Brought to you by:
Paper The following article is Open access

Strongly interacting Majorana modes in an array of Josephson junctions

and

Published 31 December 2012 © IOP Publishing and Deutsche Physikalische Gesellschaft
, , Citation Fabian Hassler and Dirk Schuricht 2012 New J. Phys. 14 125018 DOI 10.1088/1367-2630/14/12/125018

1367-2630/14/12/125018

Abstract

An array of superconducting islands with semiconducting nanowires in the right regime provides a macroscopic implementation of Kitaev's toy model for Majorana wires. We show that a capacitive coupling between adjacent islands leads to an effective interaction between the Majorana modes. We demonstrate that, even though strong repulsive interaction eventually drives the system into a Mott insulating state, the competition between the (trivial) band insulator and the (trivial) Mott insulator leads to an interjacent topological insulating state for arbitrary strong interactions.

Export citation and abstract BibTeX RIS

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

1. Introduction

Majorana zero modes are fermions which are their own antiparticles. They are believed to exist as effective particles in the middle of the gap of a topological superconductor. The recent interest in Majorana zero modes originates in their non-Abelian exchange statistics [1], which is the basis for potential applications in quantum computation [2]. Based on the theoretical proposal [3] to realize these exotic states in semiconducting nanowires with strong spin–orbit coupling in a magnetic field and in the proximity of a conventional (nontopological) superconductor, recent experimental progress has shown signatures of Majorana zero modes in the tunnelling conductance of normal conducting–superconducting [4] and superconducting–normal conducting–superconducting systems [5]. More information and references on the fast developing field can be found in the recent reviews [6].

As the nanowire is one dimensional (1D), interaction effects become important. On the one hand, employing field theoretical methods, Gangadharaiah et al [7] have argued that strong electron–electron interactions generically destroy the topological phase by suppressing the superconducting gap. On the other hand, using a combination of analytical and numerical methods, Stoudenmire et al [8] have shown that repulsive interactions significantly decrease the required Zeeman energy and increase the parameter range for which the topological phase exists. It is believed that the origin of these effects is an interaction-driven renormalization of the Zeeman gap [8, 9]. Furthermore, it has been shown that in helical liquids the scattering processes between the constituent fermion bands open gaps, which in turn lead to a stabilization of the Majorana states against interactions [10] and that (an odd number of) Majorana zero modes are in fact stable against general interactions [11]. For further studies of the effect of electron–electron interactions on Majorana zero modes in nanowires and two-chain ladders, see [12].

Recently, a macroscopic version of the Kitaev chain [13], a toy model for a nanowire supporting noninteracting Majorana modes, was proposed [14] in a 1D array of topological superconducting islands. Its advantage over microscopic implementations is that individual parameters of the effective Hamiltonian can be potentially tuned in situ. Here, we show that additional capacitances between adjacent islands lead to an effective interaction between the low-energy Majorana degrees of freedom. We present the phase diagram of the system which demonstrates that sufficiently strong repulsive interactions will drive the system from the topologically trivial phase into the topological phase supporting Majorana zero modes before eventually leading to a Mott insulating state. We discuss how the parameters of the system can be tuned by changing the gate voltages and exploit this to propose the detection of the different phases and phase boundaries in a tunnelling experiment.

2. The model

We discuss a system consisting of a 1D array of N superconducting islands (see figure 1). Because of the proximity-coupled semiconducting nanowire each of the islands has two midgap Andreev states, i.e. Majorana modes, located at the ends of the nanowire [3]. We will denote with γka and γkb the two Majorana operators on island k associated with these zero modes. The Majorana operators are Hermitian γμ = γμ and fulfil the Clifford algebra {γμ,γν} = 2δμν. The total Lagrangian L = T − VJ − VM of the system consists of three terms. Coupling of the Majorana modes on nearby islands leads to the term [13, 15] $ V_{\mathrm {M}} = \mathrm {i}\,E_{\mathrm {M}} \sum _{k=1}^N \gamma _{kb} \gamma _{k+1a} \cos [(\phi _{k+1} - \phi _k)/2]$ where ϕk is the superconducting phase of the kth island. Here and in the following, we assume for convenience periodic boundary conditions such that islands 1 and N + 1 are equivalent. Apart from the Majorana modes, the term discussed above has the additional degrees of freedom ϕk due to the condensate of Cooper pairs. Similar to [14], we eliminate these by connecting each superconducting island with a strong Josephson junction to a common (ground) superconductor. This fixes the superconducting phases (up to some quantum phase-slips discussed below) and is described by the Hamiltonian $V_{\mathrm {J}} = E_{\mathrm {J}} \sum _{k=1}^N (1- \cos \phi _k);$ here, EJ = ℏIc/2e is the effective Josephson coupling of each of the Josephson junctions with critical current Ic. In the limit EJ ≫ EM the junctions effectively pin the phases of all superconducting islands to a common value ϕk ≡ 0. As a result, VM reduces to the pure Majorana coupling

Equation (1)

Finally, the kinetic term $T = (\hbar ^2/8e^2) \sum _{k=1}^N [ C (\dot \phi _{k+1} - \dot \phi _k)^2 + C_{\mathrm {G}} \dot \phi _k^2 ] + (\hbar /2e) \sum _{k=1}^N q \dot \phi _k $ occurs due to the capacitive couplings, where C denotes the capacitance between neighbouring islands and CG = Cg + CJ the (total) capacitance to the ground. Here, CJ is the capacitance of the strong Josephson junction and Cg the capacitance to a common back gate at voltage Vg with respect to the ground superconductor. Apart from the charging energy, the back gate introduces a term proportional to the induced charge q = CgVg which is tunable with single-electron precision [16] via Vg and whose effect will become important later. The typical capacitive energy scale EC = e2/2CΣ of single islands depends on the total capacitance CΣ = 2C + CG.

Figure 1.

Figure 1. A 1D array of superconducting islands (light grey) coupled via strong Josephson junctions (EJ) to a common ground superconductor. Each island contains a pair of Majorana zero modes (white dots) at the end points of a semiconducting nanowire (dark grey). The tunnel coupling of individual electrons between the superconducting islands is proportional to the energy scale EM. A common gate voltage Vg can be used to tune the relative strength of the different terms in the Hamiltonian. The capacitive couplings between the elements are denoted by C, CJ and Cg, respectively.

Standard image

3. Mapping on an effective spin model

As we have seen, the strong coupling to the ground superconductor pins the superconducting phase differences to ϕk ≡ 0 and thus changes the energy due to the Majorana modes from VM to HM. The effect of the strong coupling to the ground superconductor on the charging energy T is more subtle. We first present the results for C = 0 before extending them to nonzero C: as charge and phase are conjugate variables the pinning of the phases ϕk greatly reduces the effect of charging [14]. An effective charging energy still arises due to quantum phase slips through the junctions. For example, changing ϕk from 0 to 2π leads to a charging energy (EJ ≫ EC) [17]

Equation (2)

Here, the tunnelling amplitude is given by ΓΔ ≃ E1/4CE3/4JeSΔ with $S_\Delta = \hbar ^{-1} \int \!\mathrm {d}\tau \,L_E =\sqrt {8E_{\mathrm {J}}/E_C}$ the dimensionless Euclidean action along the classical trajectory γΔ: ϕk∈[0,2π] in the inverted potential with LE(τ) = −L(t = i τ).4 The cosine term in (2) occurs due to the Aharonov–Casher interference of the two tunnelling paths ϕk = 0 → 2π and ϕk = 0 → − 2π which lead to an indistinguishable final state and thus interfere with a phase difference depending on the total induced charge q + enk [14, 18]. The term $n_k = \frac 12 (1- \mathcal {P}_k)\in \{0,1\}$ is the contribution to the charge due to the parity $\mathcal {P}_k = \mathrm {i}\, \gamma _{ka} \gamma _{kb} \in \{-1,1\}$ of the number of electrons on the superconducting island encoded in the state of the Majorana zero modes [14, 15]. Equation (2) is a chemical potential term: for large ΓΔ, all the fermionic states are either filled or empty and the system is in a band-insulating state.

Going away from this special point and introducing a finite cross-capacitance parameterized by η = 2C/CΣ with η∈[0,1], the classical path for a phase slip in ϕk does not only involve ϕk but also the other phases. To lowest nonvanishing order in η, we only need to take into account the change in ϕk−1 and ϕk+1 and obtain $S_\Delta = \sqrt {8 E_{\mathrm {J}}/E_C} [1 + (\pi ^2 -12)\eta ^2/96]$ , which is accurate to more than two digits all the way up to η = 1 as numerics confirms. Additionally, the effective capacitance coupling between two islands becomes important. This term is generated by simultaneous phase slips of the two phases ϕk and ϕk+1 of neighbouring islands. Due to the symmetry between island k and k + 1, we have ϕk = ϕk+1 for the classical path γU: ϕk = ϕk+1∈[0,2π] with the corresponding Euclidean action $S_U= \sqrt {16 (2- \eta ) E_{\mathrm {J}}/E_C}$ . The Aharonov–Casher interference in this case leads to the tunnelling amplitude ΓU ≃ E1/4CE3/4JeSU depending on the total charge 2q + e(nk + nk+1) of the two islands involved. Thus we obtain the interaction term

Equation (3)

We stress that (3) is an interaction term involving four Majorana operators and that the amplitude is modulated by twice the induced charge q compared to (2). We call U > 0 repulsive interaction as it prefers having an occupied site next to an empty one. For strong repulsive interactions the system is driven into a Mott-insulator or equivalently commensurate charge density wave (CDW) state.

The complete low-energy Hamiltonian in the limit EJ ≫ EM,EC is given by $H_{{\scriptscriptstyle \rm ANNNI}} = \sum _{k=1}^N (H_{Ck} + H_{Uk}) + H_{\mathrm {M}}$ , which constitutes the transverse axial next-nearest-neighbour Ising (ANNNI) model [19] as can be seen by performing a Jordan–Wigner transformation $\mathcal {P}_k = \mathrm {i}\, \gamma _{ka} \gamma _{kb} = \sigma ^z_k$ , i γkbγk+1a = σxkσxk+1, resulting in the spin Hamiltonian

Equation (4)

Here, the σx,y,zk are Pauli matrices, Δ = ΓΔcos(πq/e) and U = ΓUcos(2πq/e). We note that the energies Δ and U can be tuned via the charge q induced on the superconducting islands through the gate voltage Vg. As the fermionic Hamiltonian without the interaction term proportional to U has been intensively studied previously [13], we focus here on the effects of the interaction term. In our set-up, this term is most important in the case of C ≫ CG where η ≈ 1 and $\Gamma _U \approx \Gamma _\Delta \exp [-1.25 \sqrt {E_{\mathrm {J}}/E_C}]$ , which can be as large as 0.3 for EJ ≈ EC; note that in this regime the actions SΔ and SU are still much larger than 1 such that the semiclassical approximation employed above is valid.

4. Phase diagram

We first note that the system (4) is invariant under EM → − EM as well as Δ → − Δ due to the transformations σx,yk → (−1)kσx,yk and σx,zk → − σx,zk, respectively. Thus without loss of generality, we assume in the following EM,Δ > 0. The phase diagram of the spin model (4) contains four phases:5 a paramagnetic phase (PM) with a unique ground state with 〈σzk〉 < 0, 〈σxk〉 = 0; an antiferromagnetic phase (AFM) with doubly degenerate ground state and 〈σzk〉 = 0, 〈σxk〉∝(−1)k; an 'antiphase' (AP) with a doubly degenerate ground state with 〈σzk〉∝(−1)k; and a 'floating phase' (FP) between the AFM and the AP. For Δ = 0, the duality transform of the model (4) is a sum of two quantum Ising chains, while for the noninteracting case it reduces to a single quantum Ising chain.

As we have shown above in our realization of the ANNNI model the parameters Δ and U can be easily tuned via a gate voltage. On the other hand, the coupling EM is determined by the overlaps of the Majorana wave functions on neighbouring islands and thus fixed by the geometry of the array. Hence it is natural to consider the phase diagram as a function of Δ/EM and U/EM, which is shown in figure 2. For fixed Δ/EM > 1 and sufficiently small interaction U the system is in a trivial (band-insulating) phase corresponding to the PM in the effective spin model, which is characterized by a unique ground state with $\langle \mathcal {P}_k\rangle <0$ . By increasing U we cross into a topological phase (corresponding to the AFM) with $\langle \mathcal {P}_k\rangle =0$ and two degenerate ground states | ± 〉, distinguished by the (total) fermion parity $\prod _{k}\mathcal {P}_k |\pm \rangle = \pm |\pm \rangle $ . The parity protection (also called topological protection) originates from the fact that any fermionic perturbation conserves the fermion parity and thus cannot mix the states | ± 〉. The phase transition between the trivial and topological phases is, for U > 0, located at

Equation (5)

In particular, we find that Δ = EM + 3U/2 for |U| ≪ EM and Δ = 2U for U ≫ EM.

Figure 2.

Figure 2. Phase diagram of the Josephson junction array (4). The phase diagram is invariant under Δ → − Δ. The topological phase is characterized by a doubly degenerate ground state. In figure 3, we show the tunnelling conductance along the paths c1 and c2. Inset: Gap ΔT between the (nearly) degenerate ground states (δ ≲ 0.02 EM) and the first excited state along the blue dashed line as determined by exact diagonalization.

Standard image

At large positive U we eventually enter incommensurate and commensurate CDW (Mott-insulator) states corresponding to the FP and the AP, respectively6. This region of the phase diagram cannot be reached as long as the induced charge q on the islands is homogeneous. However, replacing the common back gate by individual gates for each island yields the model (4) with site-dependent parameters Δi = ΓΔ cos(πqi/e) and Ui = ΓU cos(π(qi + qi+1)/e) where qi = GgVig denotes the induced charge on island i. Now using qi = (−1)iq one can enter the Mott phase for q → e/2. Note that the two degenerate ground states in the Mott insulator phase have the same fermion parity and thus are not parity protected.

The main feature of the phase diagram is its strong anisotropy under U → − U. While negative interactions suppress the topological phase, for U > 0 the ordering tendencies of the first and second terms in (4) compete with each other. In particular, starting from a noninteracting point in the trivial phase, i.e. U = 0 and Δ > EM, competition between the band insulator and the Mott insulator will drive the system into the topological phase irrespective of the value of Δ/EM.

As discussed above the topological phase is characterized by the existence of a doubly degenerate ground state with different fermion parities. In a finite system this degeneracy is lifted and the value of the resulting gap δ depends on the number of islands N as well as the system parameters Δ/EM and U/EM. On the other hand, the existence of Majorana end modes is protected by the gap between the (nearly) degenerate ground states and the first excited state, which we denote by ΔT. This gap is given by ΔT = 4EM at Δ = U = 0 and decreases when going away from this point. However, as we show in the inset of figure 2, it remains of the order of EM and significantly larger than δ for U ≠ 0 as long as one stays away from the phase boundaries.

5. Tunnelling conductance

After presenting the phase diagram, we now turn to its experimental signatures in the tunnelling conductance. The parameters Δ and U can be directly tuned through the induced charge q on the islands via a common back gate voltage Vg making it possible to choose the parameters such that the path crosses one or more phase boundaries.

Specifically, we consider the path c1(q) = (ΓU cos(2πq/e),ΓΔ cos(πq/e)) with ΓU = 0.3 ΓΔ = 0.54 EM such that the starting and end points at q = 0 and e/2 lie in the topological phase, while the path enters the trivial phase in between (see the red curve in figure 2). In the following, we consider an open chain of N islands. Following [20], we couple the system to an electronic lead with the tunnel Hamiltonian $H_{\mathrm { T}} = t \gamma _{1a}\sum _{k\sigma } (c_{k\sigma } - c^{\dagger} _{k\sigma })$ , where ckσ are the annihilation operators of the electrons in the lead and t is the tunnelling amplitude. We assume a constant density of states ρ0 in the lead such that the (bare) tunnelling probability is given by Γ0 = 2πt2ρ0. We have calculated the differential Andreev conductance G(V ) using exact diagonalization for an open chain of length N = 24 taking only the two lowest energy states (with different fermion parities) into account7. In figure 3(a), the broadened conductance

Equation (6)

is plotted along c1; the broadening $\Delta \omega \simeq \max (eV,k_{\mathrm {B}} T)$ is given by the maximum of bias voltage V and temperature T. For strong coupling Γ0 to the lead as well as small broadening Δω, the phase boundaries are clearly visible as points where the conductance jumps from one to zero and vice versa. Weaker coupling to the lead will lead to a suppression of the conductance in the topological phase, i.e. unitary conductance cannot be observed. On the other hand, a larger broadening will eventually smear out all transitions. Thus in order to enable an experimental detection of the phase boundaries the energy broadening should be small (in units of EM) while the coupling to the lead has to be sufficiently strong. Recent experiments on proximity coupled nanowires [4] indicate that a Majorana coupling EM/kB ≃ 100 mK is realistic, which is well in the range of experimental accessible temperatures. The Majorana coupling sets the topological gap ΔT (see the inset of figure 2) as the other couplings (EJ and EC) can be designed in a large parameter range [21].

Figure 3.

Figure 3. Tunnelling conductance along the paths (a) c1 and (b) c2 shown in figure 2. The arrows indicate the approximate phase boundaries.

Standard image

In order to study the regime of strong interactions we use the setup with individual back gates and induced charges qi = (−1)iq. In this setup, we consider the path c2(q) = (ΓUΔ cos(πq/e)) with ΓU = 0.3 ΓΔ = 1.5 EM (see the green curve in figure 2). The conductance $\bar {G}$ along c2 is shown in figure 3(b). At q = 0 the path starts deep in the band-insulator phase and enters the topological phase at q ≈ 0.2 where the conductance becomes nonzero. When further increasing q we observe weak oscillations which are due to the finite-size lifting δ of the ground state degeneracy. For larger values of q the conductance stays zero in the incommensurate and commensurate CDW phases with the exception of the phase transition between them, where the conductance is nonzero due to finite-size effects.

Above we have shown how to realize the 1D ANNNI model using Josephson junction arrays and how to map out its phase diagram by measuring the tunnelling conductance. In this sense our proposed setup constitutes a quantum simulator for the 1D ANNNI model. In particular, the experimental control over the system parameters like the gate voltages allows us to study the effects of disorder on the phase diagram.

6. Relation to nanowires

Interacting electrons in proximity-coupled semiconducting nanowires are described by the microscopic Hamiltonian [8]

Equation (7)

with Ψ(x) = (Ψ(x),Ψ(x))T the electron field operator, m the electron mass, μ the chemical potential, α the strength of the spin–orbit coupling, EZ the Zeeman energy due to the applied magnetic field, Δs the s-wave pairing amplitude and U0 the (short-range) Coulomb interaction. For sufficiently strong Zeeman energy (compared to the other energy scales), we only need to consider a single band similar to the ANNNI model discussed above. However, projecting the Hamiltonian (7) onto a single band strongly reduces the effect of the interaction. Specifically, we find for the effective interaction strength in the single-band model U/EM = mU0α2/ℏ2LE2Z ≪ 1 with L the length of the nanowire. In this way, interacting nanowires subject to a strong Zeeman field are always in the weak coupling regime [8]. In contrast, as we showed above, the strong interaction regime for spinless fermions is readily accessible in the case of nanowires in Josephson junction arrays.

7. Conclusions

We have analysed a 1D array of Josephson junctions featuring Majorana modes, where capacitances between adjacent islands lead to interactions between the Majorana modes. We have shown that repulsive interactions generically facilitate the topological phase due to their competition with the on-site charging energies. Finally, we have proposed a tunnelling experiment to detect the phase boundaries.

Acknowledgments

We have benefited from discussions with Eran Sela and Kirill Shtengel and thank Bernd Braunecker for useful comments. This work was supported by the Alexander von Humboldt Foundation (FH) and the DFG through the Emmy–Noether program (DS).

Appendix.: Derivation of the quantum phase slip rate

In this appendix, we present more information about the derivation of the rates ΓΔ and ΓU. Starting with ΓΔ, we are interested in the event that the phase ϕk of a single island changes by 2π. The relevant tunnelling matrix element tΔ∝〈2π|e−iHt|0〉 can be evaluated in the path-integral formalism by going to imaginary time τ = i t, cf [2325],

Equation (A.1)

subject to the boundary condition ϕk(0) = 0 and $\Delta \phi _k=\phi _k(\infty ) -\phi _k(0) \in 2\pi \mathbb {Z}\setminus \{0\}$ . To exponential accuracy, the path integral is dominated by the classical paths $\phi _k^{{\mathrm {\scriptstyle cl}},n}$ which minimize the action $S_E=\hbar ^{-1}\int \!\mathrm {d}\tau \,L_E$ , i.e.

Equation (A.2)

where n is an index enumerating the different paths in the case when there are different minima of the action.

For η = 0, the relevant part of the action is well approximated by (keeping only terms which depend on ϕk)

Equation (A.3)

where ' denotes the derivative with respect to τ and we have neglected the potential proportional to EM ≪ EJ. As the action does not depend directly on τ, the energy along the classical path minimizing the action is conserved,

Equation (A.4)

For τ = 0 we have T = VJ = 0 such that E = 0 in our case.

We can express the kinetic energy in terms of the potential and obtain

Equation (A.5)

with which we can get an alternative expression for the measure (the capacitance matrix acts as a metric)

Equation (A.6)

Due to the conservation of energy, we can go over to the Euler–Maupertuis action $S_0 = S_E+ \int \!\mathrm {d}\tau \, E$ (note that in our case E = 0 such that S0 is in fact equal to SE) which can be rewritten employing (A.6) as

Equation (A.7)

which is independent of imaginary time and only depends on the path chosen [26].

The action is minimized for Δϕk = ± 2π as each additional phase slip by 2π increases S0. The expression corresponding to the first term in (A.7) is independent of ±. The classical path corresponds to increasing ϕk by 2π such that, cf [17],

Equation (A.8)

We obtain the final result

Equation (A.9)

valid up to exponential accuracy. In the main text, we use the result tΔ = ΓΔcos(πq/e) with $\Gamma _\Delta \simeq E_C^{1/4} E_{\mathrm {J}}^{3/4} \mathrm {e}^{-\sqrt {8 E_{\mathrm {J}}/E_C}}$ . In fact, the charge q should be replaced by q + enk. The reason is that due to the Majorana term the action is not 2π but only 4π periodic in ϕk. In the calculation above, we however assume the action to be 2π periodic. In fact, the action can be made 2π periodic by a gauge transformation at the expense of replacing qq + enk. More information on this rather subtle point can be found in [14, 15].

The prefactor E1/4CE3/4J of ΓΔ depends on the shape of the potential close to the turning points ϕk ≈ 0,±2π and cannot be obtained in our simple semiclassical analysis which only captures the physics up to exponential accuracy. However, the scaling of the prefactor can be obtained from summing up the instanton contributions [24] or by matching it to the exact solution of the Mathieu equation [17]. In our case, the potential is always given by $V_{\mathrm {J}} \simeq \frac 12 E_{\mathrm {J}} \phi _k^2$ for ϕk ≪ 1 such that the same prefactor E1/4CE3/4J (from the Mathieu equation) appears for all the tunnelling amplitudes.

In the case η ≠ 0, it is important to note that the phases on the different islands do not completely decouple. In lowest order in η, we need to take the phases ϕk±1 on the islands k ± 1 into account. Due to the symmetry of the problem, we know that ϕk−1(τ) = ϕk+1(τ). The relevant part of the action reads

Equation (A.10)

Following the same line of calculation as going from (A.3) to (A.7), we obtain

Equation (A.11)

where we expressed the path by giving ϕk+1 as a function of ϕk. As we are interested in processes where ϕk changes by Δϕk = ± 2π, we need to find $\phi _{k+1}(\phi _k)\colon [0,\pm 2\pi ] \mapsto \mathbb {R}$ with ϕk+1(0) = ϕk+1(±2π) such that the action is minimized. We will find the solution which corresponds to Δϕk = 2π below. The second solution with Δϕk = −2π can be obtained via the symmetry ϕi↦ − ϕi,∀i of the Lagrangian.

As the second term in (A.10) is independent of the path (it depends only on the boundary condition), we only need to minimize the first term. The extremum is attained when the function ϕk+1(ϕk) fulfils the Euler–Lagrange equation

Equation (A.12)

To first order in η (assuming that $\phi_{k+1} \in \Or(\eta)$ ), the equation assumes the form

Employing the substitution ϕk+1(ϕk) = tan(ϕk/4)f(ϕk) reduces the equation to a first-order equation in f' of the form

Equation (A.13)

which can be integrated with the solution

Equation (A.14)

Plugging the solution into (A.11) and retaining the first nonvanishing term in η yields

Equation (A.15)

Summing up the two contributions with Δϕk = ± 2π, we obtain a term proportional to cos(πq/e) as before, with the proportionality constant given by $\Gamma _\Delta = E_C^{1/4} E_{\mathrm {J}}^{3/4} \mathrm {e}^{- \sqrt {8 E_{\mathrm {J}}/E_C} [1 + (\pi ^2-12)\eta ^2/96 ]}$ .

For η ≠ 0, we get additionally a next-nearest-neighbour interaction due to phase slips where both ϕk and ϕk+1 change by 2π. In fact, due to the symmetry of the problem, we can set ϕk+1(τ) = ϕk(τ). The terms of the action which change with ϕk and ϕk+1 are given by (A.3) for each of the islands and the additional contribution of the cross capacitance. Thus, we have

Equation (A.16)

which leads to

Equation (A.17)

As Δϕk+1 = Δϕk (because the two phases slip together), the tunnelling amplitude thus assumes the form ΓU cos(2πq/e) with $\Gamma _U = E_C^{1/4} E_{\mathrm {J}}^{3/4} \mathrm {e}^{- \sqrt {16(2-\eta ) E_{\mathrm {J}}/E_C}}$ .

Footnotes

  • Because VJ ≫ VM, we take only the potential VJ into account when calculating the tunnelling action SE.

  • The Hamiltonian (4) is brought to a standard form by performing the duality transformation $\mu _k^x=\prod _{j<k}\sigma _j^z$ , μzk = σxkσxk+1. The phase diagram of the resulting model including expressions for the phase transitions has been worked out in [19, 22] in the parameters κ = U/Δ and EM/Δ.

  • We note that the existence of the FP, and thus the incommensurate CDW state, at small Δ/EM has not yet been fully established [22].

  • The two-level approximation is appropriate in the topological phase (where we have two levels separated by ΔT ≫ δ from the rest), as well as in the trivial phase (where there is a unique ground state) and in the Mott phase (where there are two degenerate ground states with the same fermion parity) as δ ≳ Δω in the latter cases such that the current vanishes.

Please wait… references are loading.
10.1088/1367-2630/14/12/125018