by equations (35) and (37), due to the periodicity of the spectrum.

Figure 3.

Figure 3. Eigenvalues and the non equidistant quantization of the bulk momentum k for a Kitaev molecule with four sites. (a) The horizontal lines mark the numerical eigenvalues ±Ej (j = 0, 1, 2, 3) and the bulk spectrum of the infinite chain is the green solid curve. The tangent-like functions follow fβ(k) and fα(k) from equation (36). The zeros of fβ,α(k) define the proper wave vectors k3,2,1 of the finite system and these cut the dispersion relation at the correct positions, such that E±(kj) = ±Ej. (b) Zoom of (a) for $k\in \left[0.5/d,\;1.2/d\right]$. The chosen parameters N = 4, t = 4 meV, Δ = 1.5 meV and μ = 0 meV lead to the bulk eigenvalues ${\pm}{E}_{j}\in \left[{\pm}4.39,\;{\pm}6.47,\;{\pm}6.89\right]$ (in meV) and to the momenta k3,2,1 approximately [0.58012/d, 0.68813/d, 1.12386/d].

Standard image High-resolution image

However, when we consider decaying or edge states this periodicity is lost (see equations (40) and (41) below) and |t| ≶ |Δ| lead to different quantization rules. The hermiticity of the Hamiltonian allows a pure imaginary momentum for μ = 0, but a simple exchange of k to iq in equation (36) does not lead to the correct results. We introduce here the functions

Equation (38)

similar to fβ,α(k) in equation (36), where m contains both ratios of Δ and t:

Equation (39)

Again, the positive sign in equation (38) belongs to the β chain and the negative one to the α chain. The exact quantization criterion is provided by the zeros of hβ,α(q),

Equation (40)

as illustrated in figure 4. The associated energies follow from the dispersion relation

Equation (41)

We notice that equation (41) is only well defined for zero or positive arguments of the square root. Indeed, all solutions of equation (40), if existent, lie always inside this range, because using hβ,α(q) = 0 in equation (41) yields

Equation (42)

Hence, each wavevector from equation (40) corresponds to two gap modes, since the gap width is $4\mathrm{min}\left\{\vert {\Delta}\vert ,\vert t\vert \right\}$ and the fraction inside equation (42) is always smaller than one. We can restrict ourselves to find only positive solutions qd, due to the time reversal symmetry. The number of physically different solutions of equation (40) is zero or two and it follows always from the equation containing the positive factor m or −m. Consequently, according to equation (38), only none or two gap modes can form and both belong to the same subchain, α or β. Moreover a solution exists if, and only if, $\vert m\vert \in \left[1,\;N+1\right]$.

Figure 4.

Figure 4. Eigenvalues and the quantised momentum q0 of the gap modes for a Kitaev molecule with four sites. (a) The horizontal lines characterise the numerical eigenvalues ±Ej (j = 0, 1, 2, 3) and the dispersion relation inside the gap (the red ellipse) is shown as function of continuous q on a finite range. Only one of both hyperbolic tangent-like functions hβ(q) or hα(q) defines a proper qd ≠ 0. (b) Zoom of (a) for $q\in \left[0.3/d,\;0.42/d\right]$. The momenta ±q0, the zeros of hα, lead to the correct associated energies, such that E±(q0) = ±E0. The chosen parameters N = 4, t = 4 meV, Δ = 1.5 meV and μ = 0 meV lead to the eigenvalues ${E}_{j,{\pm}}\in \left[{\pm}0.97,\;{\pm}4.39,\;{\pm}6.47,\;{\pm}6.89\right]$ (in meV) and to the momentum q0 ≈ 0.374 16/d.

Standard image High-resolution image

In the limiting case when |m| → 1, i.e. at the Kitaev points, the solution qd and the associated energies E± from equation (42) go to zero. The eigenstate will be a Majorana zero energy mode, see section 4.1.

In the second special case of |m| → N + 1 the solution approaches zero. The value q = 0 is only in this particular scenario a proper momentum, see appendix B.2. The momentum q = 0 yields the energies ${E}_{{\pm}}\left(0\right)={\pm}2\mathrm{min}\left\{\vert {\Delta}\vert ,\vert t\vert \right\}$, which mark exactly the gap boundaries.

Increasing the value of |m| beyond N + 1 entails the absence of imaginary solutions. The number of eigenvalues of a Kitaev chain is still 2N for a fixed number of sites and consequently equation (37) leads now to N real values for kd, instead of N − 1. In other words, the two former gap modes have moved to two extended states and their energy lies now within the bulk region of the spectrum, even though the system is still fully gaped. This effect holds for the Kitaev chain as well as for SSH chains. Physically this means, that a 'boundary' mode with imaginary momentum q and corresponding decay length ξ ∝ 1/q reached the highest possible delocalisation in the chain.

The limit of N yields always two zero energy boundary modes; since the momentum is qd = arctanh(|1/m|), due to equations (38), (40) and according to equation (42) the energy goes to zero. If we consider the odd N situation in the limit of an infinite number of sites, we have there two zero energy boundary modes as well. The results of this section are summarized in table 1.

Table 1. Overview of the quantization rule for the wave vectors of the finite Kitaev chain in different scenarios. The wavevectors k, kn, κ1,2 used together with equation (7) and q with equation (41) yield the correct finite system energies and $k,\;{k}_{n},q\in \mathbb{R}$, ${\kappa }_{1,2}\in \mathbb{C}$. Notice: n = 1, ..., N and $m=\frac{t}{{\Delta}}$ ($m=\frac{{\Delta}}{t}$) for |t| ⩾ |Δ| (|Δ| > |t|).

RequirementsQuantisation ruleZero modesEquation for eigenstate elementsMajorana character
Δ = 0:${k}_{n}d=\frac{n\pi }{N+1}$Yes, if for some n: No
   $\mu ={\mu }_{n}=2t\;\mathrm{cos}\left(\frac{n\pi }{N+1}\right)$ 
t = 0:${k}_{n}d=\frac{\pi }{2}+\frac{n\pi }{N+1}$No No
μ = 0, N odd:${k}_{n}d=\frac{n\pi }{N+1},n\ne \frac{N+1}{2}$No(67)–(69)No
$qd=\mathrm{a}\mathrm{r}\mathrm{c}\mathrm{t}\mathrm{a}\mathrm{n}\mathrm{h}\left(1/\vert m\vert \right)$Yes(71), (72)Yes
   (75)–(78)Yes
μ = 0, N even:$\mathrm{tan}\left[kd\left(N+1\right)\right]=\mp \frac{{\Delta}}{t}\mathrm{tan}\left(kd\right)$No(51), (52)No
$\mathrm{tanh}\left[qd\left(N+1\right)\right]=\mp m\;\mathrm{tanh}\left(qd\right)$Only if Δ = ±t(61)–(64)Yes
  Otherwise(51), (52)No
$t,\;{\Delta},\;\mu \in \mathbb{R}$$\frac{{\mathrm{sin}}^{2}\left[\frac{{\kappa }_{1}+{\kappa }_{2}}{2}\left(N+1\right)\right]}{{\mathrm{sin}}^{2}\left[\frac{{\kappa }_{1}-{\kappa }_{2}}{2}\left(N+1\right)\right]}=\frac{1+{\left(\frac{{\Delta}}{t}\right)}^{2}\;{\mathrm{cot}}^{2}\left(\frac{{\kappa }_{1}-{\kappa }_{2}}{2}\right)}{1+{\left(\frac{{\Delta}}{t}\right)}^{2}\;{\mathrm{cot}}^{2}\left(\frac{{\kappa }_{1}+{\kappa }_{2}}{2}\right)}$only for $\mu ={\mu }_{n}\in \mathbb{R}$ and (100), (102)No
  ${\mu }_{n}=2\sqrt{{t}^{2}-{{\Delta}}^{2}}\mathrm{cos}\left(\frac{n\pi }{N+1}\right)$(109), (110)Yes
   (112), (113)Yes

4. Eigenvectors (μ = 0)

We use the SSH-like basis to calculate the eigenvectors of the Hamiltonian equation (23) at μ = 0. The eigenvectors $\to {\psi }$ are defined with respect to the SSH-like chains α and β, see equation (23),

Equation (43)

with the feature that always either ${\to {v}}_{\beta }$ or ${\to {v}}_{\alpha }$ can be chosen to be zero, yielding the solutions ${\to {\psi }}^{\alpha }$ and ${\to {\psi }}^{\beta }$, respectively

Equation (44)

We are left to find the eigenvectors of a single tridiagonal matrix which we did basing on, and extending the results of [40]. We focus here on the edge and decaying states, while the rest of our results is in appendix C. Remember that in the SSH-like basis equation (22) the Majorana operators ${\gamma }_{j}^{A}$ and ${\gamma }_{j}^{B}$, alternate at each site, thus defining two interpenetrating 'A' and 'B' type sublattices.

4.1. Even N

We define the vectors ${\to {v}}_{\alpha }$ and ${\to {v}}_{\beta }$ via the entries

Equation (45)

Equation (46)

where x, y and $\mathcal{X},\;\mathcal{Y}$ are associated to the A and B sublattices, respectively. The internal structure of ${\to {v}}_{\alpha }$ (${\to {v}}_{\beta }$) reflects the unit cell of an SSH-like chain and thus simplifies the calculation. In the real space xl (${\mathcal{X}}_{l}$) belongs to site j = 2l − 1 and yl (${\mathcal{Y}}_{l}$) to j = 2l, where j = 1, ..., N.

Searching for solutions on the subchain α implies setting ${\to {v}}_{\beta }=\to {0}$ and solving $\left({\mathcal{H}}_{\alpha }^{\text{even}}-{E}_{{\pm}}{\mathbb{1}}_{N}\right){\to {v}}_{\alpha }=\to {0}$. The elements of ${\to {v}}_{\alpha }$ obey

Equation (47)

Equation (48)

and

Equation (49)

Equation (50)

where l runs from 1 to N/2 − 1. The solution for Δ ≠ ±t is (in agreement with [40])

Equation (51)

Equation (52)

where l = 1, ..., N/2, θ/(2d) denotes the momentum k (q) for extended (gap) states and E± is the dispersion relation associated to k (equation (35)), or q (equation (41)). The entries of the eigenvectors are essentially sine functions for the extended states

Equation (53)

and hyperbolic sine functions for the decaying states

Equation (54)

where the prefactor s depends on the ratio of Δ and t:

An illustration of ${\to {\psi }}^{\alpha }$ is given in figure 5. The allowed momenta k or q follow from the open boundary conditions

Equation (55)

The first condition is satisfied due to T0(θ) = 0 for any momentum. The second condition yields the quantization rules fα(k) = 0 and hα(q) = 0 for the α chain, see equations (37) and (40).

The eigenvector ${\to {\psi }}^{\beta }$ entails ${\to {v}}_{\alpha }=0$ and the entries of ${\to {v}}_{\beta }$ follow essentially by replacing a's and b's in the equations (51) and (52). We find

Equation (56)

Equation (57)

where j = 1, ..., N/2 and Δ ≠ ±t. The quantisation condition follows from the open boundary condition:

and k (q) obey fβ(k) = 0 (hβ(q) = 0). Further, from the quantization rules it follows that gap modes belong always to the same subchain α or β for even N.

Figure 5.

Figure 5. Visualisation of the entries of the eigenstates ${\to {\psi }}_{q,k}^{\alpha }\left(j\right)$ of the Kitaev chain with N = 42 sites and μ = 0. Panel (a) depicts the gap state ${\to {\psi }}_{q}^{\alpha }$ and (b) the lowest energy bulk state ${\to {\psi }}_{k}^{\alpha }$. The blue (orange) dots follow xl/x1 (iyl/x1) at position j = 2l − 1 (j = 2l) for l = 1, ..., N/2, while the black line is only a guide to the eye. The gap state is more localised at the edges. The extended state is largest inside the chain. The chosen parameters are t = 10 meV and Δ = 1 meV leading to q = 0.100 29/d and E = 0.0539 meV for the gap state. The shown extended state is associated with k = 1.4806/d and E = 2.6851 meV. Notice the decaying state in a) as well as the ones depicted in figures 6(a) and 7 are not Majorana states.

Standard image High-resolution image

As illustrated in figures 57 our states are symmetric w.r.t. the center of the SSH-like chains. This symmetry is visible in alternative versions of the equations (52) and (57) (l = 1, ..., N/2), whereby ${\mathit{x}}_{\frac{N}{2}+1-l}={T}_{l}^{\text{e}}\left(\theta \right)\;{\mathit{x}}_{\frac{N}{2}}$, which holds for all eigenstates. Together with equations (51), we find in general

Equation (58)

and similarly ${\mathcal{X}}_{\frac{N}{2}+1-l}={\mathcal{Y}}_{l}\;{\mathcal{X}}_{\frac{N}{2}}/{\mathcal{Y}}_{1}$. Recalling the definition of the SSH-like basis, equation (22), and introducing the operators ${\psi }_{\alpha ,\beta }^{{\dagger}}$ associated to the states ${\to {\psi }}^{\alpha ,\beta }$ in equation (44), we find the expression

Equation (59)

where vα is the norm of the vector ${\to {v}}_{\alpha }$. A similar term is found for ${\psi }_{\beta }^{{\dagger}}$. We notice that equations (59) and (60) below are true for all kinds of eigenstates, i.e. extended, decaying states and MZM, of the BdG Hamiltonian in equation (23) at μ = 0. The character (statistics) of the operators depends on whether ${\left({\psi }_{\alpha ,\beta }^{{\dagger}}\right)}^{2}$ is 0 or 1/2. The property $\left\{{\gamma }_{j}^{r},{\gamma }_{k}^{s}\right\}={\delta }_{j,k}{\delta }_{r,s}$ with (r, s ∈ {A, B}) yields

Equation (60)

The symmetry in equations (58) states that ${\left({\psi }_{\alpha }^{{\dagger}}\right)}^{2}$ is essentially determined by ${\left({x}_{N/2}/{y}_{1}\right)}^{2}$. For a ≠ 0 and consequently E ≠ 0, we find from equations (47), (48) and (58) that ${\left({x}_{N/2}/{y}_{1}\right)}^{2}=-1$, which yields ${\left({\psi }_{\alpha }^{{\dagger}}\right)}^{2}=0$. Thus the operators associated to the finite energy states ${\to {\psi }}^{\alpha }$, including the ones depicted in figures 5(a) and 6(a), obey fermionic statistics. This result holds also true in the case a = 0 and E ≠ 0 as can be seen by using the corresponding eigenstates (appendix C). Similar results hold for ${\left({\psi }_{\beta }^{{\dagger}}\right)}^{2}$.

Figure 6.

Figure 6. Illustration of the entries of the eigenstates ${\to {\psi }}_{q,k}^{\alpha }\left(j\right)$ of a Kitaev chain for N = 42 sites and μ = 0 for t = 5 meV and Δ = 1 meV. Similar to figure 5, but with a modified value for t/Δ. (a) Shows the gap mode and (b) the lowest in energy bulk mode. Notice that for the chosen parameter set the gap state is more localized than the one in figure 5. In contrast the extended state has lower weight at the ends of the chain. The gap mode (bulk state) is associated with q = 0.2027/d (k = 1.4886/d) and E = 0.6682 × 10−3 meV (E = 2.1555 meV).

Standard image High-resolution image

We turn now to Majorana zero modes, which at μ = 0 only exist at the Kitaev points Δ = ±t, see equation (32).

When Δ = t we find two zero energy modes ${\to {\psi }}_{A}^{\alpha }=\left(\begin{pmatrix}\hfill {\to {v}}_{\alpha ,A}\hfill \\ \to {0}\hfill \end{pmatrix}\right)$, ${\to {\psi }}_{B}^{\alpha }=\left(\begin{pmatrix}\hfill {\to {v}}_{\alpha ,B}\hfill \\ \to {0}\hfill \end{pmatrix}\right)$ each localised at one end of the α chain:

Equation (61)

Equation (62)

and ${\left({\psi }_{\alpha }^{{\dagger}}\right)}^{2}=1/2$ in equation (60). In contrast, both zero energy modes are on the β chain for Δ = −t. We find ${\to {\psi }}_{A}^{\beta }=\left(\begin{pmatrix}\to {0}\hfill \\ \hfill {\to {v}}_{\beta ,A}\hfill \end{pmatrix}\right)$, ${\to {\psi }}_{B}^{\beta }=\left(\begin{pmatrix}\to {0}\hfill \\ \hfill {\to {v}}_{\beta ,B}\hfill \end{pmatrix}\right)$ with

Equation (63)

Equation (64)

These states are the archetypal Majorana zero modes [1, 2]. Due to their degeneracy, these modes can be recombined into fermionic quasiparticles by appropriate linear combination, see in equations (4a) and (4b) from section 2.1.

4.2. Odd N

The composition of the eigenvectors slightly changes for the odd case compared to the even N case

Equation (65)

Equation (66)

Although both odd sized chains share the same spectrum, it is possible to find a linear combination of states which belongs to one chain only. The form of the extended states of the odd chains (Δ ≠ ±t and E± ≠ 0) does not differ much from the one of the even chain and the entries of ${\to {v}}_{\alpha }$ are

Equation (67)

Equation (68)

where ${T}_{l}^{\text{o}}$ is

Equation (69)

with knd = /(N + 1) (n = 1, ..., N, n ≠ (N + 1)/2). The exchange of a's and b's leads again to the coefficients for the chain β (see appendix C).

The significant difference between even and odd N lies in the realization of the open boundary condition. Solving $\left({\mathcal{H}}_{\alpha }^{\text{odd}}-{E}_{{\pm}}{\mathbb{1}}_{N}\right){\to {v}}_{\alpha }=\to {0}$ yields now

Equation (70)

which leads to the momenta kn.

Figure 7.

Figure 7. The decaying state ${\to {\psi }}_{q}^{\beta }$ for N = 42 sites and μ = 0. The black guiding line follows the orange (blue) dots, which correspond to ${\mathcal{X}}_{l}/{\mathcal{X}}_{1}$ ($i\;{\mathcal{Y}}_{l}/{\mathcal{X}}_{1}$) at position j = 2l − 1 (j = 2l), l = 1, ..., N/2. The difference to the edge states on subchain α is the exchanged role of Majorana operators ${\gamma }_{j}^{A}$, ${\gamma }_{j}^{B}$. The chosen parameters are t = −5 meV and Δ = 1 meV. The gap mode is associated with q = 0.2027/d and E = 0.6682 × 10−3 meV.

Standard image High-resolution image

An SSH-like chain with an odd number of sites hosts only a single zero energy mode, but α and β contribute each with one. We find on subchain α for Δ ≠ ±t

Equation (71)

and on subchain β

Equation (72)

where l runs from 1 to (N + 1)/2.

Regarding the statistics of the operators ${\psi }_{\alpha }^{{\dagger}},\;{\psi }_{\beta }^{{\dagger}}$ associated to the states ${\to {\psi }}^{\alpha },\;{\to {\psi }}^{\beta }$, we proceed like for the even N case. The use of the SSH-like basis from equation (22) and the entries of the state ${\to {\psi }}^{\alpha }$ yield now

Again, the equations (67), (68) and (70) imply a perfect compensation of the A and B sublattice contributions, yielding ${\left({\psi }_{\alpha }^{{\dagger}}\right)}^{2}=0$ for E ≠ 0. The zero energy mode, given by its entries in equation (71), leads to ${\left({\psi }_{\alpha }^{{\dagger}}\right)}^{2}=1/2$.

Further, we find that both zero energy modes ${\to {\psi }}^{\alpha ,\beta }$ have their maximum at opposite ends of the Kitaev chain and decay into the chain. To better visualize this it is convenient to introduce the decay length

Equation (73)

and remembering that the atomic site index of xl is j = 2l − 1 equation (71) yields for |t| ⩾ |Δ|

Equation (74)

For |t| ⩽ |Δ| the xl coefficients are given by the same equation without the (−1)l−1 factor. We have moreover q = ±1/ξ, where q is the imaginary momentum yielding E = 0 in equation (41). Thus the localisation of these states is determined only by t and Δ. In the parameter setting of Δ = t we find:

Equation (75)

Equation (76)

while both states exchange their position for Δ = −t

Equation (77)

Equation (78)

4.3. The particle–hole-operator

In the last section we have shown that some of the zero energy eigenstates of the BdG Hamiltonian of the finite Kitaev chain are Majorana zero modes (MZM) by exploiting the statistics of the corresponding operators ψα,β. We further corroborate this statement now by recalling that an MZM is defined as an eigenstate of the Hamiltonian $\mathcal{H}$ and of the particle hole symmetry $\mathcal{P}$. The latter acts on an eigenstate ${\to {\psi }}^{\alpha ,\beta }$ of energy E by turning it into an eigenstate of $\mathcal{H}$ of energy −E. Thus, the energy of such an exotic state has to be zero, since eigenstates associated to different energies are orthogonal.

The three symmetries, time reversal, chiral and the particle–hole symmetry, discussed in section 2.3, can be constructed in real space too. Of particular interest is their representation in the SSH-like basis. The antiunitary particle–hole symmetry is

Equation (79)

The time reversal and the chiral symmetry depend on N. If N is even we find

Equation (80)

Equation (81)

The expressions for odd N follow by removing the last line and last column in each diagonal block.

The effect of $\mathcal{P}$ from equation (79) can be seen explicitly if one considers $\mathcal{P}\;{\to {\psi }}^{\alpha }$. For N even and Δ ≠ t the elements xl, yl of ${\to {\psi }}^{\alpha }$ are given in equations (51) and (52). Here yl/x1 is pure imaginary and xl/x1 is real. Hence, ${\to {\psi }}^{\alpha }$ is not an eigenstate of $\mathcal{P}$ since the prefactor to yl is finite, i.e., E± ≠ 0. We conclude that in a finite Kitaev chain with even number of sites Majorana zero modes emerge only at the Kitaev points for μ = 0, since the states in equations (61)–(64) are eigenstates of $\mathcal{P}$ as well. In the situation of odd N and μ = 0, the eigenstates given by their elements in equations (71), (72) are Majorana zero energy modes for an appropriate choice of x1. These states can be delocalised over the entire chain, depending on their decay length ξ, while the case of full localisation is only reached at the the Kitaev points, where the MZM turn into the states given by equations (75)–(78).

5. Results for the spectrum and eigenstates at finite μ

5.1. Spectrum

The last missing situation is to consider a finite chemical potential μ. For this purpose we use the so called chiral basis ${\hat{{\Psi}}}_{c}{:=}{\left({\gamma }_{1}^{A},\;{\gamma }_{2}^{A},\dots ,\;{\gamma }_{N}^{A},\;{\gamma }_{1}^{B},\;{\gamma }_{2}^{B},\dots ,\;{\gamma }_{N}^{B}\right)}^{\text{T}}$. The Kitaev Hamiltonian transforms via ${\hat{H}}_{\text{KC}}=\frac{1}{2}\;{\hat{{\Psi}}}_{c}^{{\dagger}}\;{\mathcal{H}}_{c}\;{\hat{{\Psi}}}_{c}$, into a block off-diagonal matrix

Equation (82)

because there are no ${\gamma }_{j}^{A}\;{\gamma }_{i}^{A}$ (${\gamma }_{j}^{B}\;{\gamma }_{i}^{B}$) contributions in equation (21). The N × N matrix h is tridiagonal

Equation (83)

since the Kitaev Hamiltonian contains only nearest neighbour hoppings. Then the characteristic polynomial is [34]

Equation (84)

where, however, h and h do not commute except for t = 0 or Δ = 0. Thus, such matrices cannot be diagonalised simultaneously. Nevertheless the eigenvalues ηj (${\eta }_{j}^{{\ast}}$) of h (h) are easily derived e.g. following [33]. We find

Equation (85)

independent of whether Δ ⩾ t or t > Δ.

5.1.1. Condition for zero energy modes

Equation (85) immediately yields the criterion for hosting zero energy modes. According to equation (82), we have

Equation (86)

and we need only to focus on det(h).

If a single eigenvalue ηj of h is zero then det(h) vanishes. Thus, for a zero energy mode the chemical potential must satisfy

Equation (87)

Obviously, equation (87) cannot be satisfied for generic values of t2 − Δ2, because all other quantities are real. The only possibility is t2 ⩾ Δ2. There is only one exception for odd N, because the value n = (N + 1)/2 leads to μ = 0 in equation (87) for all values of t and Δ, in agreement with our results of section 3. This result is exact and confirms findings from [21, 22]; further it improves a similar but approximate condition on the chemical potential discussed by Zvyagin in [23].

An illustration of these discrete solutions μn, which we dub 'Majorana lines', is shown in figure 8. All paths contain the Kitaev points at μ = 0 and t = ±Δ. Further, their density is larger close to the boundary of the topological phase, as a result of the slow changes of the cosine function around 0 and π.

Figure 8.

Figure 8. Existence of zero energy solutions for growing system sizes. The red lines mark the boundaries of the topological phase diagram. Zero energy solutions correspond to the blue curves, along which the determinant of the Kitaev Hamiltonian vanishes as a function of μ, t and Δ. (a) The shown situation is for a small even N, yielding a zero determinant on the horizontal axis only at the Kitaev points t/Δ = ±1. Each blue curve departs from one of these two points. (b) The situation of small odd N is similar to the even one, but the entire μ = 0-axis is now included. (c) The solutions μn become dense for larger even N, and one sees already the filling of the non trivial phase for N. (d) Large even N behave similar to large odd N, but the latter still include the entire horizontal axis.

Standard image High-resolution image

For growing number of sites N, the density of solutions increases. In the limit N, θn = /N + 1 takes all values in [0, π] and the entire area between $\mu ={\pm}2\;\sqrt{{t}^{2}-{{\Delta}}^{2}}$ for t2 ⩾ Δ2 is now occupied with zero energy modes.

Regarding the remaining part of the topological region, we are going to show in the next section that in that parameter space only boundary modes with finite energy exist. Because the energy of these modes is decreasing exponentially with the system size, in the thermodynamic limit their energy approaches zero and the full topological region supports zero energy modes.

5.1.2. The complete spectrum of the finite Kitaev chain

To proceed we transform equation (84) into an eigenvector problem for hh,

Equation (88)

where we defined $\to {v}={\left({\xi }_{1},\;{\xi }_{2},\dots ,\;{\xi }_{N}\right)}^{\text{T}}$. Notice that we are not really interested in the eigenvector $\to {v}$ here; we simply use its entries as dummy variables to release a structure hidden in the product of h and h. The elements of h,

where n, m = 1, ..., N, allow us to calculate the product hh entry wise

Thus, importantly, equation (88) reveals a recursion formula

Equation (89)

for the components of $\to {v}$. The entries ξ are a generalisation of the Fibonacci polynomials ζj from equation (27), to which they reduce for μ = 0, and may be called Tetranacci polynomials [41, 42]. Further, we find the open boundary conditions from equation (88) to be

Equation (90)

where we used equation (89) for simplifications.

Appendix D contains the description of how to deal with those polynomials, the boundary conditions and further the connection of equation (89) to Kitaev's bulk spectrum λ = E±(k) in equation (7). Essentially one has to use similar techniques as it was done for the Fibonacci polynomials, where now the power law ansatz ξjrj leads to a characteristic equation for r of order four. Thus, we find in total four linearly independent fundamental solutions r±1,±2, which can be expressed in terms of two complex wavevectors denoted by κ1,2 through the equality

Equation (91)

These wavevectors are not independent, but coupled via

Equation (92)

For μ = 0 we can recover from equation (92) our previous results, whereby one has only pure real (k) or pure imaginary (iq) wavevectors5. Further, equations (7) and (92) yield

The linearity of the recursion formula equation (89) states that the superposition of all four fundamental solutions is the general form of ξj. Since the boundary conditions translate into a homogeneous system of four coupled equations and a trivial solution for ξj has to be avoided, we find that the determinant of the matrix describing these equations has to be zero. After some algebraic manipulations, this procedure leads finally to the full quantization rule of the Kitaev chain

Equation (93)

where we introduced the function F(κ1, κ2) as

Equation (94)

Similar quantization conditions are known for an open X–Y spin chain in transverse field [26]. Notice that the quantization rule is symmetric with respect to κ1,2. Table 1 gives an overview of the quantization rules for different parameter settings $\left({\Delta},\;t,\;\mu \right)$. The bulk eigenvalues of a finite Kitaev chain with four sites and μ ≠ 0 are shown in figure 9.

Figure 9.

Figure 9. Spectrum of the Kitaev molecule with four sites and μ ≠ 0. The green line follows the excitation spectrum from equation (7) and the horizontal lines are the numerical eigenvalues of the Kitaev chain. The momenta k3,2,1 are the proper wavevectors for μ ≠ 0 calculated from the full quantization rule, see equations (92)–(94). The dashed, light gray lines represent the wavevectors taken from the μ = 0 case to highlight the difference. The chemical potential μ is obviously changing the quantization of a finite chain. The chosen parameters are t = 4 meV, Δ = 1.5 meV and μ = 3 meV and lead to the numerical energies $\left[0.43,\;4.034,\;6.068,\;9.603\right]$ (in meV). The value of k3,2,1d is approximately $\left[0.6360,\;1.2753,\;1.6086\right]$.

Standard image High-resolution image

The previous relations open another route to finding the condition leading to modes with exact zero energy. A convenient form of equation (92) is

Equation (95)

and the dispersion relation can be transformed into

Equation (96)

Both combinations κ1 ± κ2 yield the same energy, due to equation (95). If one of the brackets in equation (96) vanishes for κ1 + κ2 (κ1κ2), the second one does so for κ1κ2 (κ1 + κ2) too. Hence, zero energy is achieved exactly if

Equation (97)

This puts restrictions on Δ, t, μ. Together with the quantization rule in equation (93), this ultimately leads to (87) and to the condition $2\sqrt{{t}^{2}-{{\Delta}}^{2}}{< }\vert \mu \vert {< }2\vert t\vert $, defining the region where exact zero modes can form.

Regarding the remaining part of the topological phase diagram, we find that in the limit N the difference between even and odd N vanishes and the part of the μ = 0 axis between t = −Δ and t = Δ for even N leads to zero energy states too, in virtue of equation (42). Analogously, the area around the origin in figure 8, defined by $2\sqrt{{t}^{2}-{{\Delta}}^{2}}{< }\vert \mu \vert $ and μ < 2|t| with μ ≠ 0 does not support zero energy modes for all finite N. Instead, this area contains solutions with exponentially small energies, see equation (96), which become zero exclusively in the limit N. The wavevectors obey κ1 = iq1, κ2 = π + iq2 with real q1,2 for Δ2 > t2, and κ1 = iq1, κ2 = iq2 otherwise, which follows from equation (93) after some manipulations; see also [26]. Thus, the entire non trivial phase hosts zero energy solutions for N.

5.2. Eigenstates

The calculation of an arbitrary wave function of the Kitaev Hamiltonian without any restriction on t, Δ, μ is performed at best in the chiral basis yielding the block off-diagonal structure in equation (82). A suitable starting point is to consider a vector $\to {w}$ in the following form

with $\to {v}=\left({\xi }_{1},\dots ,\;{\xi }_{N}\right)$, $\to {u}=\left({\sigma }_{1},\dots ,\;{\sigma }_{N}\right)$. Solving for an eigenstate with eigenvalue λ demands on $\to {v},\;\to {u}$

Equation (98)

Equation (99)

with h from equation (83). Thus, $h\;{h}^{{\dagger}}\to {v}={\lambda }^{2}\;\to {v}$, and we recover equation (88) and the entries of $\to {v}$ obey equation (89) again. In appendix E we derived the closed formula for ξj, namely

Equation (100)

where ξ−2, ..., ξ1 are the initial values of the polynomial sequence dependent on the boundary conditions, and Xi(j) inherit the selective property

Equation (101)

That these functions Xi(j) exist and that they indeed satisfy equation (100) for arbitrary values of j is discussed in appendix E.

The remaining task is to obtain the initial values ξ−2, ..., ξ1, which follow from the open boundary conditions equation (90). Further, one has one free degree of freedom, which we to choose to be the entry ξ1 of $\to {v}$. In total our initial values are ξ1, ξ0 = 0; ξ−1 = 1/a and ξ−2 follows from ξN+1 = 0 and equation (100),

Demanding further N+2N = 0 quantizes the momenta κ1,2 and in turn λ = E±(κ1,2), according to equation (93) (the relation between Xj and κ1,2 is discussed in the appendix E). Notice that the form given by equations (100) and (102) (below) hold for all eigenstates of the Kitaev BdG Hamiltonian; the distinction between extended/decaying states and MZM is made by the values of k1,2, or equivalently κ1,2.

The second part of the eigenstate $\to {w}$, i.e. $\to {u}$, follows in principle from equation (99). A simpler and faster way is to consider ${h}^{{\dagger}}\;h\;\to {u}={\lambda }^{2}\;\to {u}$. A comparison of h and h reveals that they transform into each other by exchanging a and b and switching μ into −μ. Consequently, the structure of the entries of $\to {u}$ follow essentially from the ones of $\to {v}$. Thus, we find that the σi obey equation (89) too, yielding

Equation (102)

with the same functions Xi(j). The boundary condition on $\to {u}$ reads

Proceeding as we did for $\to {v}$ yields σ0 = 0, σ−1 = 1/b and

Equation (103)

where σ1 is fixed by the first line of equation (99)

Equation (104)

The last open boundary condition N+2N = 0 is satisfied for λ = E±(κ1,2) and thus already by the construction of $\to {v}$. We notice that we assumed λ ≠ 0 in order to obtain σ1; the case λ = 0 is discussed in section 4 below.

In the limit of μ = 0 on $\to {w}$ it holds that

for all values of l; thus only two initial values for ξj and two for σj are necessary to fix the sequences.

Finally, one can prove easily that the functions Xi(j) are always real and in consequence all ξj (σj) are real (pure imaginary) if ξ1 is chosen to be real. Thus, the corresponding operators ${\psi }_{\to {w}}\;\left({\psi }_{\to {w}}^{{\dagger}}\right)$ can never square to 1/2.

6. MZM eigenvectors at finite μ

The technique outlined above (in particular, equation (104)) cannot be used directly for exact zero energy modes, because then λ = 0 in equations (98) and (99). In this section we demonstrate the Majorana nature of the zero energy solutions satisfying equation (87), and we give the explicit form of the associated MZM using a different technique, which (similar to the Chebyshev polynomials method in [27]) requires only the use of Fibonacci, not Tetranacci polynomials. This simplification is caused by the fact that setting λ = 0 decouples the two Majorana sublattices, while setting μ = 0 decouples the two SSH-like chains.

We use the SSH-like description equation (23) of the Kitaev chain where μ ≠ 0 couples both chains together. Consequently an eigenstate $\to {\psi }={\left({\to {v}}_{\alpha },\;{\to {v}}_{\beta }\right)}^{\text{T}}$ has in general no zero entries and we use the same notation for the components of ${\to {v}}_{\alpha }$, ${\to {v}}_{\beta }$ as in the sections 4.1 and 4.2.

The zero energy values are twice degenerated, as one can see from equation (86), and the associated zero modes are connected by the chiral symmetry $\mathcal{C}$. Thus, we get zero energy states by superposition ${\to {\psi }}_{A,B}\;{:=}\left(\to {\psi }{\pm}\mathcal{C}\to {\psi }\right)/2$ too. The chiral symmetry equation (80), contains an alternating pattern of ±1, such that ${\to {\psi }}_{A}\;\left({\to {\psi }}_{B}\right)$ includes only non zero entries on the Majorana sublattice A (B). Hence, ${\to {\psi }}_{A}$ (${\to {\psi }}_{B}$) contains only xl (${\mathcal{X}}_{l}$) and ${\mathcal{Y}}_{j}$ (yj) terms and the last component depends on whether N is odd or even. In the latter case we have

The form of the odd N eigenvectors is quite similar, see equations (G15) and (G16).

The composition of ${\to {\psi }}_{A}$ is illustrated in figure 10, where its entries are shown to form a sawtooth like pattern, following the entries of ${\to {\psi }}_{A}$ on both SSH-like chains.

Figure 10.

Figure 10. Illustration of the sawtooth pattern of ${\to {\psi }}_{A}$. The real space position of the entries xl (${\mathcal{Y}}_{l}$) of ${\to {\psi }}_{A}$ is at j = 2l − 1 (j = 2l) on chain α (β) and marked by the blue (light blue) spheres. The blue line connects these entries as guide to the eye.

Standard image High-resolution image

The full calculation is given in appendix G. We focus here on ${\to {\psi }}_{A}$ exclusively, because the ${\to {\psi }}_{B}$ components follow essentially from ${\to {\psi }}_{A}$ by exchanging a and b and replacing by −. The chemical potential has still to obey equation (87).

The components of the zero mode ${\to {\psi }}_{A}$ have to satisfy (l = 1, ..., N/2 − 1)

Equation (105)

Equation (106)

for even N, and the open boundary conditions are

The situation for the entries of ${\to {\psi }}_{A}$ for odd N is similar

Equation (107)

Equation (108)

where j = 1, ..., (N − 1)/2, i = 1, ..., (N − 3)/2. The open boundary condition changes to

Solving these recursive formulas leads in both cases to

Equation (109)

with θn = /(N + 1), n = 1, ..., N and

Equation (110)

where x1 is a free parameter and − a/b ⩾ 0 due to t2 ⩾ Δ2. Recalling that $a=i\left({\Delta}-t\right)$, $b=i\left({\Delta}+t\right)$, equations (109), (110) predict an oscillatory exponential decay of the coefficients xl, ${\mathcal{Y}}_{l}$. For example

Equation (111)

where the decay length is defined by $\xi =d/\left\vert \mathrm{ln}\left(\frac{t-{\Delta}}{t+{\Delta}}\right)\right\vert $, for t > Δ > 0. Summarizing: the zero energy modes ${\to {\psi }}_{A,B}$ look like small or strongly suppressed standing waves with n − 1 nodes for n = 1, ..., Nmax and Nmax = N/2 (Nmax = N − 1/2) for even (odd) N. The expressions for ${\mathcal{X}}_{l}$ and yl are obtained in a similar way

Equation (112)

Equation (113)

and ${\mathcal{X}}_{1}$ can be freely chosen. The open boundary conditions for l = 0 are satisfied by construction of ${\mathcal{Y}}_{l}$ (yl), while the remaining ones follow due to the structure of θn.

The zero mode ${\to {\psi }}_{A}$ is shown in figure 11 for a various range of parameters. For not too large ratios t/Δ > 1, the zero mode ${\to {\psi }}_{A}$ is mostly localised at one end of the Kitaev chain and decays away from it in an oscillatory way. The eigenstate ${\to {\psi }}_{B}$ is concentrated on the opposite end. While the oscillation depends on the chemical potential μn associated to the zero mode, according to equation (87), the decay length is only set by the parameters Δ and t. Thus, as the ratio of t/Δ is increased, the zero energy mode gets more and more delocalized.

Figure 11.

Figure 11. Majorana zero mode ${\to {\psi }}_{A}$ for various parameter sets. The considered parameters are denoted by different symbols on the topological phase diagram. The dark (light) blue spheres follow xl/x1 (${\mathcal{Y}}_{l}/{x}_{1}$) at position j = 2l − 1 (j = 2l) from equation (109) and (110). The decay length of the MZM increases for larger ratios of t/Δ for a fixed value of θn, until the state is delocalised over the entire system. Lowering the chemical potential, e.g. following the vertical orange line, but keeping t/Δ fixed, changes the shape of the MZM's. Large decay length and chemical potential leads to Majorana modes which have highest weight in the center of the chain.

Standard image High-resolution image

The zero energy states ${\to {\psi }}_{A,B}$ are MZM's, since they are eigenstates of the particle hole operator $\mathcal{P}$ equation (79) for real or pure imaginary values of x1, ${\mathcal{X}}_{1}$. Further, the states $\to {\psi }={\to {\psi }}_{A}+{\to {\psi }}_{B}$ and $\mathcal{C}\to {\psi }={\to {\psi }}_{A}-{\to {\psi }}_{B}$ are MZM's too. On the other hand a fermionic state is constructed with ${\to {\psi }}_{{\pm}}={\to {\psi }}_{A}{\pm}i\;{\to {\psi }}_{B}$, similar to what was found in section 4 for the μ = 0 case, or at the Kitaev points in equations (4a) and (4b) in section 2.1.

There are three limiting situations we would like to discuss: t → ±, N, and how the eigenstate changes if the sign of the chemical potential is reverted. For the first situation we notice that larger hopping amplitudes affect the decay length ξ. Because −a/b → 1 for t → ±, this implies also that ξ in that limit. Hence oscillations are less suppressed for large values of t, as illustrated in figure 11. Already a ratio of t/Δ ≈ 100 is enough to avoid a visible decay for N ≈ 20. This effect can be found as long as N is finite, for appropriate values of the ratio t/Δ.

What happens instead for larger system sizes? Regardless of how close −a/b is to 1, for a finite t, at some point the exponent j (l) in ${x}_{l}\;{\mathcal{Y}}_{l}$, ${\mathcal{X}}_{l}$ and yl leads to significantly large or small values. Thus, the state ${\to {\psi }}_{A}$ (${\to {\psi }}_{B}$) becomes more localised on the left (right) end for t > 0, and on the right (left) one for t < 0.

If we change the chemical potential to its negative value we find that yl, ${\mathcal{Y}}_{l}$ only change their sign. For odd N and for θn = π/2, i.e. μ = 0, one recovers the result in the equations (71) and (72).

7. Numerical results and impact of disorder

In this section we discuss the impact of disorder on the topological boundary states. To this extent we investigate numerically the lowest energy eigenvalues of the finite Kitaev chain.

7.1. The clean Kitaev chain

The features predicted analytically above are also clearly visible in the numerical calculations. The lowest positive energy eigenvalues E0 of a finite Kitaev chain, with the Hamiltonian given by equation (1) and varying parameters, are shown in figure 12. The phase diagram in figure 12(a) is the numerical equivalent of that shown in figure 11, but for a smaller range of t and μ. Because of the necessarily discrete sampling of the parameter space, the zero energy lines are never met exactly, hence along the Majorana lines we see only a suppression of E0. Along the border of the topological regime, μ ≲ 2t, all the boundary states in a finite system have finite energy, as shown in figure 12(b). Figure 12(c) displays E0 for fixed t/Δ = 17, as a function of μ and N. The number of near-zero energy solutions increases linearly with N, according to equation (87). It is worth noting that a spatial overlap between Majorana components of the end states in a short system does not need to lead to finite energy (cf. the right column of figure 11). The decay length ξ of the in-gap eigenstates, defined in equation (73), is determined by the ratio t/Δ and is the same both for the near-zero energy states along the Majorana lines and for the finite energy states between them. It is the maximum energy of the boundary states that decreases as E0,max ∝ exp(−Nd/ξ) as N is increased, as illustrated in figure 12(d), in agreement with equation (42) and [2, 23, 24]. The minimum energy of zero can be reached for any chain length, provided that the chemical potential is appropriately tuned.

Figure 12.

Figure 12. Numerical results for the energy E0 of the lowest positive energy state as a function of t/Δ, μ/Δ and system size N. (a) E0 as a function of t and μ for N = 20. The red line marks the boundary of the bulk topological phase. The N/2 dark lines coincide with the Majorana lines given by equation (87). (b) Zoom into the neighbourhood of the Kitaev point at μ = 0 and t = Δ, showing the absence of zero energy solutions for Δ < t in the nominally non-trivial phase. (c) E0 as a function of μ and the system size N for t/Δ = 17. The red line marks μ = 2t, the boundary of the bulk topological phase. As N increases, the number of μ yielding zero energy solutions also increases according to equation (87), and the maximum energies of the bound states decrease. (d) The values of E0 for the same set of parameters as in (c), projected onto the NE0 plane. The ground state energy E0 shows a μ dependent, oscillatory behavior in the system size N, where the maximum energies follow with very good accuracy an aμ exp(−Nd/ξ) rule for sufficiently large values of N, with ξ from equation (73) and the numerical prefactor aμ.

Standard image High-resolution image

7.2. Topological protection against Anderson disorder

One of the most sought after properties of topological states is their stability under perturbations which do not change the symmetry of the Hamiltonian. In order to see whether the non-Majorana topological modes enjoy greater or lesser topological protection than the true Majorana zero modes, we have calculated numerically the spectrum of a Kitaev chain with Anderson-type disorder as a function of μ for N = 20 and three different values of t/Δ. The disorder was modeled as an on-site energy term ɛi whose value was taken randomly from the interval [−W, W]. The energies E0, E1 of the two lowest lying states are plotted in figure 13. In all plots W = 4Δ, i.e. twice larger than the ±2Δ gap at μ = 0. Each curve is an average over 100 disorder configurations. Even with this high value of the disorder it is clear that the energy of the in-gap states is rather robust under this perturbation. For t/Δ = 2, i.e. close to the Kitaev point where the boundary states are most localized, they are nearly immune to disorder–its influence is visible only at high μ and in the energy of the first extended state. For higher ratios of t/Δ, closer to the value of N + 1 (cf equation (40) and the discussion under equation (42)), the lowest energy states seem to be strongly perturbed and the Majorana zero modes entirely lost. This is however an artifact of the averaging–the energy E0 plotted in figure 14 for several disorder strengths W shows that for any particular realization of the disorder the zero modes are always present, but their positions shift to different values of μ. The existence of these crossings is in fact protected against local perturbations and they correspond to switches of the fermionic parity [22].

Figure 13.

Figure 13. The lowest (E0) and second lowest (E1) energy eigenvalue of a Kitaev chain with N = 20 sites as a function of μ for various ratios of t/Δ. Thin lines correspond to the eigenvalues of a clean chain, thick lines to those of a chain with random on-site disorder ɛi ∈ [−4Δ, 4Δ]. The red line marks the boundary of the topological phase. Each disorder line results from averaging over a 100 disorder realizations.

Standard image High-resolution image
Figure 14.

Figure 14. The lowest energy eigenvalue E0 of a Kitaev chain with N = 20 sites as a function of μ for t/Δ = 17. Thin lines correspond to the eigenvalues of a clean chain, thick lines to those of a chain with random on-site disorder ɛi ∈ [−W, W], for two disorder realizations. The red line marks the boundary of the topological phase.

Standard image High-resolution image

8. Conclusion

Due to its apparent simplicity, the Kitaev chain is often used as the archetypal example for topological superconductivity in one dimension. Indeed, its bulk spectrum and the associated topological phase diagram are straightforward to calculate, and the presence of Majorana zero modes (MZM) at special points of the topological phase diagram, known as Kitaev points (μ = 0, t/Δ = ±1 in the notation of this paper), is easy to demonstrate. However, matters become soon complicated when generic values of the three parameters μ, Δ and t are considered.

In this work we have provided exact analytical results for the eigenvalues and eigenvectors of a finite Kitaev chain valid for any system size. Such knowledge has enabled us to gain novel insight into the properties of these eigenstates, e.g. their precise composition in terms of Majorana operators and their spatial profile.

Our analysis confirms the prediction of Kao [21], whereby for finite chemical potential (μ ≠ 0) zero energy states only exists for discrete sets of μ(Δ, t) which we dubbed 'Majorana lines'. We calculated the associated eigenvectors and demonstrated that such states are indeed MZM. Importantly, such MZM come in pairs, and because they are made up of Majorana operators of different types, they are orthogonal. In other words the energy of these modes is exactly zero, even when the two MZM are delocalized along the whole chain (which depends on the state's decay length ξ).

Beside of the Majorana lines, but still inside the topological region, finite energy boundary states exist. We studied the behavior of the energy E0 of the lowest state numerically as a function of t/Δ, μ/Δ and of the system size L = Nd. We found that with good accuracy the maximum energy E0,max ∝ exp(−L/ξ). This energy, and hence the energy of all the boundary states, tends to zero in the thermodynamic limit N. For fixed N the ratio t/Δ can be varied until the decay length ξ becomes of the order of the system size L and hence the associated E0,max is not exponentially close to zero energy.

All the boundary states in the topological region, whether of exact zero energy or not, are of topological nature, as predicted by the bulk-edge correspondence. This fact is important in the context of topological quantum computation. In fact, whether a state has exact zero energy or not is not relevant for computation purposes, as long as this state is topologically protected. For the prototypical model of the finite Kitaev chain considered in this paper this condition is best satisfied in the vicinity of the Kitaev points t/Δ = ±1, μ = 0 independent of the system's size. When a random disorder of Anderson type is introduced in the Kitaev chain, our numerical results show that the upper bound E0,max is remarkably robust, confirming the topological protection of all topological states.

The fact that even in short chains, with large spatial overlap of the Majorana components, the energy of the subgap state can still be strictly zero could have interesting implications for the experiments. For example, in STM-spectroscopy experiments with topological chains like in Nadj-Perge et al in [12] or Kim et al in [16], this would yield a zero-bias Majorana peak delocalized along the whole chain.

Although our treatment using Tetranacci polynomials for μ ≠ 0 is general (sections 5 and 6), we have dedicated special attention to two parameter choices in which the Tetranacci polynomials reduce to the generalized Fibonacci polynomials. The first case is that of zero chemical potential discussed in sections 3 and 4 of the paper, where the Kitaev chain turns out to be composed of two independent SSH-like chains. This knowledge allows one a better understanding of the difference between an even and an odd number N of sites of the chain. This ranges from different quantization conditions for the allowed momenta of the bulk states, to the presence of MZM. While MZM are always present for odd chains, they only occur at the Kitaev points for even chains. When μ is allowed to be finite, the Kitaev points develop into Majorana lines hosting MZM for both even and odd chains. In the thermodynamic limit the distinction between even and odd number of sites disappears. The fact that E = 0 at the Majorana lines decouples the two Majorana sublattices, and again allows us to use the simpler Fibonacci polynomials.

Acknowledgments

The authors thank the Elite Netzwerk Bayern for financial support via the IGK 'Topological Insulators' and the Deutsche Forschungsgemeinschaft via SFB 1277 Project B04. We acknowledge useful discussions with A Donarini, C de Morais Smith and M Wimmer.

Appendix A.: A note on Fibonacci and Tetranacci polynomials

An object of mathematical studies are the Fibonacci numbers ${F}_{n}\;\left(n\in {\mathbb{N}}_{0}\right)$ defined by

Equation (A1)

which frequently appear in nature. A more advanced sequence is the one of the Fibonacci polynomials [36] Fn(x), where

Equation (A2)

with an arbitrary complex number x which gives different weight to both terms. The polynomial character becomes obvious after a look at the first few terms

Equation (A3)

The so called generalized Fibonacci polynomials [37] are defined by

Equation (A4)

where x, y are two complex numbers. The second weight changes the first elements of the sequence in equation (A2) to

Equation (A5)

There is a general mapping between the sequences Fn(x) in equation (A2) and Fn(x, y) in equation (A4), namely

Equation (A6)

where ${F}_{n}\left(x/\sqrt{y}\right)$ obeys equation (A2) with $x/\sqrt{y}$ instead of x.

A last generalization is to consider arbitrary initial values Fi(x, y) = fi, i = 0, 1 and keeping [38]

Equation (A7)

This changes the first terms into

The Fibonacci polynomials ζ2n−1, ζ2n, epsilon2n−1, epsilon2n we consider in the spectral analysis are of the last kind with x = λ2 + a2 + b2, y = −a2 b2 (a = i(Δ − t), b = i(Δ + t)) and with different initial values for odd and even index as well as different ones for ζ and epsilon. The first terms for ζ2n−1 are

while one finds for ζ2n

The expressions for epsilon2n (epsilon2n−1) follow from the ones of ζ2n (ζ2n−1) by exchanging a and b.

A closed form for Fibonacci numbers/polynomials is called a Binet form, see for example [38]. In the case of ζn this form is given in equations (B20) and (B21).

In order to obtain the general quantization condition for the wavevectors of the Kitaev chain we face further generalizations of Fibonacci polynomials, so called Tetranacci polynomials τn, defined by

Equation (A8)

with four complex variables x0, ..., x3 and four starting values τ0, ..., τ3. These polynomials are a generalisation of Tetranacci numbers [41, 42] and their name originates from the four terms on the r.h.s. of equation (A8). The form of Tetranacci polynomials ξj we deal with in this work, is provided by equation (D4).

Appendix B.: Spectrum for μ = 0

B.1. Characteristic polynomial in closed form

The full analytic calculation of the spectrum is at best performed in the basis of Majorana operators ${\gamma }_{j}^{A\left(B\right)}$, ordered according to the chain index

Equation (B1)

Then the BdG Hamiltonian becomes block tridiagonal

Equation (B2)

where A and B are 2 × 2 matrices

Equation (B3)

Since we are interested in the spectrum, we have essentially only to calculate (and factorise) the characteristic polynomial ${P}_{\lambda }\left({\mathcal{H}}^{\text{KC}}\right)=\mathrm{det}\left(\lambda \mathbb{1}-{\mathcal{H}}_{\text{KC}}\right)$ which reads simply

Equation (B4)

at zero μ. In the following we will consider λ to be just a 'parameter', which is not necessarily real in the beginning. Further, we shall impose (only in the beginning) the restriction λ ≠ 0. However, our results will even hold without them. The validity of our argument follows from the fact that the determinant Pλ is a smooth function

Equation (B5)

in the entire parameter space $\mathbb{P}{:=}{\mathbb{C}}^{3}$, which contains a, b and λ.

The technique we want to use to evaluate Pλ is essentially given by the recursion formula of the 2 × 2 matrices Λj [43, 44]:

Equation (B6)

where j = 1, ..., N and ${P}_{\lambda }=\prod _{j=1}^{N}\mathrm{det}\left({{\Lambda}}_{j}\right)$.

The matrices B and B are pure off-diagonal matrices and since $\lambda {\mathbb{1}}_{2}$ is diagonal, one can prove that Λj has the general diagonal form of ${{\Lambda}}_{j}{:=}\left[\begin{bmatrix}{cc}\hfill {x}_{j}\hfill & \hfill 0\hfill \\ \hfill 0\hfill & \hfill {y}_{j}\hfill \end{bmatrix}\right]$ (for all j). The application of equation (B6) leads to a recursion formula for both sequences of entries

and the initial values are x1 = y1 = λ. We find xj and yj to be fractions in general, and define ζj, epsilonj, βj and δj by

to take this into account. The initial values can be set as

Equation (B7)

Equation (B8)

and after a little bit of algebra we find their growing rules to be

Equation (B9)

Equation (B10)

Equation (B11)

Equation (B12)

where j starts from 1. The definitions ζ0 := δ1 = 1 and epsilon0 := β1 = 1, enable us to get rid of the δj and βj terms inside equations (B9) and (B10). Hence

Equation (B13)

Equation (B14)

which leads to the relations

Equation (B15)

Equation (B16)

We already extended the sequences of ζj and epsilonj artificially backwards and we continue to do so, using the equations (B13) and (B14), starting from j = −1 with ζ−1 = epsilon−1 = 0. Please note there are no corresponding x0, y0 or even x−1, y−1 expressions, since they would involve division by 0.

The last duty of βj and δj is to simplify the determinant Pλ by using the equations (B8), (B11) and (B12)

Equation (B17)

which reduces the problem to finding only ζN and epsilonN.

Please note that the determinant is in fact independent of the choice of the initial values for ζ1, epsilon1, β1 and δ1 in the equations (B7) and (B8). Further, equations (B13), (B14) and (B17) together show the predicted smoothness of Pλ in $\mathbb{P}$ and all earlier restrictions are not important anymore. Finally we consider λ to be real again.

Even though it seems that we are left with the calculation of two polynomials, we need in fact only one, because both are linked via the exchange of a and b. Note that λ is considered here as a number and thus does not depend on a and b. Further, the dispersion relation is invariant under this exchange.

The connection of ζj and epsilonj for all j ⩾ −1 is

and can be proven via induction using equations (B13) and (B14). Decoupling ζj and epsilonj yields

Equation (B18)

where one identifies them as (generalized) Fibonacci polynomials [36, 37]. The qualitative difference between even and odd number of sites is a consequence of equation (B18) and the initial values for ζj.

The next step is to obtain the closed form expression of ζj (epsilonj), the so called Binet form. We focus exclusively on ζj.

One way to keep the notation easier is to introduce x := λ2 + a2 + b2, y := a2b2, vj := ζ2j and uj := ζ2j−1, such that uj (vj) obey

The Binet form can be obtained by using a power law ansatz ujrj, leading to two fundamental solutions

Equation (B19)

Please note that this square root is always well defined, which can be seen in the simplest way by setting λ to zero. Consequently, the difference between r1 and r2 is never zero.

A general solution of uj (vj) can be achieved with a superposition of r1,2 with some coefficients c1,2,

due to the linearity of their recursion formula. Both constants c1 and c2 are fixed by the initial values for ζj, for example u0 = ζ−1 = 0 and u1 = ζ1 = λ and similar for vj. After some simplifications, we finally arrive at

Equation (B20)

Equation (B21)

in agreement with [3638]. The validity of the solutions is guaranteed by a proof via induction, where one needs mostly the properties of r1,2 to be the fundamental solutions. The exchange of a and b leads to the expressions

Equation (B22)

Equation (B23)

where we used that r1,2 is symmetric in a and b. At this stage we have the characteristic polynomial in closed form for all Δ, t and more importantly for all sizes N at zero μ.

We can already anticipate the twice degenerated eigenvalues of the odd sized Kitaev chain, because from the closed forms of epsilonj and ζj it follows immediately

Equation (B24)

Notice that equation (B24) is important to derive the characteristic polynomial via the SSH description of the Kitaev BdG Hamiltonian at μ = 0 and to show the equivalence to the approach used here. It is recommended to use the determinant formula in [45] together with equations (B9) and (B10) for the proof.

The main steps of the factorisation are mentioned in the next section.

B.2. Factorisation of generalized Fibonacci polynomials

The trick to factorise our Fibonacci polynomials [36, 37] bases on the special form of r1,2. The ansatz is to look for the eigenvalues λ in the following form

Equation (B25)

which is actually the definition of θ. The hermiticity of the Hamiltonian enforces real eigenvalues and consequently θ can be chosen either real, describing extended solutions, or pure imaginary, which is connected to decaying states. The ansatz leads to an exponential form of the fundamental solutions

and we consider $\theta \in \mathbb{R}$ first. Thus, we find the eigenvalues for odd N

One obvious solution is λ = 0. The introduction of 2kd = θ, where d is the lattice constant of the Kitaev chain, leads to:

Equation (B26)

and solutions inside the first Brillouin-zone are given by

where n runs from 1, ..., N without (N + 1)/2. Please note that equation (B26) cannot be satisfied for N = 1.

The even N case requires more manipulations. We first rearrange equation (B21) as

The expressions λ2 + b2r1,2 are simplified to

In the end ζ2j becomes

Equation (B27)

Note that the competition of Δ and t is hidden inside the square root

affecting both the quantization condition and the dispersion relation E±(k) = λ(θ), which follow from equation (B25). However, both situations lead to the same result, because the momenta and the spectrum are shifted by π/2 (with respect to kd). From ζN it follows:

or in shorter form

Equation (B28)

for even N. The polynomial epsilonN can be treated in the same way leading to

Equation (B29)

From equation (B25) follows the bulk spectrum for all N,

in agreement with equations (34) and (35).

The case of decaying states is similar, but not just done by replacing k by iq. The following case is only valid for even N, since we have already all 2N eigenvalues of the odd N case.

Our ansatz is modified to

by an additional minus sign, which is important to find the decaying state solutions. After some manipulations ζN = 0 yields the quantization conditions

Equation (B30)

Equation (B31)

where qd = θ/2. The conditions for qd = 0 as solution, corresponding to infinite decay length ξ, turn out to be ±t/Δ = N + 1 (if |t| ⩾ |Δ|) or ±Δ/t = N + 1 (else) and follow by applying the limit qd → 0 on equations (B30) and (B31).

A last simplification can be done for qd ≠ 0

where we introduced

The criterion to find a wave vector is that (−m) ⩾ 1, but not larger than N + 1, which leads then to exactly two solutions ±q and otherwise to none. The corresponding eigenvalues can be obtained from

which can be zero. The results for epsilonN can be obtained by replacing t with −t everywhere.

Appendix C.: Eigenvectors for zero μ

The simplest way to calculate the eigenstates of the Kitaev Hamiltonian is the use of the SSH-like basis for μ = 0 from equation (23). We define the eigenvector $\to {\psi }$ as

for all N to respect the structure of the Hamiltonian. Moreover, one can search for solutions belonging only to one block ${\left({\to {v}}_{\alpha },\;{\to {0}}_{\beta }\right)}^{\text{T}}$ or ${\left({\to {0}}_{\alpha },\;{\to {v}}_{\beta }\right)}^{\text{T}}$, without any restriction. In other words either is ${\to {v}}_{\alpha }$ zero or ${\to {v}}_{\beta }$ and we will mention only non zero entries from now on. We report here only about the calculation of ${\to {v}}_{\alpha }$, because the one for ${\to {v}}_{\beta }$ can be performed analogously.

The general idea behind the eigenvector calculation of tridiagonal matrices is given in [40], but we consider here all possible configurations of parameters.

C.1. N even

The sublattice vectors are defined via the N entries

Solving $\left({\mathcal{H}}_{\alpha }^{\text{even}}-\lambda \;\mathbb{1}\right)\;{\to {v}}_{\alpha }=0$ leads to

Equation (C1)

Equation (C2)

and

Equation (C3)

Equation (C4)

where l runs from 1 to (N/2) − 1. The coupled equations (C1)–(C4) for the entries of the eigenvector are continuous in all parameters. However, resolving to the xl's and yl's may lead to problems for certain values of Δ, t and λ.

Case 1. |Δ| ≠ |t|. The parameter setting excludes λ = 0, as we found from our spectral analysis in section 3. Equations (C3) and (C4) are used to disentangle x's and y's. Both sequences obey

Equation (C5)

where l = 1, ..., N/2. Thus the y's and x's are Fibonacci polynomials [36, 37]. The difference to the previous ones found for the spectrum is that the new version can be dimensionless in physical units, depending on the initial values. The transformation formula [37] to pass from the unitless recursion formula to the other one is given by equation (A6).

The Binet form of the dimensionless sequences is obtained with same treatment as for the spectrum. The power ansatz ylfl yields the fundamental solutions f1,2, obeying

Equation (C6)

Equation (C7)

Equation (C8)

Due to the linearity of the recursion formula, the most generic ansatz for yl is

where the constants c1, 2 follow from the initial values of y1, 2. The calculation of both constants leads to

where Tl is simply [40]

Analogously we find

A short comment on the initial values y1,2. A hermitian matrix is always diagonalisable, regardless of degeneracies in its spectrum and an eigenvector is well defined only up to the prefactor. Consequently we have the freedom to choose one component of ${\to {v}}_{\alpha }$. This choice will in turn define all remaining initial values.

Consider for example x1 to be a fixed value of our choice. We find y1, x2 and y2 to be

The y2 can be rewritten as ${y}_{2}={y}_{1}\left[{f}_{1}+{f}_{2}\right]$ which leads to a simpler form of all yl's [40]

Equation (C9)

After a bit of algebra, one finds xl to be

Equation (C10)

So far we found the general solutions of the recursion formulas equations (C1)–(C4). The comparison of equations (C2) and (C3) leads to

Equation (C11)

because the recursion formulas themselves do not care about any index limitation. The last equation means only that the wave function of a finite system has to vanish outside, at the boundary, yielding the quantization rule.

The extended states can be obtained with

where equations (C6) and (C7) relate kd and λ, and Tl is recast as

Equation (C12)

The last equation for Tl yields via equations (C10) and (C11) the quantization condition. Thus, the momenta k obey

where each solution defines two states with the energy E± from equation (35).

The decaying states depend strongly on the interplay of Δ and t. The ansatz is

where s is defined as

Finally the coefficient Tl becomes

The proper q, if existent, leads to two states and satisfies

where m is

Equation (C13)

In total we have already all N non normalized states with respect to the chain α and this approach holds as long as |Δ| ≠ |t|.

The remaining cases start again from the equations (C1)–(C4).

Case 2. Eigenvectors at the Kitaev point. We consider now Δ = −t, or b = 0, and we have to solve

where l runs from 1 to (N − 2)/2. A zero energy mode is obviously not existing on the α subchain, because λ = 0 would lead to ${\to {v}}_{\alpha }=0$ which is not an eigenvector by definition. These zero modes belong to the subchain β for Δ = −t. The only possible eigenvalues for the extended modes of the α chain are λ = ±2t [1, 2], see equation (35). Recalling a = −2it, leads to N/2 independent solutions of dimerised pairs (xl, yl) with yl = ∓ixl and the signs are with respect to the eigenvalues.

The last cases belong to Δ = t (a = 0), where we search for the solution of

where l runs from 1 to (N − 2)/2. The first (second) line clearly states that either λ is zero and/or x1 (yN/2). The zero λ means on the one hand that most entries vanish x2 = x3 = ⋯ = xN/2 = 0 and y1 = y2 = ⋯ = y(N−2)/2 = 0, since b = 2it ≠ 0 to avoid a trivial Hamiltonian. On the other hand we have two independent solutions, first

and second

describing the isolated MZM's at opposite ends of the chain. In the case of λ = ±2t, we have N − 2 independent solutions in form of pairs (yl, xl+1) with yl = ±ixl+1.

The non trivial solutions for ${\to {v}}_{\beta }$ follow by replacing ${x}_{l}\to {\mathcal{X}}_{l}$, ${y}_{l}\to {\mathcal{Y}}_{l}$ and t → −t everywhere.

C.2. N odd

The eigenvectors have similar shape

but the last entry is different compared to the even N case. Although both subchains have the same spectrum, it is possible to consider a superposition of eigenstates of the full Hamiltonian which belongs to only one chain, for example α. We consider ${\to {v}}_{\beta }$ to be zero.

The eigenvector system for ${\to {v}}_{\alpha }$ reads

and

with $l=1,\;\;\dots ,\;\;\frac{N-3}{2}$ and $i=1,\;\;\dots ,\;\;\frac{N-1}{2}$.

If we consider a, b and λ all to be different from zero, we find again that the entries of ${\to {v}}_{\alpha }$ are Fibonacci polynomials obeying the same recursion formula as in the even N case and lead to the same solution

where $l=1,\;\;\dots ,\;\frac{N+1}{2}$ and Tl, (i) is as before. The ansatz f1 = e2i kd, f2 = e−2i kd for the extended states influences Tl (Ti analogously)

and leads via

to the equidistant quantization $k\;\equiv \;{k}_{n}=\frac{n\;\pi }{N+1}$ with n = 1, ..., (N − 1)/2, due to the number of eigenvectors of the single SSH-like chain. Both chains share the same spectrum for odd N and thus we have in total n = 1, ..., N, n ≠ (N + 1)/2.

We report here shortly on all other parameter situations.

  • (a)  
    If we consider a and b to be different from zero, but λ = 0, we find only one state
    Equation (C14)
    and l runs from 1 to (N − 1)/2.
  • (b)  
    If Δ = t, i.e. a = 0, but λ = ±2t ≠ 0, we find (N − 1)/2 solutions (yl, xl+1) with yl = ±ixl+1, l = 1, ..., (N − 1)/2 and x1 = 0 for all.The zero mode of this setting (Δ = t) is an MZM localized on x1 = 1, while all other components are zero.
  • (c)  
    If Δ = −t (b = 0) and λ ≠ 0 we find (N − 1)/2 solutions of the form (xl, yl) with yl = ±ixl l = 1, ..., (N − 1)/2 and ${x}_{\frac{N+1}{2}}=0$ for all of them. The MZM is localised at ${x}_{\frac{N+1}{2}}=1$ for b = 0.The results for the α chain follow again by replacing ${x}_{l}\to {\mathcal{X}}_{l}$, ${y}_{l}\to {\mathcal{Y}}_{l}$ and t → −t.

Appendix D.: Spectrum for finite μ

The BdG Hamiltonian, expressed in the chiral basis ${\hat{{\Psi}}}_{c}={\left({\gamma }_{1}^{A},\;{\gamma }_{2}^{A},\;\;\dots ,\;\;{\gamma }_{N}^{A},\;{\gamma }_{1}^{B},\;{\gamma }_{2}^{B},\;\;\dots ,\;\;{\gamma }_{N}^{B}\right)}^{\text{T}}$ leads via ${\hat{H}}_{\text{kc}}=\frac{1}{2}{\hat{{\Psi}}}_{c}^{{\dagger}}\;{\mathcal{H}}_{c}\;{\hat{{\Psi}}}_{c}$ to

Equation (D1)

where the matrix h is

Equation (D2)

As mentioned in section 5, we look for a solution of

Equation (D3)

with $\to {v}={\left({\xi }_{1},\;{\xi }_{2}\;\;\dots ,\;\;{\xi }_{N}\right)}^{\text{T}}$ to find the general quantization rule. The entries of the matrix hh are

and equation (D3) becomes the Tetranacci sequence

Equation (D4)

where j = 1, ..., N − 5. The missing four boundary terms are

We extend the Tetranacci sequence from j = − to j = , i.e. the index limitations in equation (D4) can be ignored, while $\to {v}$ still contains only ξ1, ..., ξN. Consequently, we can simplify the boundary conditions by using the recursion formula and further any restriction like N > 3 does not exist. We find

Equation (D5)

Equation (D6)

Equation (D7)

The procedure we followed in the context of Fibonacci polynomials was to obtain a closed form with the ansatz ξj = rj, r ≠ 0. So we do here on starting from equation (D4). Thus, the characteristic equation for r reads

and we have to find all four zeros to determine ξj in the end. We introduce two new variables

Equation (D8)

Equation (D9)

to simplify the expressions in the following. The characteristic equation becomes

Dividing by r2 (r ≠ 0) and defining S := r + r−1 leads to

Equation (D10)

where we can read out the solutions S1,2

Equation (D11)

The definition of S amounts to an equation for r

Thus one can insert the solutions S1,2 and solve for r. We find

Equation (D12)

yielding directly rjrj = 1. Here, we choose the ansatz

Equation (D13)

which is actually the definition of κ1,2. Since the coefficients S1,2 contain λ through the variable ζ, this is in the end an ansatz for λ. The expression for λ follows easily from equation (D10) by inserting equation (D13). Using the definiton of η and resolving for ζ first and in a second step for λ we finally arrive at Kitaev's bulk formula

Notice that by construction we have $\lambda =\lambda \left({\kappa }_{1}\right)=\lambda \left({\kappa }_{2}\right)$. Alternatively the sum of S1 and S2 leads via equations (D11), (D13) to

Equation (D14)

The use of equation (D14) on the dispersion relation will indeed yield

Equation (D15)

Let us return to ξj. Since the recursion formula in equation (D4) is linear, a superposition of all four solutions r±1, r±2

Equation (D16)

is still a solution with some coefficients ${c}_{1,2,3,4}\in \mathbb{C}$. From equation (D12) it follows

Equation (D17)

and thus

Equation (D18)

Further, equation (D15) implies that we consider a combination of states of the same energy. The usually following step would be to fix these constants, requiring four initial values. We can use e.g. ξ1 as free parameter. Further setting ξ0 = ξN+1 = 0, ξ−1 = (b/a)ξ1 as the boundary conditions yield a sufficient number of constraints.

The remaining condition N = N+2 yields the quantization rule then. However, if one is not interested in the state $\to {v}$ or in the general eigenstates of the Kitaev chain, but only in the quantization rule, one can use a much simpler approach. Using our ansatz for ξj from equation (D18) and being aware of the fact that the boundary conditions yield a homogeneous system, we find

where the boundary matrix B is

Demanding $\mathrm{det}\left(B\right)=0$ avoids a trivial solution and leads to the quantization rule in equations (93) and (94).

Appendix E.: The closed formula of Tetranacci polynomials

The goal here is to obtain the general solutions of a polynomial sequence ξj, $j\in \mathbb{Z}$, which obeys

Equation (E1)

with arbitrary initial values. We consider here ξ−2, ..., ξ1, other choices are possible too, to be the initial values. We want to determine a closed form expression for all ξj's. Similar to equation (D18), the general solution is given by a superposition of the four fundamental solutions r±i from equation (D12)

Equation (E2)

with some constants c1, ..., c4 which follow from ξ−2, ..., ξ1 via

Equation (E3)

Solving equation (E3) and factorising ξj into contributions of ξ−2, ..., ξ1 yields

Equation (E4)

The functions Xi(j) obey by construction (via equation (E3))

Equation (E5)

for all values of λ, μ, a, b and further obey equation (E1).

Despite the short form of ξj in equation (E4), the formulas of Xi(j) tend to be lengthy, such that we first introduce a short hand notation for their main pieces. We define

Equation (E6)

Equation (E7)

where the r.h.s of both equalities arise due to riri = 1 for i = 1, 2. With S1,2 from equation (D11) we find the Xi(j) to be

Equation (E8)

Equation (E9)

Equation (E10)

Equation (E11)

where $\overline{\sigma }$ is meant as 'not σ', e.g. if σ = 1 then we have $\overline{\sigma }=2$ and vice versa. The presence of S1,2 in the form of Xi(j) arises due to the definition of F1,2, since they are Fibonacci polynomials with inital values F1,2(0) = 0, F1,2(1) = 1 and obey

The proof is done by induction over j and using the relation between r±i and Si according to the equations (D11) and (D12).

The formulas for Xi(j) and Fi are exact and hold for all values of μ, a, b (t, Δ) and for all values of λ, regardless of whether equation (E1) describes an eigenvector problem or not. Notice that μ = 0 is a special situation, since

leads to

for μ = 0. Thus, we find

yielding

for all values of l.

The closed formula of ξj can be used in multiple ways. In the context of eigenvectors the exponential form of the fundamental solutions r±i according to equation (D17) is the direct connection to the momenta κ1,2, and their values follow from the quantisation rule in equation (93). The corresponding value of λ = E± (κ1,2) follows then from equation (7). The form of F1,2 transforms into a ratio of sin(κ1,2 j)/sin(κ1,2). However, once the energy E±(κ1,2) is known, the explicit use of κ1,2 is not important, since r±i follow also directly from equation (D12).

Appendix F.: The zeros of the determinant

Our first step is to calculate the determinant of the Kitaev chain in closed form. We use the chiral basis where the BdG Hamiltonian is given by equations (D1) and (D2). The determinant is obviously

Equation (F1)

and we need only the determinant of h. The calculation is performed with a sequence of polynomials [45] h0, ..., hN

Equation (F2)

with the initial values h0 = 1, h1 = − and the determinant of h is

Equation (F3)

We notice the Fibonacci character [3638] of the sequence in equation (F2) and continue with the calculation of the Binet form. The ansatz hjRj ($R\in \mathbb{C}{\backslash}\left\{0\right\}$) leads to

and the solutions R1,2 are

Equation (F4)

Our ansatz holds for all parameter choices of μ, Δ and t and R1,2 obey

Equation (F5)

Equation (F6)

The general form of hj is given by a superposition of R1 and R2

Equation (F7)

and n1,2 are fixed by the initial values. The calculation can be simplified by extending the sequence hj backwards with equation (F2), because h−1 = 0. The use of h−1 and h0 leads to

yielding the closed form of hj

We find the determinant of the Kitaev chain to be

Equation (F8)

for all values of $\mu ,\;t,\;{\Delta}\in \mathbb{R}$. The determinant does not vanish in general, due to equation (F4), but only for a specific combination of the parameters μ, t, Δ.

In the following we consider t and Δ to be fixed values of our choice and we search for the values of μ such that the determinant vanishes. The Fibonacci character of hN enables us to factorize the determinant [36, 37] and leads automatically to the zeros. The factorization follows from equation (F4) and the starting point is the square root:

We have to consider in general three cases

  • (a)  
    t2 ⩾ Δ2 and 4(t2 −Δ2) ⩾ μ2,
  • (b)  
    t2 ⩾ Δ2 and 4(t2 −Δ2) ⩽ μ2,
  • (c)  
    t2 ⩽ Δ2 and 4(t2 − Δ2) ⩽ μ2,

and we introduce the procedure in detail with the first scenario.

F.1. Case (a)

The most general form for μ is

Equation (F9)

where the function f(θ) accounts for all possible ratios of μ and $\sqrt{{t}^{2}-{{\Delta}}^{2}}$. The case (a) enforces the function f(θ) to be real valued, because both μ and $\sqrt{{t}^{2}-{{\Delta}}^{2}}$ are real. Further, we find that

Equation (F10)

since 4(t2 − Δ2) − μ2 ⩾ 0. Please note that equation (F10) needs only to hold for θ on a finite set. From all possible functions f(θ), a convenient choice is f = cos(θ). The reason behind our specific choice is the form of R1,2, because f leads in $\sqrt{4\;ab\;-{\mu }^{2}}$ to

and R1,2(f) become

Simplifications lead to

Let us focus on the determinant. We find ${R}_{1}^{j}-{R}_{2}^{j}$ to be

Consequently the determinant reads

Equation (F11)

and vanishes for θ = /(N + 1) (n = 1, ..., N) or t2 = Δ2. Since Δ, t and θ define together with f1 = cos(θ) the chemical potential, we find that the determinant of the Kitaev chain is zero if, and only if:

  • (a)  
    $\mu =2\;\sqrt{{t}^{2}-{{\Delta}}^{2}}\;\mathrm{cos}\left(\frac{n\;\pi }{N+1}\right)$,
  • (b)  
    μ = 0 and t2 = Δ2,

for n = 1, ..., N, t2 ⩾ Δ2 and for all N. A feature of odd N is the value n = N + 1/2 yielding θ = π/2, i.e. μ = 0 for all values of Δ, t for t2 ⩾ Δ2. In fact μ = 0 holds for odd N everywhere, as we already know from previous discussion in appendix B.

We found all zeros in case (a) and we continue with (b).

F.2. Case (b)

We follow the same way of argumentation as above, but we have to keep in mind that t2 − Δ2 ⩾ 0, and 4(t2 − Δ2) < μ2. The first step is to reshape the square root in R1,2

Equation (F12)

where we find a similar situation as in the previous scenario. Our ansatz is

Equation (F13)

where the function g(θ) is real and obeys

Equation (F14)

since μ2 ⩾ 4(t2 − Δ2). The candidates of our choice are g±(θ) = ±cosh(θ), where θ is real. The square root becomes now

and we find R1,2(g+) to be

Simplifications yield

and the determinant becomes:

Equation (F15)

The determinant vanishes only if t2 = Δ2, which by virtue of equation (F13) implies μ = 0, because the fraction of the hyperbolic sine functions is always positive. The use of g1,− = −cosh(θ) leads to equation (F15) again.

F.3. Case (c)

We consider here Δ2t2 and 4(t2 − Δ2) − μ2 ⩽ 0. We start by manipulating the square root in R1,2

Equation (F16)

Our ansatz is $\mu =2\;\sqrt{{{\Delta}}^{2}-{t}^{2}}\;v\left(\theta \right)$ with a real valued function v(θ), without further restrictions, because

in view of μ2 ⩾ −4(Δ2t2). The square root in R1,2 becomes in general

and one sees immediately that v(θ) = sinh(θ), $\theta \in \mathbb{R}$ is an appropriate choice. We find for R1,2 the form

where the negative sign in front of the exponential forces us to distinguish between even and odd N. The determinant reads finally

and it is never zero, except for Δ2 = t2 at μ = 0.

F.4. Discussion of completeness of all scenarios

In summary, for μ ≠ 0, we have only non trivial, zero determinants in case (a). How can one be sure that no zero is missed especially in the settings (b) and (c)? This follows immediately from equation (F8), because the determinant vanishes only if

Consequently we need first of all |R1| = |R2|. The second part is to find the proper phase factors and all of them lie on a circle with radius |R1| in the complex plane. We have found non trivial solutions only for scenario (a).

In total, we found all conditions $\mathrm{det}\left({\mathcal{H}}_{\text{KC}}\right)=0$. The general case is when the chemical potential is

Equation (F17)

with t2 ⩾ Δ2 and n = 1, ..., N, i.e. the chemical potential corresponds to the energy levels of a linear chain with hopping $\sqrt{{t}^{2}-{{\Delta}}^{2}}$. The case μ = 0 and t2 = Δ2 is included in equation (F17).

Further, the determinant of a Kitaev chain with odd number of sites is zero if μ = 0 for all values of Δ and t.

Appendix G.: The zero energy eigenstates

The presence of zero energy modes is marked by $\mathrm{det}\left({\mathcal{H}}_{\text{KC}}\right)=0$ and a natural question is to investigate their topological character, be it trivial or non-trivial. Hence, we have first to obtain these states. We use here again the SSH-like basis, e.g. the Hamiltonian from equation (23) for μ ≠ 0. We keep the notation for the eigenvector $\to {\psi }={\left({\to {v}}_{\alpha },\;{\to {v}}_{\beta }\right)}^{\text{T}}$ with

for even N, but unlike in the previous calculation both SSH-like chains are coupled now. We consider first N even, because the odd N solutions have the same shape, as it turns out later. Further, we derive the general eigenvector problem including even non zero modes. Solving $\left(\lambda \;{\mathbb{1}}_{2N}-{\mathcal{H}}_{\text{KC}}^{\text{SSH}}\right)\;\to {\psi }=\to {0}$ translates to

The reason to keep λ first inside the calculation is the diagonal structure of τ, τ and ${\mathbb{1}}_{N}$ as well as the entry structure of ${\to {v}}_{\alpha }$ and ${\to {v}}_{\beta }$, which enables us to identify easily the new contributions of $\tau \;{\to {v}}_{\beta }$ and ${\tau }^{{\dagger}}\;{\to {v}}_{\alpha }$ in comparison to the μ = 0, i.e. τ = 0, case from appendix C. The difficulty to write down $\left(\lambda \;{\mathbb{1}}_{2N}-{\mathcal{H}}_{\text{KC}}^{\text{SSH}}\right)\;\to {\psi }=\to {0}$ reduces to taking the correct signs of the μ terms. We have to solve (l = 1, ..., N − 1)

Equation (G1)

Equation (G2)

Equation (G3)

Equation (G4)

from ${\mathcal{H}}_{\alpha }\;{\to {v}}_{\alpha }+\tau \;{\to {v}}_{\beta }=\lambda \;{\to {v}}_{\alpha }$ and

Equation (G5)

Equation (G6)

Equation (G7)

Equation (G8)

from ${\mathcal{H}}_{\beta }\;{\to {v}}_{\beta }+{\tau }^{{\dagger}}\;{\to {v}}_{\alpha }=\lambda \;{\to {v}}_{\beta }$. Extending the sequences xl, yl, ${\mathcal{X}}_{l}$ and ${\mathcal{Y}}_{l}$ backwards leads to simplifications in the open boundary conditions

As we see from the particle–hole operator, see equation (79), an MZM requires either fully real or fully imaginary entries, which is not true for a generic solution of the eigenvector system equations (G1)–(G8) of the Kitaev Hamiltonian for λ ≠ 0. Thus, zero energy is essential for an MZM.

Zero energy has one advantage, because the chiral partner of a zero mode is itself a zero mode and superpositions of both will simplify the eigenvector problem even more. Acting with $\mathcal{C}$ from equation (80) on $\to {\psi }$, all yl (${\mathcal{X}}_{l}$) go into −yl ($-{\mathcal{X}}_{l}$), while all xl (${\mathcal{Y}}_{l}$) remain the same. Hence, ${\to {\psi }}_{A}{:=}\left(\to {\psi }+\mathcal{C}\;\to {\psi }\;\right)/2$ reads

and '|' marks the boundary of both SSH-like chains. Similar ${\to {\psi }}_{B}{:=}\left(\to {\psi }-\mathcal{C}\;\to {\psi }\;\right)/2$ is

As we see, we decomposed $\to {\psi }$ into ${\to {\psi }}_{A,B}$. The decomposition is optional, but ${\to {\psi }}_{A}$ (${\to {\psi }}_{B}$) has only non zero weight on A type (B type) Majorana positions ${\gamma }_{j}^{A}$ (${\gamma }_{j}^{B}$) in the SSH-like basis, as depicted in figure 10. Thus, ${\to {\psi }}_{A}$ obeys (S+)

while ${\to {\psi }}_{B}$ satisfies (S−)

and l runs from 1 to N − 1. As we see, (S+) turns into (S−) by exchanging a's and b's, μ into −μ and the standard letters into the calligraphic ones. Thus, we need only to solve one set of equations and the solution of the second follows immediately.

We focus on (S+) and we ignore the index limitations during the following calculation. Decoupling leads to

Equation (G9)

Equation (G10)

Fibonacci polynomials [3638]. The Binet form needs initial values and we have to think about the number of free entries we have here. These degrees of freedom are given by the dimension of the zero energy subspace, i.e. the number of zero energy states. So far, the chiral symmetry implies their pairwise presence, but not their absolute quantity. Each zero of the determinant is twice degenerated, as we see from equation (F8). Hence, we have in total only two zero energy modes and each has one unspecified entry. We choose x1 as a fixed number.

The naive choice would be to take x1, x2, ${\mathcal{Y}}_{1}$ and ${\mathcal{Y}}_{2}$ as initial values, where the last three are expressed in terms of x1. Instead we use the l = 0, 1 expressions and introduce x0 via (S+)

because x1 is our choice and ${\mathcal{Y}}_{0}=0$. We find x0 = x1b/a. The term y1 follows from (S+)

which reduces to y1 = −iμx1/b.

The Binet form follows again from a power ansatz xlzl. The fundamental solutions for both sequences are

We use equation (F17) to get $2\;ab-{\mu }^{2}=-2ab\;\mathrm{cos}\left(2\;\frac{n\pi }{N+1}\right)$ and we obtain

Finally, we have

with θn := /(N + 1). The general solution is given by the superposition of z1 and z2

and we find both coefficients with x1 and x0 to be

With this xl becomes

Using the expressions for z1,2, we find

or in the most compact form

Equation (G11)

Similar, we obtain ${\mathcal{Y}}_{l}$

Equation (G12)

which simplifies to

where −a/b is always positive since t2 ⩾ Δ2. The last step is to check if the open boundary conditions are satisfied. Obviously ${\mathcal{Y}}_{0}=0$ holds and we get for ${x}_{\frac{N}{2}+1}$ the form

Hence, the vector ${\to {\psi }}_{A}$ is an eigenvector of the Kitaev BdG Hamiltonian. The vector ${\to {\psi }}_{B}$ has the entries

Equation (G13)

and

Equation (G14)

where ${\mathcal{X}}_{1}$ is free to choose. The case of odd N is similar. We use

and $\to {\psi }={\left({\to {v}}_{\alpha },\;{\to {v}}_{\beta }\right)}^{\text{T}}$. The vectors ${\to {\psi }}_{A,B}=\left(\to {\psi }\;{\pm}\;C\to {\psi }\;\right)$ become now

Equation (G15)

Equation (G16)

As we see, we have to respect different index limitations for ${x}_{j}\;\left({\mathcal{X}}_{j}\right)$ and ${\mathcal{Y}}_{i}$ (yi), but apart from this small change everything else remains as in the even N case. The vector ${\to {\psi }}_{A}$ obeys now

with j = 1, ..., (N − 1)/2, i = 1, ..., (N − 3)/2 and ${\to {\psi }}_{B}$ satisfies

The only important change compared to the even N case are the new open boundary conditions, while the Fibonacci character remains. Hence, we ignore the index limitation during the calculation of those entries as in the even N case and we get the same results for xl, ${\mathcal{X}}_{l}$, yl and ${\mathcal{Y}}_{l}$, see equations (G11)–(G14).

The boundary conditions are satisfied, since ${y}_{0}={\mathcal{Y}}_{0}=0$,

and ${y}_{\frac{N+1}{2}}=0$. A last check for the odd N case is done by choosing n = (N + 1)/2, i.e. θn = π/2, which leads back to the old μ = 0 limit. Applying θnπ/2 on xl leads to

after some steps, while all ${\mathcal{Y}}_{l}\propto \mu $ are zero. Similar we find ${\mathcal{X}}_{l}$ from xl upon changing a with b, while yl = 0 for all l. Hence, we recover our result for the α (β) chain, see equations (71) and (72).

The remaining questions is whether these zero energy modes are Majorana zero modes or not. The use of the particle hole operator in the SSH-like basis from equation (79), i.e. complex conjugation, reveals that the expressions xl/x1, ${\mathcal{Y}}_{l}/{x}_{1}$, ${\mathcal{X}}_{l}/{\mathcal{X}}_{1}$ and ${y}_{l}/{\mathcal{X}}_{1}$ are always real quantities, for both even and odd N. Thus ${\to {\psi }}_{A}$ (${\to {\psi }}_{B}$) is an MZM if x1 (${\mathcal{X}}_{1}$) is either real or pure imaginary.

The MZM mode ${\to {\psi }}_{A}$ (${\to {\psi }}_{B}$) has non-zero weight on ${\gamma }_{j}^{A}$ (${\gamma }_{j}^{B}$). Superpositions of both vectors can be MZM too if the coefficients are chosen properly. For example $\to {\psi }={\to {\psi }}_{A}+{\to {\psi }}_{B}$ has no zero entry. Hence, it is a mixed type MZM (for the correct choice of x1 and ${\mathcal{X}}_{1}$).

Footnotes

  • Note that λ can be zero and it will for odd N. Hence, the standard formula to calculate the determinant of a partitioned 2 × 2 matrix can not be used here, because it requires the inverse of one diagonal block. We use instead Silvester's formula [34]: $\text{det}\left[\begin{matrix}\hfill A\hfill & \hfill B\hfill \\ \hfill C\hfill & \hfill D\hfill \end{matrix}\right]=\text{det}\left[AD-CB\right],$ where A, B, C, D are square matrices of the same size and the only requirement is [C, D] = 0.

  • In fact μ = 0 supports complex wavevectors too, but their real part has to be zero or μ/2, i.e. one has to use iq or $\frac{\pi }{2}+iq$.

Please wait… references are loading.
10.1088/1361-648X/ab8bf9