This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Majorana differential shot noise and its universal thermoelectric crossover

Sergey Smirnov P. N. Lebedev Physical Institute of the Russian Academy of Sciences, 119991 Moscow, Russia 1) [email protected]
2) [email protected]
3) [email protected]
Abstract

Nonequilibrium states driven by both electric bias voltages VV and temperature differences ΔT\Delta T (or thermal voltages eVTkBΔTeV_{T}\equiv k_{B}\Delta T) are unique probes of various systems. Whereas average currents I(V,VT)I(V,V_{T}) are traditionally measured in majority of experiments, an essential part of nonequilibrium dynamics, stored particularly in fluctuations, remains largely unexplored. Here we focus on Majorana quantum dot devices, specifically on their differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V, and demonstrate that in contrast to the differential electric or thermoelectric conductance, I(V,VT)/V\partial I(V,V_{T})/\partial V or I(V,VT)/VT\partial I(V,V_{T})/\partial V_{T}, it reveals a crossover from thermoelectric to pure thermal nonequilibrium behavior. It is shown that this Majorana crossover in S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V is induced by an interplay of the electric and thermal driving, occurs at an energy scale determined by the Majorana tunneling amplitude, and exhibits a number of universal characteristics which may be accessed in solely noise experiments or in combination with measurements of average currents.

I Introduction

Topological superconductors interacting with nanoscopic setups provide a feasible technological platform to entangle a specific quantum core of these setups with non-Abelian Majorana bound states (MBSs) Yu. Kitaev (2001); Alicea (2012); Leijnse and Flensberg (2012); Sato and Fujimoto (2016); Aguado (2017); Lutchyn et al. (2018); Val’kov et al. (2022) imitating particle-antiparticle paradigm of the Abelian Majorana fermions Majorana (1937) known in the particle physics. Such Majorana entangled setups are understood (without a strict relation to the specific meaning of ”entanglement” in quantum information) as those where Majorana and non-Majorana degrees of freedom are coupled via a certain quantum mechanical mechanism. They represent a special class of condensed matter systems which are very attractive both from theoretical and experimental perspectives. On one side, they admit an observation of diverse physical phenomena governed essentially by Majorana entangled states and, on the other side, they may function as elementary blocks integrated in various fault tolerant topological quantum computing Yu. Kitaev (2003) schemes designed to process the quantum nonlocality supported by MBSs.

Nanoscopic setups where MBSs are entangled with the quantum degrees of freedom involved in experimental measurements reveal various remarkable characteristics many of which may be accessed in quantum transport experiments. Such experiments deal with nonequilibrium states which may be generated by bias voltages VV or temperature differences ΔT\Delta T, expressed equivalently through the corresponding thermal voltages VTV_{T}, defined as eVTkBΔTeV_{T}\equiv k_{B}\Delta T. Here Majorana features are predicted within the framework of mean currents I(V)I(V) induced by mostly bias voltages Liu and Baranger (2011); Fidkowski et al. (2012); Prada et al. (2012); Pientka et al. (2012); Lin et al. (2012); Lee et al. (2013); Kundu and Seradjeh (2013); Vernek et al. (2014); Ilan et al. (2014); Cheng et al. (2014); Lobos and Das Sarma (2015); Peng et al. (2015); Sharma and Tewari (2016); van Heck et al. (2016); Sarma et al. (2016); Lutchyn and Glazman (2017); Liu et al. (2017); Huang et al. (2017); Liu et al. (2018); Lai et al. (2019); Tang and Mao (2020); Zhang and Spånslätt (2020); Chi et al. (2021); Wang and Huang (2021); Galambos et al. (2022); Jin and Li (2022); Zou et al. (2023a) or, to a lesser extent, mean currents I(V,VT)I(V,V_{T}) induced by also temperature differences Leijnse (2014); López et al. (2014); Khim et al. (2015); Ramos-Andrade et al. (2016); Smirnov (2020); He et al. (2021); Giuliano et al. (2022); Buccheri et al. (2022); Majek et al. (2022); Bondyopadhaya and Roy (2022); Zou et al. (2022). Experimental efforts on mean currents I(V)I(V) induced by bias voltages Mourik et al. (2012); Nadj-Perge et al. (2014); Wang et al. (2022) are aimed to measure the differential conductance I(V)/V\partial I(V)/\partial V. Of particular interest here is the zero bias limit of the differential conductance (linear conductance) which should attain a certain universal value predicted theoretically for a specific Majorana entangled setup. Although such mean current experiments are well developed and should be performed in the first place, unfortunately, they may be controversial Yu et al. (2021); Frolov (2021) in detecting MBSs and, as a consequence, other types of quantum transport measurements, or perhaps sequences of measurements Ziesen et al. (2023), are currently in demand. Particularly, one is interested in those physical observables which demonstrate in a given setup a Majorana driven behavior which is qualitatively different from the behavior of the mean currents measured in the same Majorana entangled setup.

An attractive quantum transport alternative to the mean value of a current flowing through a nanoscopic setup is to study the random deviations of this current from its mean value, that is the current fluctuations, characterized, for example, by the shot noise S>S^{>}. Here majority of Majorana shot noise proposals assume nonequilibrium states originating from bias voltages Liu et al. (2015a, b); Haim et al. (2015); Valentini et al. (2016); Zazunov et al. (2016); Smirnov (2017); Jonckheere et al. (2019); Smirnov (2019a); Manousakis et al. (2020); Feng and Zhang (2022); Smirnov (2022) and explore the behavior of S>(V)S^{>}(V) at small and large VV. As in mean current experiments, where the differential conductance I(V)/V\partial I(V)/\partial V provides an access to an averaged Majorana universality, the differential shot noise S>(V)/V\partial S^{>}(V)/\partial V allows one to reveal a universal fluctuation behavior governed by Majorana entangled states.

One may also avoid resorting to nonequilibrium behavior and address Majorana entangled states in corresponding equilibrium nanoscopic setups by means of quantum thermodynamic tools such as the entropy of these setups Smirnov (2015); Sela et al. (2019); Smirnov (2021a, b); Ahari et al. (2021). Recent experimental and theoretical activities Hartman et al. (2018); Kleeorin et al. (2019); Pyurbeeva and Mol (2021); Child et al. (2022a); Han et al. (2022); Pyurbeeva et al. (2022); Child et al. (2022b) on the entropy of nanoscale and mesoscale systems demonstrate that this engrossing approach may become a powerful and univocal technique which will not be subject to further controversy similar to the one about the Majorana differential conductance.

Nevertheless, presently quantum transport is a more appealing framework within which experimentalists have at their disposal well established technologies verified in diverse nanoscopic setups for a long period of time. Moreover, quantum transport techniques have a wider space of control due to numerous additional parameters utilized to maintain various kinds of nonequilibrium states in which a broad spectrum of physical observables is available for performing experiments. Thus applying the quantum transport framework to Majorana entangled nanoscopic setups provides vast freedom in exploring Majorana phenomena in nonequilibrium. In particular, among measurements of other physical observables, shot noise experiments in various nonequilibrium states are expected to qualitatively enrich the existing results on mean currents in Majorana entangled nanoscopic setups.

Here we focus on the shot noise in nonequilibrium states produced by bias voltages VV and thermal voltages VTV_{T} in a quantum dot (QD) whose degrees of freedom are entangled with MBSs of a topological superconductor. Specifically, we explore the differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V which, as has been discussed above, inspects universal Majorana fluctuation behavior. So far it is not much known about this physical observable when both VV and VTV_{T} excite competing current flows in a Majorana setup. Indeed, whereas the differential thermoelectric shot and quantum noise, S>(V,VT)/VT\partial S^{>}(V,V_{T})/\partial V_{T}, have been addressed Smirnov (2018, 2019b) in presence of both bias voltages and temperature differences, the differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V remains to a large extent unexplored for Majorana entangled setups in nonequilibrium states driven by both VV and VTV_{T}. It should be noted that in nonequilibrium states induced only by bias voltages VV the differential shot noise has been studied in combination with the differential conductance. In particular, in Ref. Cao et al. (2023) it is demonstrated that in presence of MBSs a dip of the differential shot noise is always accompanied by a peak of the differential conductance. This behavior has also been observed earlier in Ref. Smirnov (2022) (see its Fig. 4, namely, the insets of the upper panel). As mentioned above, it is important to find for a given Majorana entangled setup physical observables whose behavior has a character qualitatively different from the one of the mean current or its derivative physical quantities such as the differential conductance I(V,VT)/V\partial I(V,V_{T})/\partial V or differential thermoelectric conductance I(V,VT)/VT\partial I(V,V_{T})/\partial V_{T} whose behavior may be obtained in the same setup. We demonstrate that the differential shot noise is one of such physical observables which is distinguished by the presence of a crossover from a thermoelectric to pure thermal nonequilibrium behavior. It is shown that whereas the differential shot noise passes through its crossover, the differential conductance and differential thermoelectric conductance do not exhibit any crossover or any other peculiarity. Thus, in contrast to I(V,VT)/V\partial I(V,V_{T})/\partial V and I(V,VT)/VT\partial I(V,V_{T})/\partial V_{T}, the differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V brings a nonequilibrium energy scale having a pure fluctuation nature meaning that it cannot be revealed within measurements limited only by the mean current. Besides being of fundamental interest, the energy scale associated with the nonequilibrium crossover in S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V is shown to be of practical importance in expressing quantitatively the fluctuation universality of Majorana entangled states via a number of measurable universal ratios which would be of interest for future experiments.

The paper is organized as follows. In Section II we discuss a theoretical model of an experimentally feasible nanoscopic setup where MBSs are entangled with a QD whose nonequilibrium states are generated by both a bias voltage and thermal voltage. Results of numerical analysis performed with high accuracy for the differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V, differential conductance I(V,VT)/V\partial I(V,V_{T})/\partial V and differential thermoelectric conductance I(V,VT)/VT\partial I(V,V_{T})/\partial V_{T} are presented in Section III where it is demonstrated that, in contrast to I(V,VT)/V\partial I(V,V_{T})/\partial V and I(V,VT)/VT\partial I(V,V_{T})/\partial V_{T}, one observes in S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V a crossover from a thermoelectric to pure thermal nonequilibrium behavior. The energy scale where this crossover takes place and a number of universal Majorana ratios involving this energy scale are also shown in this section. Finally, with Section IV we make conclusions and discuss possible outlooks.

II Theoretical model of a Majorana entangled quantum dot and the differential shot noise in thermoelectric nonequilibrium

We start with a description of a setup which, on one side, is technologically feasible Deng et al. (2016, 2018) and, on the other side, involves a basic mechanism of a Majorana entanglement which is sufficient to demonstrate a remarkable nonequilibrium behavior of the differential shot noise in presence of both bias voltages and temperature differences. To this end, let us consider a noninteracting QD,

H^QD=ϵddd.\hat{H}_{QD}=\epsilon_{d}d^{\dagger}d. (1)

The QD is nondegenerate and its energy level ϵd\epsilon_{d} is tunable by a proper gate voltage. The choice of a setup with a noninteracting QD is quite a realistic assumption to explore universal Majorana phenomena at low energies. Indeed, the spin degeneracy is assumed to be removed by an external magnetic field which excludes a possible interfering of an interaction induced Kondo universal behavior, well-known in experiment and theory of spin-degenerate QDs Ralph and Buhrman (1994); Goldhaber-Gordon et al. (1998); Meir et al. (1993); Wingreen and Meir (1994); Smirnov and Grifoni (2011a, b); Niklas et al. (2016), with the low-energy Majorana universal effects, which are of interest in this work. Numerical renormalization group calculations Ruiz-Tijerina et al. (2015) have demonstrated that interacting spin-degenerate QDs in external magnetic fields behave similar to noninteracting nondegenerate QDs exhibiting, for example, the linear conductance e2/2he^{2}/2h which results entirely from the Majorana entangled states. Thus Eq. (1) is a proper model to explore low-energy Majorana quantum transport, in particular, in nonequilibrium states resulting from bias voltages and temperature differences López et al. (2014).

Two normal noninteracting metallic contacts, denoted below as left (LL) and right (RR),

H^C=l={L,R}kϵkclkclk,\hat{H}_{C}=\sum_{l=\{L,R\}}\sum_{k}\epsilon_{k}c_{lk}^{\dagger}c_{lk}, (2)

are connected to the QD via tunneling processes,

H^QDC=l={L,R}𝒯lkclkd+H.c.\hat{H}_{QD-C}=\sum_{l=\{L,R\}}\mathcal{T}_{l}\sum_{k}c_{lk}^{\dagger}d+\text{H.c.} (3)

In Eq. (2) the continuous energy spectrum ϵk\epsilon_{k} gives rise to a density of states of the contacts νC(ϵ)\nu_{C}(\epsilon) which is in general energy dependent. However, around the Fermi energy one usually with a good accuracy assumes that its energy dependence plays no essential role for quantum transport and thus νC(ϵ)ν0/2\nu_{C}(\epsilon)\approx\nu_{0}/2. In Eq. (3) one additionally assumes that the tunneling matrix elements do not depend on the set kk of the quantum numbers used to describe the states in the contacts, 𝒯kl𝒯l\mathcal{T}_{kl}\approx\mathcal{T}_{l}. The relevant energy scales brought about by the tunneling between the QD and contacts are Γl=πν0|𝒯l|2\Gamma_{l}=\pi\nu_{0}|\mathcal{T}_{l}|^{2}. By proper gate voltages one may achieve the symmetric coupling ΓL=ΓR=Γ/2\Gamma_{L}=\Gamma_{R}=\Gamma/2 which will be assumed below.

Each contact is assumed to be in its own equilibrium state with the corresponding Fermi-Dirac distribution,

fl(ϵ)=1exp(ϵμlkBTl)+1.f_{l}(\epsilon)=\frac{1}{\exp\bigl{(}\frac{\epsilon-\mu_{l}}{k_{B}T_{l}}\bigr{)}+1}. (4)

Here the chemical potentials,

μL,R=μ0±eV/2,\mu_{L,R}=\mu_{0}\pm eV/2, (5)

are specified by the bias voltage VV such that eV<0eV<0 and the temperature of the left contact is higher than the temperature of the right contact, that is

TL=T+ΔT(hot contact),TR=T(cold contact),\begin{split}&T_{L}=T+\Delta T\quad\text{(hot contact)},\\ &T_{R}=T\quad\text{(cold contact)},\end{split} (6)

assuming T,ΔT0T,\Delta T\geqslant 0. The QD is out of equilibrium when either V0V\neq 0 or ΔT0\Delta T\neq 0.

A topological superconductor hosting two MBSs γ1,2\gamma_{1,2} at its ends,

H^TS=12iξγ2γ1,γ1,2=γ1,2,{γi,γj}=2δij,\hat{H}_{TS}=\frac{1}{2}i\xi\gamma_{2}\gamma_{1},\quad\gamma_{1,2}^{\dagger}=\gamma_{1,2},\quad\{\gamma_{i},\gamma_{j}\}=2\delta_{ij}, (7)

interacts with the QD,

H^QDTS=ηdγ1+H.c.,\hat{H}_{QD-TS}=\eta^{*}d^{\dagger}\gamma_{1}+\text{H.c.}, (8)

implementing a direct entanglement of the QD’s degrees of freedom with the Majorana mode γ1\gamma_{1} of the topological superconductor. In Eq. (7) the parameter ξ\xi is an energetic measure of how strong the two Majorana modes overlap with each other. When ξ\xi is small the MBSs are well separated whereas large values of ξ\xi model a situation where the two MBSs merge into a single Dirac fermion. In Eq. (8) the Majorana tunneling amplitude |η||\eta| specifies the strength of the Majorana entanglement.

A schematic summary of the above theoretical formulation of the setup, based on Eqs. (1)-(8), is illustrated in the inset of Fig. 1.

The Hamiltonian of the setup, H^=H^QD+H^C+H^QDC+H^TS+H^QDTS\hat{H}=\hat{H}_{QD}+\hat{H}_{C}+\hat{H}_{QD-C}+\hat{H}_{TS}+\hat{H}_{QD-TS}, allows us to formulate the problem in terms of the Keldysh field integral Altland and Simons (2010), a convenient tool to calculate various correlation functions. Other technical tools based, e.g., on quantum master equations Xu et al. (2022) may also be considered as alternative approaches to the problem. Within the Keldysh field integral formalism one may straightforwardly derive the shot noise from the Keldysh generating functional,

Z[Jlq(t)]=𝒟[ψ¯,ψ;ϕ¯,ϕ;ζ¯,ζ]eiSK[Jlq(t)],Z[J_{lq}(t)]=\int\mathcal{D}[\bar{\psi},\psi;\bar{\phi},\phi;\bar{\zeta},\zeta]e^{\frac{i}{\hbar}S_{K}[J_{lq}(t)]}, (9)

which is a field integral over the Grassmann fields of the QD (ψ¯q(t),ψq(t)\bar{\psi}_{q}(t),\psi_{q}(t)), contacts (ϕ¯lkq(t),ϕlkq(t)\bar{\phi}_{lkq}(t),\phi_{lkq}(t)) and topological superconductor (ζ¯q(t),ζq(t)\bar{\zeta}_{q}(t),\zeta_{q}(t)) whose temporal arguments are on the real axis and q=±q=\pm specifies, respectively, the forward or backward branch of the Keldysh contour. At zero source fields the Keldysh generating functional is determined by the Keldysh action SK(0)SK[Jlq(t)=0]S_{K}^{(0)}\equiv S_{K}[J_{lq}(t)=0] and is equal to unity, Z[Jlq(t)=0]=1Z[J_{lq}(t)=0]=1. The Keldysh action SK[Jlq(t)]S_{K}[J_{lq}(t)],

SK[Jlq(t)]=SQD[ψ¯,ψ]+SC[ϕ¯,ϕ]+STS[ζ¯,ζ]+SQDC[ψ¯,ψ;ϕ¯,ϕ]+SQDTS[ψ¯,ψ;ζ¯,ζ]+SO[ψ¯,ψ;ϕ¯,ϕ;Jlq(t)],\begin{split}&S_{K}[J_{lq}(t)]=S_{QD}[\bar{\psi},\psi]+S_{C}[\bar{\phi},\phi]+S_{TS}[\bar{\zeta},\zeta]\\ &+S_{QD-C}[\bar{\psi},\psi;\bar{\phi},\phi]+S_{QD-TS}[\bar{\psi},\psi;\bar{\zeta},\zeta]\\ &+S_{O}[\bar{\psi},\psi;\bar{\phi},\phi;J_{lq}(t)],\end{split} (10)

is the sum of, respectively, the actions describing the QD, contacts, topological superconductor, tunneling between the QD and contacts, tunneling between the QD and topological superconductor and the source action added to generate an observable of interest, in particular, the mean current and shot noise. The actions SQDS_{QD}, SCS_{C} and STSS_{TS} are of the standard matrix form Altland and Simons (2010) in the retarded-advanced space. The actions SQDCS_{QD-C}, SQDTSS_{QD-TS} and SOS_{O} have the following form:

SQDC=𝑑tl={L,R}k,q[𝒯lqϕ¯lkq(t)ψq(t)+G.c.],S_{QD-C}=-\!\int_{-\infty}^{\infty}\!\!\!\!\!dt\sum_{l=\{L,R\}}\sum_{k,q}[\mathcal{T}_{l}q\bar{\phi}_{lkq}(t)\psi_{q}(t)+\text{G.c.}], (11)
SQDTS=dt{ηqq[ψ¯q(t)ζq(t)+ψ¯q(t)ζ¯q(t)]+G.c.},\begin{split}S_{QD-TS}&=-\int_{-\infty}^{\infty}dt\{\eta^{*}\sum_{q}q[\bar{\psi}_{q}(t)\zeta_{q}(t)\\ &+\bar{\psi}_{q}(t)\bar{\zeta}_{q}(t)]+\text{G.c.}\},\end{split} (12)
SO=𝑑tl={L,R}qJlq(t)Ilq(t),S_{O}=-\int_{-\infty}^{\infty}dt\sum_{l=\{L,R\}}\sum_{q}J_{lq}(t)I_{lq}(t), (13)

where G.c. denotes the Grassmann conjugated terms and Ilq(t)I_{lq}(t) is the current operator in the Grassmann representation,

Ilq(t)=iek(𝒯lϕ¯lkq(t)ψq(t)G.c.).I_{lq}(t)=\frac{ie}{\hbar}\sum_{k}\bigl{(}\mathcal{T}_{l}\bar{\phi}_{lkq}(t)\psi_{q}(t)-\text{G.c.}\bigr{)}. (14)

The form of the source action in Eq. (13) implies that one derives the mean current and current-current correlations via proper differentiations,

Ilq(t)0=iδZ[Jlq(t)]δJlq(t)|Jlq(t)=0,\langle I_{lq}(t)\rangle_{0}=i\hbar\frac{\delta Z[J_{lq}(t)]}{\delta J_{lq}(t)}\biggl{|}_{J_{lq}(t)=0}, (15)
Ilq(t)Ilq(t)0=(i)2δ2Z[Jlq(t)]δJlq(t)δJlq(t)|Jlq(t)=0,\langle I_{lq}(t)I_{l^{\prime}q^{\prime}}(t^{\prime})\rangle_{0}=(i\hbar)^{2}\frac{\delta^{2}Z[J_{lq}(t)]}{\delta J_{lq}(t)\delta J_{l^{\prime}q^{\prime}}(t^{\prime})}\biggl{|}_{J_{lq}(t)=0}, (16)

where

iIliqi(ti)0𝒟[ψ¯,ψ;ϕ¯,ϕ;ζ¯,ζ]eiSK(0)iIliqi(ti).\begin{split}&\langle\prod_{i}I_{l_{i}q_{i}}(t_{i})\rangle_{0}\\ &\equiv\int\mathcal{D}[\bar{\psi},\psi;\bar{\phi},\phi;\bar{\zeta},\zeta]e^{\frac{i}{\hbar}S_{K}^{(0)}}\prod_{i}I_{l_{i}q_{i}}(t_{i}).\end{split} (17)

Choosing the left contact as the one where measurements of the mean current

I(V,VT)=ILq(t)0,I(V,V_{T})=\langle I_{Lq}(t)\rangle_{0}, (18)

and correlations

S>(t,t;V,VT)=δIL(t)δIL+(t)0S^{>}(t,t^{\prime};V,V_{T})=\langle\delta I_{L-}(t)\delta I_{L+}(t^{\prime})\rangle_{0} (19)

of the current fluctuations

δILq(t)=ILq(t)I(V,VT)\delta I_{Lq}(t)=I_{Lq}(t)-I(V,V_{T}) (20)

are performed, one obtains the shot noise S>(V,VT)S^{>}(V,V_{T}) in the left contact as the zero frequency Fourier transform of S>(t,t;V,VT)=S>(tt;V,VT)S^{>}(t,t^{\prime};V,V_{T})=S^{>}(t-t^{\prime};V,V_{T}),

S>(ω;V,VT)=𝑑teiωtS>(t;V,VT),S>(V,VT)=S>(ω=0;V,VT).\begin{split}&S^{>}(\omega;V,V_{T})=\int_{-\infty}^{\infty}dt\,e^{i\omega t}S^{>}(t;V,V_{T}),\\ &S^{>}(V,V_{T})=S^{>}(\omega=0;V,V_{T}).\end{split} (21)

As it has already been mentioned in Section I, majority of quantum transport experiments deal with mean currents, Eq. (18), specifically, with their differential characteristics such as the differential conductance or, less often, differential thermoelectric conductance, I(V,VT)/V\partial I(V,V_{T})/\partial V or I(V,VT)/VT\partial I(V,V_{T})/\partial V_{T}, respectively. These quantities have universal units of e2/he^{2}/h and thus provide direct access to universal properties of MBSs. Likewise, experiments dealing with shot noises, Eq. (21), and their derivatives, may access universal fluctuation behavior of Majorana entangled states via, for example, the differential shot noise, S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V, having universal units of e3/he^{3}/h. Whereas the Majorana universality of S>/V\partial S^{>}/\partial V is still an experimental challenge for Majorana entangled setups, the differential shot noise has already been successfully measured to probe other types of fluctuation universality, for example the universality of the Kondo noise in quantum dots Basset et al. (2012). Although experiments on current fluctuations are more complicated than those measuring mean currents, results of such noise measurements provide a much more detailed microscopic structure of various nanoscopic setups.

Below we obtain the differential conductance I(V,VT)/V\partial I(V,V_{T})/\partial V, differential thermoelectric conductance I(V,VT)/VT\partial I(V,V_{T})/\partial V_{T} and differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V by means of numerical calculations based on finite differences used to approximate the corresponding derivatives. Here we would like to emphasize that although the above theoretical model is noninteracting, numerical calculations of S>(V,VT)S^{>}(V,V_{T}) and I(V,VT)I(V,V_{T}) are nevertheless necessary. The point is that after obtaining closed analytic expressions for S>(V,VT)S^{>}(V,V_{T}) and I(V,VT)I(V,V_{T}), which is possible because the Keldysh field integral is quadratic in the fermionic fields, it still remains to perform integrals in the energy domain (see the Appendix in Ref. Smirnov (2018)) in these analytic expressions. These integrals are hard to calculate analytically, especially, at finite temperature differences (characterized by finite thermal voltages VTV_{T}), that is when the Fermi-Dirac distributions in Eq. (4) cannot be approximated by step-like functions. In general, calculations of the differential shot noise are more time consuming than those which would be necessary to get just the shot noise. Whereas a certain numerical accuracy may be sufficient to get curves looking smooth enough for S>(V,VT)S^{>}(V,V_{T}), using the same numerical data to calculate S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V may result in numerical errors leading to a chaotic dataset. Thus to obtain S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V with an accuracy that allows to identify dependence on various parameters as well as corresponding coefficients, the calculation of S>(V,VT)S^{>}(V,V_{T}) should be done with a proper precision. Clearly, a higher degree of numerical accuracy leads to a notable increase of the computational time but still makes it possible to perform for the setup described in this section a detailed analysis of S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V, in particular, its universal Majorana thermoelectric crossover discussed thoroughly in the next section.

III Numerical analysis of the differential shot noise and its thermoelectric crossover

In this section we present numerical results for the differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V and demonstrate that, in contrast to the differential conductance I(V,VT)/V\partial I(V,V_{T})/\partial V and differential thermoelectric conductance I(V,VT)/VT\partial I(V,V_{T})/\partial V_{T}, it exhibits a crossover from one type of nonequilibrium behavior to a qualitatively different one. This crossover occurs in the regime

ΓeVT|eV|ξ,\Gamma\gg eV_{T}\gg|eV|\gg\xi, (22)

and in the most part of this section we focus on quantum transport in this regime except for the last part where we show that the crossover disappears for large values of the Majorana overlap energy ξ\xi.

In Fig. 1 we show numerical results obtained for the differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V as a function of the thermal voltage VTV_{T} for different values of the bias voltage VV. As can be seen, at a certain value of the thermal voltage VT=VT,minV_{T}=V_{T,\text{min}} each of the three curves has a characteristic minimum (shown by the corresponding circle) which represents a crossover from a thermoelectric to pure thermal nonequilibrium behavior. Indeed, the decreasing, or thermoelectric, branch depends on both the electric driving VV and thermal driving VTV_{T} with the asymptotic behavior shown by the inclined dashed line whereas the increasing, or pure thermal, branch depends only on the thermal driving VTV_{T} and does not

Refer to caption
Figure 1: Differential shot noise S>/V\partial S^{>}/\partial V as a function of the thermal voltage VTV_{T} for three different values of the bias voltage: |eV|/Γ=1012|eV|/\Gamma=10^{-12} (black), |eV|/Γ=1011|eV|/\Gamma=10^{-11} (red), |eV|/Γ=1010|eV|/\Gamma=10^{-10} (blue). The other parameters have the following values: ϵd/Γ=101\epsilon_{d}/\Gamma=10^{-1}, kBT/Γ=1012k_{B}T/\Gamma=10^{-12}, |η|/Γ=1|\eta|/\Gamma=1, ξ/Γ=1014\xi/\Gamma=10^{-14}. The Majorana device presented above the curves illustrates schematically the physical setup, described in the main text, Eqs. (1)-(8), assuming eV<0eV<0 and ΔT>0\Delta T>0.

depend on the electric driving VV. Our numerical analysis shows that the asymptotics of the thermoelectric and pure thermal nonequilibrium branches are, respectively, given by the following analytic expressions:

S>(V,VT)V=e3h18|eV|eVT,for VT:|eV|eVTeVT,min,\begin{split}&\frac{\partial S^{>}(V,V_{T})}{\partial V}=\frac{e^{3}}{h}\frac{1}{8}\frac{|eV|}{eV_{T}},\\ &\text{for $V_{T}$:}\quad|eV|\ll eV_{T}\ll eV_{T,\text{min}},\end{split} (23)

and

S>(V,VT)V=e3h1ln(2)4ϵd(eVT)η2,for VT:eVT,mineVTΓ,\begin{split}&\frac{\partial S^{>}(V,V_{T})}{\partial V}=\frac{e^{3}}{h}\frac{1-\ln(2)}{4}\frac{\epsilon_{d}(eV_{T})}{\eta^{2}},\\ &\text{for $V_{T}$:}\quad eV_{T,\text{min}}\ll eV_{T}\ll\Gamma,\end{split} (24)

both of which we are able to reproduce with any desired numerical accuracy. Similarly, in the whole range of the thermal voltage VTV_{T}, restricted by the regime specified in Eq. (22), our numerical calculations show that the analytic expression for the differential shot noise is given by the sum of Eqs. (23) and (24),

S>(V,VT)V=e3h[18|eV|eVT+1ln(2)4ϵd(eVT)|η|2],for VT:|eV|eVTΓ.\begin{split}&\frac{\partial S^{>}(V,V_{T})}{\partial V}=\frac{e^{3}}{h}\biggl{[}\frac{1}{8}\frac{|eV|}{eV_{T}}+\frac{1-\ln(2)}{4}\frac{\epsilon_{d}(eV_{T})}{|\eta|^{2}}\biggr{]},\\ &\text{for $V_{T}$:}\quad|eV|\ll eV_{T}\ll\Gamma.\end{split} (25)

An analytic derivation of Eq. (25) is a complicated task which we would like to address in a separate paper. We note that a proper analytic analysis may provide corrections to Eq. (25) and show under which conditions these corrections start to play an essential role. Also using the Sommerfeld expansion Ashkroft and Mermin (1976), one may analytically derive the differential shot noise in the complementary regime where eVT|eV|eV_{T}\ll|eV|. However, within the specified regime, Eq. (22), the analytic expression in Eq. (25) has been confirmed with any desired numerical precision. This means that the stronger the inequalities in Eq. (22) are fulfilled, the more digits after the decimal point are reproduced numerically for any given value obtained analytically from Eq. (25).

From Eq. (25) we find that eVT,mineV_{T,\text{min}} depends on the bias voltage VV, gate voltage ϵd\epsilon_{d} and Majorana tunneling amplitude |η||\eta|,

eVT,min={12[1ln(2)]|eV||η|2ϵd}12.eV_{T,\text{min}}=\biggl{\{}\frac{1}{2[1-\ln(2)]}\frac{|eV||\eta|^{2}}{\epsilon_{d}}\biggr{\}}^{\frac{1}{2}}. (26)

The differential shot noise at VT=VT,minV_{T}=V_{T,\text{min}} is obtained from Eqs. (25) and (26) which lead to the following result:

S>(V,VT)V|VT=VT,min=e3h[1ln(2)8ϵd|eV||η|2]12.\frac{\partial S^{>}(V,V_{T})}{\partial V}\bigg{|}_{V_{T}=V_{T,\text{min}}}=\frac{e^{3}}{h}\biggl{[}\frac{1-\ln(2)}{8}\frac{\epsilon_{d}|eV|}{|\eta|^{2}}\biggr{]}^{\frac{1}{2}}. (27)
Refer to caption
Figure 2: Differential shot noise S>/V\partial S^{>}/\partial V as a function of the thermal voltage VTV_{T} for three different values of the gate voltage: ϵd/Γ=102\epsilon_{d}/\Gamma=10^{-2} (black), ϵd/Γ=101\epsilon_{d}/\Gamma=10^{-1} (red), ϵd/Γ=1\epsilon_{d}/\Gamma=1 (blue). The other parameters have the following values: |eV|/Γ=1010|eV|/\Gamma=10^{-10}, kBT/Γ=1012k_{B}T/\Gamma=10^{-12}, |η|/Γ=1|\eta|/\Gamma=1, ξ/Γ=1014\xi/\Gamma=10^{-14}.

Note, that the thermoelectric branch, Eq. (23), is universal because it does not depend on the gate voltage ϵd\epsilon_{d} and depends only on the ratio between the electric and thermal driving, VV and VTV_{T}, respectively.

According to Eqs. (26) and (27) both the location of the crossover, VT,minV_{T,\text{min}}, and the value of S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V at VT=VT,minV_{T}=V_{T,\text{min}} depend on VV. This suggests that the crossover results from an interplay between the electric and thermal driving. Moreover, the dependence of the crossover on ϵd\epsilon_{d} in Eqs. (26) and (27) (see also Fig. 2) is another indication that the crossover emerges from a competition of the two flows excited, respectively, by the electric and thermal driving. Obviously, due to the particle-hole symmetry, the current cannot be excited exclusively by the thermal driving VTV_{T} when ϵd=0\epsilon_{d}=0. The pure thermal driving VTV_{T} induces a finite current only when ϵd0\epsilon_{d}\neq 0.

The straight solid line in Fig. 1 shows both the locations VT,minV_{T,\text{min}} of the crossovers and the values of S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V at VT=VT,minV_{T}=V_{T,\text{min}} parameterized by the bias voltage VV according to Eqs. (26) and (27). The figure clearly shows that the crossovers (highlighted by the circles), obtained from the numerical calculations, reside perfectly on the analytic straight solid line. As expected, at low energies, eVT|eV|eV_{T}\ll|eV| (that is outside the regime in Eq. (22)), the differential shot noise reaches its universal asymptotic unitary value e3/4he^{3}/4h shown by the horizontal dashed line (see also Refs. Liu et al. (2015a); Smirnov (2017)).

The numerical results presented in Fig. 2 show the differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V as a function of the thermal voltage VTV_{T} for different values of the gate voltage ϵd\epsilon_{d}. As in Fig. 1, each of the three curves possesses a crossover (shown by the corresponding circle) from a thermoelectric to pure thermal nonequilibrium behavior. As mentioned above, on the left side of the crossover the decreasing, or thermoelectric, branch is universal: it depends on both the electric driving VV and thermal driving VTV_{T} via their ratio and Fig. 2 explicitly demonstrates that it does not depend on the gate voltage ϵd\epsilon_{d}. On the right side of the crossover the increasing, or pure thermal, branch, which is driven only by the thermal voltage VTV_{T}, is obviously not universal. Indeed, the figure clearly shows that this branch depends on the gate voltage ϵd\epsilon_{d}. The inclined dashed line shows the asymptotic behavior of the pure thermal nonequilibrium branch, in particular, its dependence on the gate voltage ϵd\epsilon_{d}. Note, that in fact this pure thermal nonequilibrium branch is not universal because of two reasons. The first reason, the dependence on ϵd\epsilon_{d}, was already mentioned above. The second reason is that this branch additionally depends on the Majorana tunneling amplitude |η||\eta| as it is also shown in its asymptotic behavior.

In addition to the dependence on VV, both the location of the crossover, VT,minV_{T,\text{min}}, and the value of S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V at VT=VT,minV_{T}=V_{T,\text{min}} depend on ϵd\epsilon_{d}. The straight solid line shows both the locations VT,minV_{T,\text{min}} of the crossovers and the values of S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V at VT=VT,minV_{T}=V_{T,\text{min}} parameterized by the gate voltage ϵd\epsilon_{d} according to Eqs. (26) and (27). As in the case of the parametric dependence on VV, one clearly sees that the numerically obtained crossovers, marked by the circles, also reside perfectly on the analytic straight line resulting from the parametric dependence on ϵd\epsilon_{d}. Here the universality of the differential shot noise at low energies, eVT|eV|eV_{T}\ll|eV|, is explicitly visible: the asymptotic low-energy behavior is obviously independent of ϵd\epsilon_{d} and is characterized by the unitary value e3/4he^{3}/4h shown by the horizontal dashed line.

The nonequilibrium Majorana crossover in the differential shot noise has a number of universal properties which may quantitatively be expressed via a number of ratios taking universal values. For example, according to Eqs. (26) and (27) the ratio

R1eVT,min|eV|S>(V,VT)V|VT=VT,minR_{1}\equiv\frac{eV_{T,\text{min}}}{|eV|}\frac{\partial S^{>}(V,V_{T})}{\partial V}\bigg{|}_{V_{T}=V_{T,\text{min}}} (28)

is independent of VV and takes the universal value

R1(M)=e34hR_{1}^{(M)}=\frac{e^{3}}{4h} (29)

for bias voltages satisfying Eq. (22).

Refer to caption
Figure 3: Dimensionless ratio R3R_{3}, defined in Eq. (33), which relates the thermoelectric (VT1V_{T_{1}} is on the left side of the crossover) and pure thermal (VT2V_{T_{2}} is on the right side of the crossover) branches of the differential shot noise S>/V\partial S^{>}/\partial V via its value at the crossover VT=VT,minV_{T}=V_{T,\text{min}}. Upper (Lower) panel: R3R_{3} as a function of VT1V_{T_{1}} (VT2V_{T_{2}}) at a fixed value of VT2V_{T_{2}} (VT1V_{T_{1}}). Here the values of the parameters are as follows: ϵd/Γ=101\epsilon_{d}/\Gamma=10^{-1}, |eV|/Γ=1010|eV|/\Gamma=10^{-10}, kBT/Γ=1012k_{B}T/\Gamma=10^{-12}, |η|/Γ=1|\eta|/\Gamma=1, ξ/Γ=1014\xi/\Gamma=10^{-14}, and eVT1/Γ=107eV_{T_{1}}/\Gamma=10^{-7} (lower panel), eVT2/Γ=103eV_{T_{2}}/\Gamma=10^{-3} (upper panel). For the above values of the parameters one gets from Eq. (26) that eVT,min/Γ4105eV_{T,\text{min}}/\Gamma\approx 4\cdot 10^{-5} (see the vertical dashed line).

For the mean current at low bias voltages in Ref. Smirnov (2020) it has been found that

I(V,VT)VT=e2hπ212ϵd(eVT)|η|2.\frac{\partial I(V,V_{T})}{\partial V_{T}}=\frac{e^{2}}{h}\frac{\pi^{2}}{12}\frac{\epsilon_{d}(eV_{T})}{|\eta|^{2}}. (30)

Using Eqs. (26), (27) and (30) one finds that the dimensionless ratio

R2S>(V,VT)V|VT=VT,mineVT,min2I(V,VT)VT2R_{2}\equiv\frac{\frac{\partial S^{>}(V,V_{T})}{\partial V}\bigg{|}_{V_{T}=V_{T,\text{min}}}}{eV_{T,\text{min}}\frac{\partial^{2}I(V,V_{T})}{\partial V_{T}^{2}}} (31)

becomes universal,

R2(M)=[1ln(2)]6π2R_{2}^{(M)}=[1-\ln(2)]\frac{6}{\pi^{2}} (32)

under the conditions in Eq. (22).

If one takes two values of the thermal voltage, VT1V_{T_{1}} and VT2V_{T_{2}}, such that VT1V_{T_{1}} belongs to the thermoelectric branch and VT2V_{T_{2}} belongs to the pure thermal branch, then from Eqs. (23), (24) and (27) it follows that the dimensionless ratio

R3VT1VT2S>(V,VT)V|VT=VT1S>(V,VT)V|VT=VT2(S>(V,VT)V|VT=VT,min)2R_{3}\equiv\frac{V_{T_{1}}}{V_{T_{2}}}\frac{\frac{\partial S^{>}(V,V_{T})}{\partial V}\bigg{|}_{V_{T}=V_{T_{1}}}\frac{\partial S^{>}(V,V_{T})}{\partial V}\bigg{|}_{V_{T}=V_{T_{2}}}}{\biggl{(}\frac{\partial S^{>}(V,V_{T})}{\partial V}\bigg{|}_{V_{T}=V_{T,\text{min}}}\biggr{)}^{2}} (33)

is universal,

R3(M)=14,R_{3}^{(M)}=\frac{1}{4}, (34)

that is independent of VT1V_{T_{1}} (VT2V_{T_{2}}) at fixed VT2V_{T_{2}} (VT1V_{T_{1}}). This is demonstrated in Fig. 3 where the fixed values of VT1V_{T_{1}} (lower panel) and VT2V_{T_{2}} (upper panel) are respectively chosen such that |eV|eVT1eVT,min|eV|\ll eV_{T_{1}}\ll eV_{T,\text{min}} and eVT,mineVT2ΓeV_{T,\text{min}}\ll eV_{T_{2}}\ll\Gamma. As one can see in the upper panel

Refer to caption
Figure 4: Differential shot noise S>/V\partial S^{>}/\partial V (upper panel), differential thermoelectric conductance I/VT\partial I/\partial V_{T} (middle panel) and their ratio R4R_{4} (lower panel) as functions of the thermal voltage VTV_{T}. Here the results have been obtained for the following values of the parameters: ϵd/Γ=101\epsilon_{d}/\Gamma=10^{-1}, |eV|/Γ=1012|eV|/\Gamma=10^{-12}, kBT/Γ=1012k_{B}T/\Gamma=10^{-12}, |η|/Γ=1|\eta|/\Gamma=1, ξ/Γ=1014\xi/\Gamma=10^{-14}. According to Eq. (26), these values of the parameters give eVT,min/Γ4106eV_{T,\text{min}}/\Gamma\approx 4\cdot 10^{-6}.

of Fig. 3, on the left (thermoelectric) side of the crossover the numerically obtained dimensionless ratio R3R_{3} has a perfect plateau with R3=1/4R_{3}=1/4 (shown by the horizontal dashed line) as expected in the range |eV|eVT1eVT,min|eV|\ll eV_{T_{1}}\ll eV_{T,\text{min}} for any fixed value of VT2V_{T_{2}} taken from the range eVT,mineVT2ΓeV_{T,\text{min}}\ll eV_{T_{2}}\ll\Gamma. As demonstrated in the lower panel of Fig. 3, also on the right (pure thermal) side of the crossover the numerical curve develops a clear plateau on which R3=1/4R_{3}=1/4 (horizontal dashed line) in the range eVT,mineVT2ΓeV_{T,\text{min}}\ll eV_{T_{2}}\ll\Gamma for any fixed value of VT1V_{T_{1}} taken from the range |eV|eVT1eVT,min|eV|\ll eV_{T_{1}}\ll eV_{T,\text{min}}. In contrast, when VT1V_{T_{1}} moves away from the thermoelectric branch (left side of the crossover), that is when eVT1|eV|eV_{T_{1}}\lesssim|eV| or VT1VT,minV_{T_{1}}\gtrsim V_{T,\text{min}}, the ratio R3R_{3} must deviate from the value 1/41/4. The numerical results (solid curve) in the upper panel show that indeed deviations from the plateau R3=1/4R_{3}=1/4 occur when eVT1|eV|eV_{T_{1}}\lesssim|eV| or VT1VT,minV_{T_{1}}\gtrsim V_{T,\text{min}}. Similarly, when VT2V_{T_{2}} is not located on the pure thermal branch (right side of the crossover), that is when VT2VT,minV_{T_{2}}\lesssim V_{T,\text{min}} or eVT2ΓeV_{T_{2}}\gtrsim\Gamma, the ratio R3R_{3} must also shift away from the plateau on which it reaches the value 1/41/4. As anticipated, the solid curve resulting from numerical calculations demonstrates that its plateau-like behavior in the lower panel breaks in the domains where VT2VT,minV_{T_{2}}\lesssim V_{T,\text{min}} or eVT2ΓeV_{T_{2}}\gtrsim\Gamma.

Now let us consider only the pure thermal nonequilibrium branch of the differential shot noise. From Eqs. (24) and (30) one finds that the ratio

R4S>(V,VT)VI(V,VT)VTR_{4}\equiv\frac{\frac{\partial S^{>}(V,V_{T})}{\partial V}}{\frac{\partial I(V,V_{T})}{\partial V_{T}}} (35)

becomes universal,

R4(M)=e3[1ln(2)]π2,R_{4}^{(M)}=e\frac{3[1-\ln(2)]}{\pi^{2}}, (36)

in the range of the pure thermal nonequilibrium branch. In Fig. 4 we present numerical results for the ratio R4R_{4}. From the upper and middle panels one immediately sees the qualitative difference between the behavior of respectively the differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V and differential thermoelectric conductance I(V,VT)/VT\partial I(V,V_{T})/\partial V_{T}. Indeed, whereas the differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V exhibits in its minimum a crossover separating a thermoelectric branch (left side of the crossover) from a pure thermal branch (right side of the crossover), the differential thermoelectric conductance I(V,VT)/VT\partial I(V,V_{T})/\partial V_{T} does not demonstrate any crossover and has only one, pure thermal, nonequilibrium branch. As has been discussed above, in contrast to the universal thermoelectric branch of S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V, its pure thermal branch is not universal because of its dependence on the gate voltage ϵd\epsilon_{d} and the Majorana tunneling amplitude |η||\eta| as can be seen in Eq. (24). Comparing the two non-universal pure thermal branches of the differential shot noise and differential thermoelectric conductance, Eqs. (24) and (30), respectively, one sees that both of them depend linearly on the thermal voltage VTV_{T} and have identical parametric dependence on the gate voltage ϵd\epsilon_{d} and Majorana tunneling amplitude |η||\eta|. Thus, although the pure thermal branches of the differential shot noise and differential thermoelectric conductance are not universal when considered separately from each other, their ratio R4R_{4} in Eq. (35), must be universal, that is it must be independent of the thermal voltage VTV_{T}, gate voltage ϵd\epsilon_{d} and Majorana tunneling amplitude |η||\eta|. Moreover, according to Eq. (36), one expects that in the range of VTV_{T} corresponding to the pure thermal branch of S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V the ratio R4R_{4} must be equal to 3[1ln(2)]/π23[1-\ln(2)]/\pi^{2} in the universal units of the elementary charge ee. The numerical results presented in the lower panel confirm this expectation: on the right (pure thermal) side of the crossover of S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V the ratio R4R_{4} exhibits a plateau-like behavior in the range eVT,mineVTΓeV_{T,\text{min}}\ll eV_{T}\ll\Gamma with R4=3e[1ln(2)]/π2R_{4}=3e[1-\ln(2)]/\pi^{2} on the plateau. At this point we would also like to note that, similarly to the differential thermoelectric conductance I(V,VT)/VT\partial I(V,V_{T})/\partial V_{T}, the differential conductance I(V,VT)/V\partial I(V,V_{T})/\partial V does not exhibit any crossover. Our numerical calculations show that it remains almost independent of VTV_{T} and retains its Majorana fractional value I(V,VT)/V=e2/2h\partial I(V,V_{T})/\partial V=e^{2}/2h up to eVTΓeV_{T}\sim\Gamma where it starts to decrease and becomes strongly suppressed, i.e. I(V,VT)/Ve2/2h\partial I(V,V_{T})/\partial V\ll e^{2}/2h, for eVTΓeV_{T}\geqslant\Gamma.

Refer to caption
Figure 5: Ratio R4R_{4} defined in Eq. (35) (duplicated also below the solid curve in the figure) as a function of the Majorana overlap energy ξ\xi. Here we put ϵd/Γ=101\epsilon_{d}/\Gamma=10^{-1}, |eV|/Γ=1012|eV|/\Gamma=10^{-12}, eVT/Γ=103eV_{T}/\Gamma=10^{-3}, kBT/Γ=1012k_{B}T/\Gamma=10^{-12}, |η|/Γ=1|\eta|/\Gamma=1.

To see what happens when the two MBSs are not well separated, we have performed numerical calculations for larger values of the Majorana overlap energy ξ\xi. Our results show that the above discussed crossover and universal values of the ratios R1,2,3,4R_{1,2,3,4} disappear. For example, Fig. 5 shows the ratio R4R_{4} as a function of ξ\xi. For well separated MBSs the values of the parameters are chosen to drive the system into the regime where it stays within the plateau shown in the lower panel of Fig. 4, that is when the ratio between the pure thermal branch of the differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V and differential thermoelectric conductance I(V,VT)/VT\partial I(V,V_{T})/\partial V_{T} takes its universal value, R4(M)=3e[1ln(2)]/π2R_{4}^{(M)}=3e[1-\ln(2)]/\pi^{2}. As Fig. 5 clearly demonstrates, for small values of the Majorana overlap energy ξ\xi the ratio between the pure thermal nonequilibrium branches of S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V and I(V,VT)/VT\partial I(V,V_{T})/\partial V_{T} is equal to its universal Majorana value R4(M)R_{4}^{(M)}. However, when ξ\xi grows, the two MBSs significantly merge into a single Dirac fermion and cannot be probed separately anymore. In this situation the universal nonequilibrium Majorana behavior breaks. As a consequence, the ratio R4R_{4} significantly deviates from its universal Majorana plateau R4(M)R_{4}^{(M)}. Moreover, for large values of the Majorana overlap energy ξ\xi both the thermoelectric and pure thermal nonequilibrium branches of the differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V are destroyed and the notion of the crossover discussed above loses its sense as one would expect for a phenomenon having a Majorana nature.

Refer to caption
Figure 6: Differential shot noise S>/V\partial S^{>}/\partial V in the absence of MBSs is shown as a function of the thermal voltage VTV_{T} for three different values of the gate voltage: ϵd/Γ=101\epsilon_{d}/\Gamma=10^{-1} (black), ϵd/Γ=0.74×101\epsilon_{d}/\Gamma=0.74\times 10^{-1} (red), ϵd/Γ=1.18×101\epsilon_{d}/\Gamma=1.18\times 10^{-1} (blue). The other parameters have the following values: |eV|/Γ=1010|eV|/\Gamma=10^{-10}, kBT/Γ=1012k_{B}T/\Gamma=10^{-12}, |η|=ξ=0|\eta|=\xi=0.

Finally, to demonstrate that the universal Majorana thermoelectric crossover in the differential shot noise represents a specific behavior strikingly distinct from what is observed in conventional systems without coupling to MBSs, we have computed S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V in the absence of the topological superconductor. The Hamiltonian of the setup without the topological superconductor is obtained from our Hamiltonian if one sets |η|=ξ=0|\eta|=\xi=0. The results obtained for this setup are shown in Fig. 6. As can be seen, in the absence of MBSs the dependence of the differential shot noise on VTV_{T} is qualitatively different from the Majorana induced behavior in two respects.

First, as Fig. 6 shows, the differential shot noise becomes a monotonic function of VTV_{T} for |eV|eVTΓ|eV|\ll eV_{T}\ll\Gamma. Its monotonically decreasing character makes S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V negative at some point (that is why we avoid using the logarithmic scale for the yy-axis). In this sense the differential shot noise is not qualitatively singled out because the differential electric and thermoelectric conductances are also monotonically decreasing functions of VTV_{T} for |eV|eVTΓ|eV|\ll eV_{T}\ll\Gamma in the absence of MBSs. Specifically, for |eV|eVTΓ|eV|\ll eV_{T}\ll\Gamma the differential electric conductance is almost independent of VTV_{T} (and is equal to a value which depends on the gate voltage ϵd\epsilon_{d}) up to eVTΓeV_{T}\sim\Gamma where it starts to quickly decrease, whereas the differential thermoelectric conductance is, unlike the Majorana case, always negative and decreases linearly with VTV_{T} (that is its absolute value grows). In contrast, when MBSs are present, the differential shot noise is qualitatively singled out by its nonmonotonic behavior characterized by a minimum, specifying the thermoelectric crossover, as opposed to the differential electric and thermoelectric conductances having monotonic behavior exhibiting no minima or maxima for |eV|eVTΓ|eV|\ll eV_{T}\ll\Gamma.

Second, the three curves in Fig. 6, corresponding to three different values of the gate voltage ϵd\epsilon_{d}, demonstrate that in the whole range of the thermal voltage VTV_{T} the differential shot noise is not universal, that is in the absence of MBSs S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V depends on ϵd\epsilon_{d} for any value of VTV_{T}. This nonuniversal behavior of S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V is qualitatively different from what has been demonstrated in Fig. 2 where coupling to MBSs makes the thermoelectric branch (left side of the crossover) of S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V universal that is independent of ϵd\epsilon_{d}. In contrast, without coupling to MBSs even small variations of the gate voltage ϵd\epsilon_{d} produce large changes in the differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V in the whole range of the thermal voltage VTV_{T} as it is clearly seen in Fig. 6.

IV Conclusion

We have explored the differential shot noise S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V in a Majorana entangled QD device driven out of equilibrium by both the bias voltage VV and thermal voltage VTV_{T}. The numerical analysis of high precision has been used to reveal the existence of a crossover in the behavior of S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V as a function of VTV_{T} and identify its analytic form. In particular, it has been shown that this crossover results from an interplay between the two types of nonequilibrium fluctuations induced by respectively VV and VTV_{T} and separates thermoelectric nonequilibrium behavior of the differential shot noise from its pure thermal nonequilibrium behavior. The energy scale of the crossover as well as its nonequilibrium fluctuation nature invisible for mean current probes have been identified and the crossover dependences on the gate voltage, bias voltage and Majorana tunneling amplitude have been explicitly shown. Additionally, various universal Majorana ratios R1,2,3,4R_{1,2,3,4} involving the energy scale of the crossover have been provided for a future experimental access to universal fluctuation behavior of Majorana entangled states within either pure noise measurements, ratios R1,3R_{1,3}, or in combination with measurements of mean currents, ratios R2,4R_{2,4}. It has been found that the crossover is destroyed when the two MBSs of the topological superconductor start to overlap and merge into a single Dirac fermion. This results in a disappearance of the universal Majorana plateaus in the ratios R1,2,3,4R_{1,2,3,4} as has been exemplified via numerical calculations for R4R_{4}. Finally, we have demonstrated that whereas for Majorana entangled states the differential shot noise has a nonmonotonic behavior characterized by a minimum with universal properties, in conventional systems without coupling to MBSs the differential shot noise is a monotonic and nonuniversal function in the whole range of VTV_{T}. Thus, in contrast to Majorana entangled states, in setups without MBSs the monotonic differential shot noise is not qualitatively different from the differential electric and thermoelectric conductances which are also monotonic functions of VTV_{T} in the absence of MBSs.

For an experimental verification of the theoretical results presented in this work one might consider the devices studied in Refs. Deng et al. (2016, 2018). These devices are based on InAs nanowires covered by an Al layer grown by molecular beam epitaxy. The Al layer is the superconductor which is used to induce a topological superconducting state in the InAs nanowire whose ends are assumed to host MBSs γ1,2\gamma_{1,2}. To couple γ1\gamma_{1} to a QD the Al layer is etched on one end of the InAs nanowire. This bare part of InAs is the place where one forms a QD coupled to the Majorana state γ1\gamma_{1} with the coupling strength |η||\eta|. As explained in Ref. Deng et al. (2018), the occupancy (or the energy level ϵd\epsilon_{d} in our context) of the QD is tuned by proper gate voltages. In addition to the setup in Refs. Deng et al. (2016, 2018), one may also form two independent normal metallic contacts coupled to the QD with the coupling strength Γ\Gamma. These two independent normal metallic contacts may, in general, have different chemical potentials μL,R\mu_{L,R} and different temperatures TL,RT_{L,R}. To measure the differential shot noise one could try to adapt, for example, the technology from Ref. Basset et al. (2012) based on coupling of a setup to a quantum noise detector. Here for the quantum noise detector one also uses Al as a superconductor and thus it might be compatible with the above technology Deng et al. (2016, 2018) for topological superconductivity. The setup in Ref. Basset et al. (2012) is a carbon nanotube. It may be replaced with the InAs nanowire from Refs. Deng et al. (2016, 2018). One possible problem here is that measurements in Ref. Basset et al. (2012) assume finite frequencies. Nevertheless, one may still measure the differential shot noise if the resonant frequencies in Ref. Basset et al. (2012) are made smaller than all the relevant energy scales of our setup. This might be achieved, for example, by increasing the length of the transmission lines in Ref. Basset et al. (2012) or by other relevant techniques.

Among possible outlooks we would like to mention setups with Aharonov-Bohm fluxes Zou et al. (2023b) or setups where both MBSs are directly entangled with a QD whose nonequilibrium states are governed by bias voltages and temperature differences. Majorana interference effects in such setups will emerge through the Majorana tunneling phases forming a complex interplay with the two competing flows induced by respectively VV and VTV_{T} and the fate of the Majorana crossover in S>(V,VT)/V\partial S^{>}(V,V_{T})/\partial V in this situation is an interesting and important problem. The results presented in this work have been obtained assuming that interactions between the Majorana entangled setup and its environment are sufficiently weak. Under certain circumstances, however, such interactions may have a significant impact on the shot noise via corresponding inelastic processes Krainov et al. (2022) and thus represent a challenge for future models where MBSs are coupled to an external environment. Another possibility is to study the differential shot noise in nonequilibrium setups with poor man’s MBSs Tsintzis et al. (2022); Dvir et al. (2023) which may arise inside QDs when one fine-tunes parameters of such setups to locate their states as close as possible to their sweet spots.

Acknowledgments

The author thanks Reinhold Egger for valuable comments.

References

  • Yu. Kitaev (2001) A. Yu. Kitaev, “Unpaired Majorana fermions in quantum wires,” Phys.-Usp. 44, 131 (2001).
  • Alicea (2012) J. Alicea, “New directions in the pursuit of Majorana fermions in solid state systems,” Rep. Prog. Phys. 75, 076501 (2012).
  • Leijnse and Flensberg (2012) M. Leijnse and K. Flensberg, “Introduction to topological superconductivity and Majorana fermions,” Semicond. Sci. Technol. 27, 124003 (2012).
  • Sato and Fujimoto (2016) M. Sato and S. Fujimoto, “Majorana fermions and topology in superconductors,” J. Phys. Soc. Japan 85, 072001 (2016).
  • Aguado (2017) R. Aguado, “Majorana quasiparticles in condensed matter,” La Rivista del Nuovo Cimento 40, 523 (2017).
  • Lutchyn et al. (2018) R. M. Lutchyn, E. P. A. M. Bakkers, L. P. Kouwenhoven, P. Krogstrup, C. M. Marcus,  and Y. Oreg, “Majorana zero modes in superconductor-semiconductor heterostructures,” Nat. Rev. Mater. 3, 52 (2018).
  • Val’kov et al. (2022) V. V. Val’kov, M. S. Shustin, S. V. Aksenov, A. O. Zlotnikov, A. D. Fedoseev, V. A. Mitskan,  and M. Yu. Kagan, “Topological superconductivity and Majorana states in low-dimensional systems,” Phys. Usp. 65, 2 (2022).
  • Majorana (1937) E. Majorana, “Teoria simmetrica dell’elettrone e del positrone,” Nuovo Cimento 14, 171 (1937).
  • Yu. Kitaev (2003) A. Yu. Kitaev, “Fault-tolerant quantum computation by anyons,” Ann. Phys. 303, 2 (2003).
  • Liu and Baranger (2011) D. E. Liu and H. U. Baranger, “Detecting a Majorana-fermion zero mode using a quantum dot,” Phys. Rev. B 84, 201308(R) (2011).
  • Fidkowski et al. (2012) L. Fidkowski, J. Alicea, N. H. Lindner, R. M. Lutchyn,  and M. P. A. Fisher, “Universal transport signatures of Majorana fermions in superconductor-Luttinger liquid junctions,” Phys. Rev. B 85, 245121 (2012).
  • Prada et al. (2012) E. Prada, P. San-Jose,  and R. Aguado, “Transport spectroscopy of NSNS nanowire junctions with Majorana fermions,” Phys. Rev. B 86, 180503(R) (2012).
  • Pientka et al. (2012) F. Pientka, G. Kells, A. Romito, P. W. Brouwer,  and F. von Oppen, “Enhanced zero-bias Majorana peak in the differential tunneling conductance of disordered multisubband quantum-wire/superconductor junctions,” Phys. Rev. Lett. 109, 227006 (2012).
  • Lin et al. (2012) C.-H. Lin, J. D. Sau,  and S. Das Sarma, “Zero-bias conductance peak in Majorana wires made of semiconductor/superconductor hybrid structures,” Phys. Rev. B 86, 224511 (2012).
  • Lee et al. (2013) M. Lee, J. S. Lim,  and R. López, “Kondo effect in a quantum dot side-coupled to a topological superconductor,” Phys. Rev. B 87, 241402(R) (2013).
  • Kundu and Seradjeh (2013) A. Kundu and B. Seradjeh, “Transport signatures of Floquet Majorana fermions in driven topological superconductors,” Phys. Rev. Lett. 111, 136402 (2013).
  • Vernek et al. (2014) E. Vernek, P. H. Penteado, A. C. Seridonio,  and J. C. Egues, “Subtle leakage of a Majorana mode into a quantum dot,” Phys. Rev. B 89, 165314 (2014).
  • Ilan et al. (2014) R. Ilan, J. H. Bardarson, H.-S. Sim,  and J. E. Moore, “Detecting perfect transmission in Josephson junctions on the surface of three dimensional topological insulators,” New J. Phys. 16, 053007 (2014).
  • Cheng et al. (2014) M. Cheng, M. Becker, B. Bauer,  and R. M. Lutchyn, “Interplay between Kondo and Majorana interactions in quantum dots,” Phys. Rev. X 4, 031051 (2014).
  • Lobos and Das Sarma (2015) A. M. Lobos and S. Das Sarma, “Tunneling transport in NSN Majorana junctions across the topological quantum phase transition,” New J. Phys. 17, 065010 (2015).
  • Peng et al. (2015) Y. Peng, F. Pientka, Y. Vinkler-Aviv, L. I. Glazman,  and F. von Oppen, “Robust Majorana conductance peaks for a superconducting lead,” Phys. Rev. Lett. 115, 266804 (2015).
  • Sharma and Tewari (2016) G. Sharma and S. Tewari, “Tunneling conductance for Majorana fermions in spin-orbit coupled semiconductor-superconductor heterostructures using superconducting leads,” Phys. Rev. B 93, 195161 (2016).
  • van Heck et al. (2016) B. van Heck, R. M. Lutchyn,  and L. I. Glazman, “Conductance of a proximitized nanowire in the Coulomb blockade regime,” Phys. Rev. B 93, 235431 (2016).
  • Sarma et al. (2016) S. Das Sarma, A. Nag,  and J. D. Sau, “How to infer non-Abelian statistics and topological visibility from tunneling conductance properties of realistic Majorana nanowires,” Phys. Rev. B 94, 035143 (2016).
  • Lutchyn and Glazman (2017) R. M. Lutchyn and L. I. Glazman, “Transport through a Majorana island in the strong tunneling regime,” Phys. Rev. Lett. 119, 057002 (2017).
  • Liu et al. (2017) C.-X. Liu, J. D. Sau, T. D. Stanescu,  and S. Das Sarma, “Andreev bound states versus Majorana bound states in quantum dot-nanowire-superconductor hybrid structures: Trivial versus topological zero-bias conductance peaks,” Phys. Rev. B 96, 075161 (2017).
  • Huang et al. (2017) H. Huang, Q.-F. Liang, D.-X. Yao,  and Z. Wang, “Majorana ϕ0\phi_{0}-junction in a disordered spin-orbit coupling nanowire with tilted magnetic field,” Physica C: Superconductivity and its Applications 543, 22 (2017).
  • Liu et al. (2018) C.-X. Liu, J. D. Sau,  and S. Das Sarma, “Distinguishing topological Majorana bound states from trivial Andreev bound states: Proposed tests through differential tunneling conductance spectroscopy,” Phys. Rev. B 97, 214502 (2018).
  • Lai et al. (2019) Y.-H. Lai, J. D. Sau,  and S. Das Sarma, “Presence versus absence of end-to-end nonlocal conductance correlations in Majorana nanowires: Majorana bound states versus Andreev bound states,” Phys. Rev. B 100, 045302 (2019).
  • Tang and Mao (2020) L.-W. Tang and W.-G. Mao, “Detection of Majorana bound states by sign change of the tunnel magnetoresistance in a quantum dot coupled to ferromagnetic electrodes,” Front. Phys. 8, 147 (2020).
  • Zhang and Spånslätt (2020) G. Zhang and C. Spånslätt, “Distinguishing between topological and quasi Majorana zero modes with a dissipative resonant level,” Phys. Rev. B 102, 045111 (2020).
  • Chi et al. (2021) F. Chi, T.-Y. He,  and G. Zhou, “Photon-assisted average current through a quantum dot coupled to Majorana bound states,” J. Nanoelectron. Optoelectron. 16, 1325 (2021).
  • Wang and Huang (2021) Z.-H. Wang and W.-C. Huang, “Dual negative differential of heat generation in a strongly correlated quantum dot side-coupled to Majorana bound states,” Front. Phys. 9, 727934 (2021).
  • Galambos et al. (2022) T. H. Galambos, F. Ronetti, B. Hetényi, D. Loss,  and J. Klinovaja, “Crossed Andreev reflection in spin-polarized chiral edge states due to the Meissner effect,” Phys. Rev. B 106, 075410 (2022).
  • Jin and Li (2022) J. Jin and X.-Q. Li, “Master equation approach for transport through Majorana zero modes,” New J. Phys. 24, 093009 (2022).
  • Zou et al. (2023a) W.-K. Zou, N.-W. Li,  and F.-L. Chong, “Charge and spin transports through a normal lead coupled to an s-wave superconductor and Majorana fermions,” Phys. Status Solidi B , 2200472 (2023a).
  • Leijnse (2014) M. Leijnse, “Thermoelectric signatures of a Majorana bound state coupled to a quantum dot,” New J. Phys. 16, 015029 (2014).
  • López et al. (2014) R. López, M. Lee, L. Serra,  and J. S. Lim, “Thermoelectrical detection of Majorana states,” Phys. Rev. B 89, 205418 (2014).
  • Khim et al. (2015) H. Khim, R. López, J. S. Lim,  and M. Lee, “Thermoelectric effect in the Kondo dot side-coupled to a Majorana mode,” Eur. Phys. J. B 88, 151 (2015).
  • Ramos-Andrade et al. (2016) J. P. Ramos-Andrade, O. Ávalos-Ovando, P. A. Orellana,  and S. E. Ulloa, “Thermoelectric transport through Majorana bound states and violation of Wiedemann-Franz law,” Phys. Rev. B 94, 155436 (2016).
  • Smirnov (2020) S. Smirnov, “Dual Majorana universality in thermally induced nonequilibrium,” Phys. Rev. B 101, 125417 (2020).
  • He et al. (2021) T.-Y. He, H. Sun,  and G. Zhou, “Photon-assisted Seebeck effect in a quantum dot coupled to Majorana zero modes,” Front. Phys. 9, 687438 (2021).
  • Giuliano et al. (2022) D. Giuliano, A. Nava, R. Egger, P. Sodano,  and F. Buccheri, “Multiparticle scattering and breakdown of the Wiedemann-Franz law at a junction of NN interacting quantum wires,” Phys. Rev. B 105, 035419 (2022).
  • Buccheri et al. (2022) F. Buccheri, A. Nava, R. Egger, P. Sodano,  and D. Giuliano, “Violation of the Wiedemann-Franz law in the topological Kondo model,” Phys. Rev. B 105, L081403 (2022).
  • Majek et al. (2022) P. Majek, K. P. Wójcik,  and I. Weymann, “Spin-resolved thermal signatures of Majorana-Kondo interplay in double quantum dots,” Phys. Rev. B 105, 075418 (2022).
  • Bondyopadhaya and Roy (2022) N. Bondyopadhaya and D. Roy, “Nonequilibrium electrical, thermal and spin transport in open quantum systems of topological superconductors, semiconductors and metals,” J. Stat. Phys. 187, 11 (2022).
  • Zou et al. (2022) W.-K. Zou, Q. Wang,  and H.-K. Zhao, “Aharonov-Bohm oscillations in the Majorana fermion modulated charge and heat transports through a double-quantum-dot interferometer,” Phys. Lett. A 443, 128219 (2022).
  • Mourik et al. (2012) V. Mourik, K. Zuo, S. M. Frolov, S. R. Plissard, E. P. A. M. Bakkers,  and L. P. Kouwenhoven, “Signatures of Majorana fermions in hybrid superconductor-semiconductor nanowire devices,” Science 336, 1003 (2012).
  • Nadj-Perge et al. (2014) S. Nadj-Perge, I. K. Drozdov, J. Li, H. Chen, S. Jeon, J. Seo, A. H. MacDonald, B. A. Bernevig,  and A. Yazdani, “Observation of Majorana fermions in ferromagnetic atomic chains on a superconductor,” Science 346, 602 (2014).
  • Wang et al. (2022) Z. Wang, H. Song, D. Pan, Z. Zhang, W. Miao, R. Li, Z. Cao, G. Zhang, L. Liu, L. Wen, R. Zhuo, D. E. Liu, K. He, R. Shang, J. Zhao,  and H. Zhang, “Plateau regions for zero-bias peaks within 5% of the quantized conductance value 2e2/h2e^{2}/h,” Phys. Rev. Lett. 129, 167702 (2022).
  • Yu et al. (2021) P. Yu, J. Chen, M. Gomanko, G. Badawy, E. P. A. M. Bakkers, K. Zuo, V. Mourik,  and S. M. Frolov, “Non-Majorana states yield nearly quantized conductance in proximatized nanowires,” Nat. Phys. 17, 482 (2021).
  • Frolov (2021) S. Frolov, “Quantum computing’s reproducibility crisis: Majorana fermions,” Nature (London) 592, 350 (2021).
  • Ziesen et al. (2023) A. Ziesen, A. Altland, R. Egger,  and F. Hassler, “Statistical Majorana bound state spectroscopy,” Phys. Rev. Lett. 130, 106001 (2023).
  • Liu et al. (2015a) D. E. Liu, M. Cheng,  and R. M. Lutchyn, “Probing Majorana physics in quantum-dot shot-noise experiments,” Phys. Rev. B 91, 081405(R) (2015a).
  • Liu et al. (2015b) D. E. Liu, A. Levchenko,  and R. M. Lutchyn, “Majorana zero modes choose Euler numbers as revealed by full counting statistics,” Phys. Rev. B 92, 205422 (2015b).
  • Haim et al. (2015) A. Haim, E. Berg, F. von Oppen,  and Y. Oreg, “Current correlations in a Majorana beam splitter,” Phys. Rev. B 92, 245112 (2015).
  • Valentini et al. (2016) S. Valentini, M. Governale, R. Fazio,  and F. Taddei, “Finite-frequency noise in a topological superconducting wire,” Physica E 75, 15 (2016).
  • Zazunov et al. (2016) A. Zazunov, R. Egger,  and A. Levy Yeyati, “Low-energy theory of transport in Majorana wire junctions,” Phys. Rev. B 94, 014502 (2016).
  • Smirnov (2017) S. Smirnov, “Non-equilibrium Majorana fluctuations,” New J. Phys. 19, 063020 (2017).
  • Jonckheere et al. (2019) T. Jonckheere, J. Rech, A. Zazunov, R. Egger, A. Levy Yeyati,  and T. Martin, “Giant shot noise from Majorana zero modes in topological trijunctions,” Phys. Rev. Lett. 122, 097003 (2019).
  • Smirnov (2019a) S. Smirnov, “Majorana finite-frequency nonequilibrium quantum noise,” Phys. Rev. B 99, 165427 (2019a).
  • Manousakis et al. (2020) J. Manousakis, C. Wille, A. Altland, R. Egger, K. Flensberg,  and F. Hassler, “Weak measurement protocols for Majorana bound state identification,” Phys. Rev. Lett. 124, 096801 (2020).
  • Feng and Zhang (2022) G.-H. Feng and H.-H. Zhang, “Probing robust Majorana signatures by crossed Andreev reflection with a quantum dot,” Phys. Rev. B 105, 035148 (2022).
  • Smirnov (2022) S. Smirnov, “Revealing universal Majorana fractionalization using differential shot noise and conductance in nonequilibrium states controlled by tunneling phases,” Phys. Rev. B 105, 205430 (2022).
  • Smirnov (2015) S. Smirnov, “Majorana tunneling entropy,” Phys. Rev. B 92, 195312 (2015).
  • Sela et al. (2019) E. Sela, Y. Oreg, S. Plugge, N. Hartman, S. Lüscher,  and J. Folk, “Detecting the universal fractional entropy of Majorana zero modes,” Phys. Rev. Lett. 123, 147702 (2019).
  • Smirnov (2021a) S. Smirnov, “Majorana entropy revival via tunneling phases,” Phys. Rev. B 103, 075440 (2021a).
  • Smirnov (2021b) S. Smirnov, “Majorana ensembles with fractional entropy and conductance in nanoscopic systems,” Phys. Rev. B 104, 205406 (2021b).
  • Ahari et al. (2021) M. T. Ahari, S. Zhang, J. Zou,  and Y. Tserkovnyak, “Biasing topological charge injection in topological matter,” Phys. Rev. B 104, L201401 (2021).
  • Hartman et al. (2018) N. Hartman, C. Olsen, S. Lüscher, M. Samani, S. Fallahi, G. C. Gardner, M. Manfra,  and J. Folk, “Direct entropy measurement in a mesoscopic quantum system,” Nat. Phys. 14, 1083 (2018).
  • Kleeorin et al. (2019) Y. Kleeorin, H. Thierschmann, H. Buhmann, A. Georges, L. W. Molenkamp,  and Y. Meir, “How to measure the entropy of a mesoscopic system via thermoelectric transport,” Nat. Commun. 10, 5801 (2019).
  • Pyurbeeva and Mol (2021) E. Pyurbeeva and J. A. Mol, “A thermodynamic approach to measuring entropy in a few-electron nanodevice,” Entropy 23, 640 (2021).
  • Child et al. (2022a) T. Child, O. Sheekey, S. Lüscher, S. Fallahi, G. C. Gardner, M. Manfra,  and J. Folk, “A robust protocol for entropy measurement in mesoscopic circuits,” Entropy 24, 417 (2022a).
  • Han et al. (2022) C. Han, Z. Iftikhar, Y. Kleeorin, A. Anthore, F. Pierre, Y. Meir, A. K. Mitchell,  and E. Sela, “Fractional entropy of multichannel Kondo systems from conductance-charge relations,” Phys. Rev. Lett. 128, 146803 (2022).
  • Pyurbeeva et al. (2022) E. Pyurbeeva, J. A. Mol,  and P. Gehring, “Electronic measurements of entropy in meso- and nanoscale systems,” Chem. Phys. Rev. 3, 041308 (2022).
  • Child et al. (2022b) T. Child, O. Sheekey, S. Lüscher, S. Fallahi, G. C. Gardner, M. Manfra, A. Mitchell, E. Sela, Y. Kleeorin, Y. Meir,  and J. Folk, “Entropy measurement of a strongly coupled quantum dot,” Phys. Rev. Lett. 129, 227702 (2022b).
  • Smirnov (2018) S. Smirnov, “Universal Majorana thermoelectric noise,” Phys. Rev. B 97, 165434 (2018).
  • Smirnov (2019b) S. Smirnov, “Dynamic Majorana resonances and universal symmetry of nonequilibrium thermoelectric quantum noise,” Phys. Rev. B 100, 245410 (2019b).
  • Cao et al. (2023) Z. Cao, G. Zhang, H. Zhang, Y.-X. Liang, W.-X. He, K. He,  and D. E. Liu, “Towards strong evidence of Majorana bound state by a combined measurement of tunneling conductance and current noise,” arXiv:2301.06451v1  (2023).
  • Deng et al. (2016) M. T. Deng, S. Vaitiekėnas, E. B. Hansen, J. Danon, M. Leijnse, K. Flensberg, J. Nygård, P. Krogstrup,  and C. M. Marcus, “Majorana bound state in a coupled quantum-dot hybrid-nanowire system,” Science 354, 1557 (2016).
  • Deng et al. (2018) M.-T. Deng, S. Vaitiekėnas, E. Prada, P. San-Jose, J. Nygård, P. Krogstrup, R. Aguado,  and C. M. Marcus, “Nonlocality of Majorana modes in hybrid nanowires,” Phys. Rev. B 98, 085125 (2018).
  • Ralph and Buhrman (1994) D. C. Ralph and R. A. Buhrman, “Kondo-assisted and resonant tunneling via a single charge trap: A realization of the Anderson model out of equilibrium,” Phys. Rev. Lett. 72, 3401 (1994).
  • Goldhaber-Gordon et al. (1998) D. Goldhaber-Gordon, H. Shtrikman, D. Mahalu, D. Abusch-Magder, U. Meirav,  and M. A. Kastner, “Kondo effect in a single-electron transistor,” Nature (London) 391, 156 (1998).
  • Meir et al. (1993) Y. Meir, N. S. Wingreen,  and P. A. Lee, “Low-temperature transport through a quantum dot: The Anderson model out of equilibrium,” Phys. Rev. Lett. 70, 2601 (1993).
  • Wingreen and Meir (1994) N. S. Wingreen and Y. Meir, “Anderson model out of equilibrium: Noncrossing-approximation approach to transport through a quantum dot,” Phys. Rev. B 49, 11040 (1994).
  • Smirnov and Grifoni (2011a) S. Smirnov and M. Grifoni, “Slave-boson Keldysh field theory for the Kondo effect in quantum dots,” Phys. Rev. B 84, 125303 (2011a).
  • Smirnov and Grifoni (2011b) S. Smirnov and M. Grifoni, “Kondo effect in interacting nanoscopic systems: Keldysh field integral theory,” Phys. Rev. B 84, 235314 (2011b).
  • Niklas et al. (2016) M. Niklas, S. Smirnov, D. Mantelli, M. Margańska, N.-V. Nguyen, W. Wernsdorfer, J.-P. Cleuziou,  and M. Grifoni, “Blocking transport resonances via Kondo many-body entanglement in quantum dots,” Nat. Commun. 7, 12442 (2016).
  • Ruiz-Tijerina et al. (2015) D. A. Ruiz-Tijerina, E. Vernek, L. G. G. V. Dias da Silva,  and J. C. Egues, “Interaction effects on a Majorana zero mode leaking into a quantum dot,” Phys. Rev. B 91, 115435 (2015).
  • Altland and Simons (2010) A. Altland and B. Simons, Condensed Matter Field Theory, 2nd ed. (Cambridge University Press, Cambridge, 2010).
  • Xu et al. (2022) Y. Xu, J. Jin, S. Wang,  and Y. J. Yan, “Memory-effect-preserving quantum master equation approach to noise spectrum of transport current,” Phys. Rev. E 106, 064130 (2022).
  • Basset et al. (2012) J. Basset, A. Y. Kasumov, C. P. Moca, G. Zaránd, P. Simon, H. Bouchiat,  and R. Deblock, “Measurement of quantum noise in a carbon nanotube quantum dot in the Kondo regime,” Phys. Rev. Lett. 108, 046802 (2012).
  • Ashkroft and Mermin (1976) N. W. Ashkroft and N. D. Mermin, Solid State Physics (Saunders College, Philadelphia, 1976).
  • Zou et al. (2023b) W.-K. Zou, Q. Wang,  and H.-K. Zhao, “Dynamic heat and charge transports through double-quantum-dot-interferometer modulated by Majorana bound states and time-oscillating Aharonov-Bohm flux,” J. Phys.: Condens. Matter 35, 165303 (2023b).
  • Krainov et al. (2022) I. V. Krainov, A. P. Dmitriev,  and N. S. Averkiev, “Shot noise in resonant tunneling: Role of inelastic scattering,” Phys. Rev. B 106, 245421 (2022).
  • Tsintzis et al. (2022) A. Tsintzis, R. S. Souto,  and M. Leijnse, “Creating and detecting poor man’s Majorana bound states in interacting quantum dots,” Phys. Rev. B 106, L201404 (2022).
  • Dvir et al. (2023) T. Dvir, G. Wang, N. van Loo, C.-X. Liu, G. P. Mazur, A. Bordin, S. L. D. ten Haaf, J.-Y. Wang, D. van Driel, F. Zatelli, X. Li, F. K. Malinowski, S. Gazibegovic, G. Badawy, E. P. A. M. Bakkers, M. Wimmer,  and L. P. Kouwenhoven, “Realization of a minimal Kitaev chain in coupled quantum dots,” Nature 614, 445 (2023).