10
arXiv:1509.05756v1 [cond-mat.mes-hall] 18 Sep 2015 Time dependent heat flow in interacting quantum conductors Guillem Rossell´ o, 1, 2 Rosa L´ opez, 2 and Jong Soo Lim 3 1 Institut f¨ ur Theorie der Statistischen Physik, RWTH Aachen University, D-52056 Aachen, JARA - Future Information Technologies, Germany 2 Institut de F´ ısica Interdisciplin`aria i de Sistemes Complexos IFISC (CSIC-UIB), E-07122 Palma de Mallorca, Spain 3 School of Physics, Korea Institute for Advanced Study, Seoul 130-722, Korea We derive the frequency-resolved heat current expression in the linear response regime for a setup comprised of reservoir, interacting central site, and tunneling barrier under the action of a time de- pendent electrical signal. We exploit the frequency parity properties of response functions to obtain the heat current expression for interacting quantum conductors. Importantly, the corresponding heat formula, valid for arbitrary AC frequencies, can describe photon-assisted heat transport. In particular, we analyze the heat transfer for an interacting multilevel conductor (a carbon nanotube quantum dot) coupled to a single reservoir. We show that the electrothermal admittance can reverse its sign by properly tunning the AC frequency. I. INTRODUCTION The expeditious advance of circuit miniaturization re- quires the knowledge of heat flow in quantum devices in response to electrical fields. 1 Due to this necessity thermoelectrical transport in quantum devices is a re- search field that has vigorously relaunched nowadays. 2 Nevertheless, so far, the activity has focused on driv- ing electrical currents by oscillatory forces 3,4 for charge qubit manipulation 5 , quantum emitter generation 6,7 or quantum tomography purposes. 8 Low frequency mea- surements of the electrical admittance in quantum RC circuits provide information about the spectroscopic (C G quantum capacitance) and resistive (R G relaxation re- sistance) properties of coherent conductors 9–11 in which R G = h/2e 2 takes an universal value. 12–16 Interactions, such as charging effects 17 or Kondo correlations 18,19 do not alter the universality of R G , measured ex- perimentally, in quantum capacitors, 9–11,20,21 carbon nanotubes, 22 superconducting junctions, 23 and quantum dots. 24 In contrast, time-resolved heat transport has been poorly investigated. 25–27 Solely, stationary or time- averaged heat flows have been analyzed in more detail. 28 The understanding of time dependent heat currents opens an avenue of creating circuit architectures where heat absorption or emission events are finely tunable via electrical and thermal time dependent signals. Recently, the linear response for the charge and heat fluxes to electrical and thermal AC signals was computed for a quantum capacitor showing that heat flows can be de- layed or elapsed with respect to the AC pulse depending on the dot gate position. 26 Later on, Ludovico et al. 27 showed, for a slow AC modulation and noninteracting conductors, that time-dependent heat flow J R (t) needs to explicitly consider the energy stored and relaxed at the tunnel coupling region when a tunnel Hamiltonian description is employed. The goal that we face consists of the calculation of the heat flux in a completely dif- ferent regime. We derive the time dependent heat flow for interacting and multi-orbital conductors and arbi- FIG. 1: Schematic of the multi-orbital interacting quantum capacitor. It is shown two situations where the AC voltage is applied either (a) to the dot or (b) to the reservoir. Both cases are equivalent by means of a gauge transformation. U (t) denotes the time-dependent internal potential of the dot as response of the charge injected by the AC driving field V (t). We model such response with a capacitance C. We illustrate in (b) the induced photon-assisted emission and absorption processes (shown by wiggle lines) in the heat by the action of V (t). QR, QD, and QT are the energy change rates in time at the reservoir, quantum dot, and tunneling barriers. trary AC frequencies. Since our heat formula holds for arbitrary high frequencies, it is able to describe photon- assisted tunneling processes occurring in the heat trans- port. Our only limitation is that we are restricted to the linear response regime, i.e., low AC driving amplitudes. We confine our interests to the Coulomb blockade regime and do not consider higher-order correlations like Kondo effect, we implicitly assume a temperature higher than the Kondo scale (T K ), i.e., T T K . For our calcula- tions, we employ the nonequilibrium (Keldysh) Green’s function formalism 29 which allows us to include electron- electron interactions, charging effects, in a feasible way within the so-called Hartree-Fock regime. To address these issues, we focus on a prototypical interacting multi-orbital quantum circuit – a quantum capacitor– formed by a carbon nanotube quantum dot that is coupled to a single terminal being modulated by an AC voltage, as shown in Fig. 1. Carbon nanotubes

Time dependent heat flow in interacting quantum conductors

  • Upload
    la

  • View
    215

  • Download
    2

Embed Size (px)

DESCRIPTION

Time dependent heat flow in interacting quantum conductors

Citation preview

Page 1: Time dependent heat flow in interacting quantum conductors

arX

iv:1

509.

0575

6v1

[co

nd-m

at.m

es-h

all]

18

Sep

2015

Time dependent heat flow in interacting quantum conductors

Guillem Rossello,1, 2 Rosa Lopez,2 and Jong Soo Lim3

1Institut fur Theorie der Statistischen Physik, RWTH Aachen University,D-52056 Aachen, JARA - Future Information Technologies, Germany

2Institut de Fısica Interdisciplinaria i de Sistemes Complexos IFISC (CSIC-UIB), E-07122 Palma de Mallorca, Spain3School of Physics, Korea Institute for Advanced Study, Seoul 130-722, Korea

We derive the frequency-resolved heat current expression in the linear response regime for a setupcomprised of reservoir, interacting central site, and tunneling barrier under the action of a time de-pendent electrical signal. We exploit the frequency parity properties of response functions to obtainthe heat current expression for interacting quantum conductors. Importantly, the correspondingheat formula, valid for arbitrary AC frequencies, can describe photon-assisted heat transport. Inparticular, we analyze the heat transfer for an interacting multilevel conductor (a carbon nanotubequantum dot) coupled to a single reservoir. We show that the electrothermal admittance can reverseits sign by properly tunning the AC frequency.

I. INTRODUCTION

The expeditious advance of circuit miniaturization re-quires the knowledge of heat flow in quantum devicesin response to electrical fields.1 Due to this necessitythermoelectrical transport in quantum devices is a re-search field that has vigorously relaunched nowadays.2

Nevertheless, so far, the activity has focused on driv-ing electrical currents by oscillatory forces3,4 for chargequbit manipulation5, quantum emitter generation6,7 orquantum tomography purposes.8 Low frequency mea-surements of the electrical admittance in quantum RCcircuits provide information about the spectroscopic (CG

quantum capacitance) and resistive (RG relaxation re-sistance) properties of coherent conductors9–11 in whichRG = h/2e2 takes an universal value.12–16 Interactions,such as charging effects17 or Kondo correlations18,19

do not alter the universality of RG , measured ex-perimentally, in quantum capacitors,9–11,20,21 carbonnanotubes,22 superconducting junctions,23 and quantumdots.24

In contrast, time-resolved heat transport has beenpoorly investigated.25–27 Solely, stationary or time-averaged heat flows have been analyzed in more detail.28

The understanding of time dependent heat currentsopens an avenue of creating circuit architectures whereheat absorption or emission events are finely tunable viaelectrical and thermal time dependent signals. Recently,the linear response for the charge and heat fluxes toelectrical and thermal AC signals was computed for aquantum capacitor showing that heat flows can be de-layed or elapsed with respect to the AC pulse dependingon the dot gate position.26 Later on, Ludovico et al.27

showed, for a slow AC modulation and noninteractingconductors, that time-dependent heat flow JR(t) needsto explicitly consider the energy stored and relaxed atthe tunnel coupling region when a tunnel Hamiltoniandescription is employed. The goal that we face consistsof the calculation of the heat flux in a completely dif-ferent regime. We derive the time dependent heat flowfor interacting and multi-orbital conductors and arbi-

FIG. 1: Schematic of the multi-orbital interacting quantumcapacitor. It is shown two situations where the AC voltageis applied either (a) to the dot or (b) to the reservoir. Bothcases are equivalent by means of a gauge transformation. U(t)denotes the time-dependent internal potential of the dot asresponse of the charge injected by the AC driving field V (t).We model such response with a capacitance C. We illustratein (b) the induced photon-assisted emission and absorptionprocesses (shown by wiggle lines) in the heat by the action ofV (t). QR, QD, and QT are the energy change rates in timeat the reservoir, quantum dot, and tunneling barriers.

trary AC frequencies. Since our heat formula holds forarbitrary high frequencies, it is able to describe photon-assisted tunneling processes occurring in the heat trans-port. Our only limitation is that we are restricted to thelinear response regime, i.e., low AC driving amplitudes.We confine our interests to the Coulomb blockade regimeand do not consider higher-order correlations like Kondoeffect, we implicitly assume a temperature higher thanthe Kondo scale (TK), i.e., T ≫ TK . For our calcula-tions, we employ the nonequilibrium (Keldysh) Green’sfunction formalism29 which allows us to include electron-electron interactions, charging effects, in a feasible waywithin the so-called Hartree-Fock regime.To address these issues, we focus on a prototypical

interacting multi-orbital quantum circuit– a quantumcapacitor– formed by a carbon nanotube quantum dotthat is coupled to a single terminal being modulated byan AC voltage, as shown in Fig. 1. Carbon nanotubes

Page 2: Time dependent heat flow in interacting quantum conductors

2

exhibit charging effects due to the formation of quantumdots inside the tube.30 The valley degrees of freedom,corresponding to the K and K’ Dirac points in graphenein addition to the spin indices, lead to a four-fold en-ergy level degeneracy. Such degeneracy can be liftedby the presence of an external magnetic field B.31 Fur-thermore, the nanotube curvature yields a spin-orbit in-teraction resulting in split time-reversal dot level pairs.Therefore, carbon nanotube quantum dots act as multi-orbital interacting conductors where Coulomb interac-tions give rise to Coulomb blockade phenomena. Evenmore importantly, carbon nanotube quantum dots havebeen demonstrated to be perfect platforms to investigatethe frequency-resolved transport when they are embed-ded in electromagnetic cavities,22 precisely the issue thatour work addresses.In Sec. II, we describe the theoretical model for

an interacting multi-orbital quantum capacitor and theHartree-Fock decomposition of electron-electron interac-tion. We derive the time-dependent heat flux for thereservoir in the linear response regime and obtain theelectrothermal admittance. For completeness we also dis-cuss the electrical admittance. In Sec. III, we start byproving that our formalism can exhibit Coulomb block-ade phenomena. The observations for a single orbitalquantum dot are briefly discussed and then results for amulti-orbital one are presented. Here, we consider a fourlevel degenerated carbon nanotube quantum dot whichsplits due to the presence of magnetic field and spin-orbitinteraction. For this system, we show the RC parametersand investigate the electrical and electrothermal admit-tances. Our findings are summarized in Sec. IV.

II. TIME-DEPENDENT TRANSPORT

FORMULATION

The purpose is to derive the linear response heat cur-rent for an interacting conductor coupled to a reservoirthat oscillates with AC voltage signal in the Coulombblockade regime. We first propose a model Hamiltoniandescribing a multi-orbital quantum capacitor which is at-tached to a single reservoir. Using the Keldysh formalismfor nonequilibrium Green’s functions, we have the linearresponse heat flux at the reservoir subject to an oscillat-ing electrical signal. Remarkably, the heat flux can becast in terms of the conductor Green’s function.

A. Model of Hamiltonian

The starting point of our derivation is the tunnelHamiltonian description of a quantum capacitor. Thequantum capacitor Hamiltonian H is split into threeparts, namely the reservoir part (HR), the dot contri-bution (HD), and the tunneling term (HT ), i.e.,

H = HR +HD +HT . (1)

More concretely, the reservoir part reads

HR =∑

k,σ

[ǫkσ − (µ+ eV (t))] c†kσckσ , (2)

where µ = EF + eVdc is the chemical potential, e > 0 theunit charge, and V (t) is the electrical voltage modula-tion [see Fig. 1 (b)]. As shown in Fig. 1, notice that thisdescription is equivalent to the situation where the ACvoltage is applied to the dot since a gauge transformationconnects both situations and we thus employ two situa-tions interchangeably. Without loss of generality, we set

Vdc = 0. The operator c†kσ(ckσ) creates (annihilates) anelectron with wavevector k and spin σ in the reservoir.For the dot contribution describing an interacting systemwith n levels, we have

HD =∑

n,σ

ǫnσd†nσdnσ + EC [Nd +N (t)]

2, (3)

where d†nσ(dnσ) corresponds to the creation (annihila-tion) operator for a dot electron in the nth level withspin σ, ǫnσ denotes the single-particle energies, andEC = e2/2C is the electrostatic charging energy (C isthe dot geometrical capacitance). Here, the dot occu-pation operator reads Nd =

∑n,σ d

†nσdnσ. Finally, the

tunneling term hybridizes the reservoir and the dot sub-systems according to

HT =∑

n,k,σ

tnkσ

(c†kσcnσ +H.c

). (4)

where tnkσ denotes the tunneling amplitude.When charge is injected in the dot by an external

source V (t), a polarization charge is created to keep thedot as a neutral charge object. In a simple electrostaticpicture, we model such polarization charge with a capac-itance C leading to32

eNd(t) = CU(t) , (5)

in which U(t) is the internal potential of the dot givingrise to a time-dependent potential inside the dot. Then,the dot Hamiltonian in Eq. (3), up to linear order inU(t), can be written as

HD =∑

n,σ

[ǫnσ + eU(t)] d†nσdnσ + ECN2d . (6)

Since N2d =

∑m,σ

∑n,σ′ d†mσdmσd

†nσ′dnσ′ is a quartic op-

erator, the Hamiltonian cannot be solved without intro-ducing some proper approximation.

B. Hartree Fock approximation

We thus perform the Hartree-Fock approximation inthe dot Hamiltonian. For the Hartree approximation,N2

d is decoupled in the form

[N2d ]Hartree = 2

m,σ

n,σ′

d†mσdmσ〈d†nσ′dnσ′ 〉 , (7)

Page 3: Time dependent heat flow in interacting quantum conductors

3

while for the Fock approximation one has

[N2d ]Fock = −2

m,σ

n,σ′

d†mσdnσ′ 〈d†mσdnσ′〉 . (8)

Considering Hartree and Fock approximations (HF) to-gether, the dot Hamiltonian then becomes

HD =∑

m,n,σ

ǫmnσ(t)d†mσdnσ , (9)

where

ǫmnσ(t) = δm,n

ǫmσ + eU(t) + 2EC

l,s( 6=m,σ)

〈d†lsdls〉

− 2EC〈d†mσdnσ〉, (10)

with σ =↓ / ↑ for σ =↑ / ↓.An important observation is in order when Eq. (9)

is considered. For illustration, we assume that there isonly a single orbital (ǫmσ = ǫdσ). We allow the pres-ence of a small external magnetic field ǫdσ = ǫd + σ∆Z

(∆Z the Zeeman energy) in order to break explicitly spindegeneracy, i.e., Ndσ 6= Ndσ. For a single orbital, sincethe exchange interaction is absent between electrons withsame spins the Fock term disappears such that the dot’senergy level can be simplified as

ǫdσ(t) = ǫd + σ∆Z + eU(t) + 2EC〈d†σdσ〉 . (11)

We envision the level is occupied by the spin σ electronwith energy ǫdσ + eU(t). When another electron withopposite spin σ enters the level, its energy is increasedby the charging energy 2EC . Therefore, as a functionof the Fermi energy EF the dot occupation then doesnot change continuously, but shows plateaus and discon-tinuous jumps. These jumps are the clear evidence ofCoulomb blockade phenomenon. We illustrate how HFapproximation captures charging effects by considering asingle orbital quantum dot and plotting its total occu-pation Nd when the Fermi energy varies in Fig. 2. Weconsider the absence of AC signal. The dot occupationshows a plateau of Nd ≈ 1 in the Coulomb gap region,i.e., ǫd . EF . ǫd + 2EC . This result corroborates thefact that our HF description can reproduce charging ef-fects and thus Coulomb blockade properly. We emphasizethat our results are not equivalent or expected when theyare compared with previous findings using noninteractingmodels.

C. Heat admittance

Our main interest is focused on the rate of heat atthe reservoir. For such purpose, we evaluate the timederivative of each component of the Hamiltonian whichis given by

QS =d

dt〈HS〉 =

i

~〈[H,HS ]〉+

⟨∂HS

∂t

⟩, (12)

0 5 10E

F/Γ

0

0.5

1

1.5

2

Nd (

occu

patio

n)

2 EC

FIG. 2: Quantum dot occupation Nd versus the Fermi energyEF for a single orbital dot in the Coulomb blockade regime.The dot’s energy level is placed at ǫdσ = ǫd + σ∆Z withǫd = 0 and ∆Z = 0.5Γ. Rest of parameters: 2EC = 10Γand kBT = 0.04Γ.

where S = {R,D, T } is referred to R reservoir, D dot,and T tunneling term. Notice that in Eq. (12) the lastterm is the power supplied by an external source andmust be subtracted from the definition of the heat rateQS(t) (see below). Besides, since the Hamiltonian oper-ator H commutes with itself it is fulfilled

i

~〈[H,H]〉 = QR +QD +QT = 0 . (13)

As mentioned, our goal is to compute QR. WhenCoulomb interaction is taken into account, we obtain therate of heat at the reservoir, QR, in terms of interact-ing dot Green’s functions. Since the direct calculation ofQR is cumbersome, our strategy is first to compute theenergy change rates at the dot and tunnel barrier usingEq. (13)

QR = −(QD +QT ) . (14)

We recall that our calculation applies only for the linearresponse regime and thus we keep only the leading ordercontributions in the fields V (t) and U(t). We employ thenonequilibrium Keldysh Green’s function formalism forthe calculation of the energy change rates in time.29 Westart with the time derivative of HD first,

d

dt〈HD(t)〉 =

m,n,σ

[∂tǫmnσ(t)] 〈d†mσ(t)dnσ(t)〉

+∑

m,n,σ

ǫmnσ(t)∂t〈d†mσ(t)dnσ(t)〉 . (15)

Page 4: Time dependent heat flow in interacting quantum conductors

4

The first term on the right hand side denotes the powerdeveloped by the AC source and does not contribute tothe heat flow. The second term represents the energyflux going into the dot which can be written in terms ofthe dot Green’s function

QD(t) = −iTr[ǫdσ(t)∂tG

<dσ,dσ(t, t)

]. (16)

Hereafter, bold face symbols denote matrices such thattheir (m,n) components are

[ǫdσ(t)]m,n = ǫmnσ(t) ,

[G<dσ,dσ(t, t

′)]m,n = i〈d†nσ(t′)dmσ(t)〉 .

(17)

The trace means summations over energy levels and spinindices, i.e., Tr =

∑m,n

∑σ. Secondly, from the defini-

tion of HT , it can be shown that

〈HT 〉 = −iTr[G<

dσ,dσ(t, t)tσ + t∗σG<cσ,dσ(t, t)

], (18)

and thus the time variation of energy stored or relaxedat tunneling barrier QT (t) = ∂t〈HT 〉 is given by

QT (t) = −i∂tTr[t∗σG

<cσ,dσ(t, t) +G<

dσ,cσ(t, t)tσ]. (19)

Here, the reservoir-dot and dot-reservoir Green’s func-tions are given, respectively, by

[G<cσ,dσ(t, t

′)]k,n = i〈d†nσ(t′)ckσ(t)〉 ,

[G<dσ,cσ(t, t

′)]n,k = i〈c†kσ(t′)dnσ(t)〉 ,

(20)

and [t∗σ]n,k = t∗nkσ, and Tr =∑

k,n,σ.Using the standard technique of equation of motion,

the reservoir-dot Green’s function G<cσ,dσ(t, t) can now

be recast in terms of solely the dot Green’s functions asfollows,

t∗σG<cσ,dσ(t, t) =

∫dt1~

[Σrσ(t, t1)G

<dσ,dσ(t1, t)

+Σ<σ (t, t1)G

adσ,dσ(t1, t)] . (21)

Here, we have defined

Σrσ(t1, t2) = −

i

2~Γδ(t1 − t2) ,

Σ<σ (t1, t2) = iΓe−iφV (t1,t2)f(t1 − t2)

(22)

as the retarded and lesser self-energies17 that contain thetime-dependent fields. We have assumed a momentum-independent tunneling amplitude tnkσ = tnσ leading tohybridization strength given by [Γ]m,n = 2πρ0t

∗mσtnσ

with ρ0 = 1/2D the density of states of the reservoirand D the reservoir bandwidth. In such self-energies, itis defined

φV (t1, t2) =

∫ t1

t2

dt

~eV (t) , (23)

as the time-dependent phase due to the AC external po-tential and

f(t1 − t2) =

∫dǫ

2πe−iǫ(t1−t2)/~f(ǫ) , (24)

being f(t1 − t2) the Fermi-Dirac distribution function inthe time domain. Similarly, the dot-reservoir Green’sfunction G<

dσ,cσ(t, t)tσ is given by

G<dσ,cσ(t, t)tσ =

∫dt1~

[Grdσ,dσ(t, t1)Σ

<σ (t1, t)

+G<dσ,dσ(t, t1)Σ

aσ(t1, t)] . (25)

Employing Eqs. (21) and (25), we find QT is also ex-pressed only in terms of the dot Green’s functions.Once the heat energy rates for the dot and the tun-

neling parts (QD, and QT ) are expressed in terms of thedot Green’s function, the next step consists of comput-ing such Green’s functions in the presence of the time-dependent AC signal. For the dot’s retarded/advancedGreen’s function, we have17

Gr/adσ,dσ(t, t

′) = e−iφU (t,t′)Gr/a,eqdσ,dσ (t− t′), (26)

where φU (t, t′) =

∫ t

t′(dt1/~) eU(t1) and Gr/a,eqdσ,dσ (t1 − t2)

denotes the dot’s retarded/advanced Green’s function inequilibrium, i.e., in the absence of AC signal. Finally, thedot’s lesser Green’s function is obtained by means of

G<dσ,dσ(t, t

′) =∫

dt1~

∫dt2~

Grdσ,dσ(t, t1)Σ

<σ (t1, t2)G

adσ,dσ(t2, t

′) .

(27)

Inserting Eqs. (21)-(27) into Eqs. (16) and (19), it canbe shown, after some cumbersome algebra, that to linearorder in V and U the energy change rates in frequencydomain become

QD(ω) = ieωTr{∫

2πǫdσA(ǫ, ~ω)[V (~ω)− U(~ω)]

},

(28a)

QT (ω) = ieωTr{∫

2π(~ω + 2ǫ− 2ǫdσ)A(ǫ, ~ω)

× [V (~ω)− U(~ω)]}, (28b)

where V (~ω), and U(~ω) are the Fourier transforms ofV (t), and U(t) respectively. Here,

A(ǫ, ~ω) = Υ(~ω, ǫ)F(ǫ, ~ω) (29)

with

Υ(~ω, ǫ) =[G

r,eqdσ,dσ(ǫ+ ~ω)ΓGa,eq

dσ,dσ(ǫ)],

F(ǫ, ~ω) =f(ǫ+ ~ω)− f(ǫ)

~ω.

(30)

The dot’s retarded/advanced Green’s function in equilib-rium and in frequency domain is given by

Gr/adσ,dσ(ǫ) =

1

ǫ− ǫdσ ± iΓ/2, (31)

Page 5: Time dependent heat flow in interacting quantum conductors

5

with

[ǫdσ]m,n = δm,n

ǫmσ + 2EC

l,s( 6=m,σ)

〈d†lsdls〉

− 2EC〈d†mσdnσ〉 . (32)

Although Eq. (31) has the same structure with its nonin-teracting retarded/advanced Green’s function, the charg-ing energy is included in the denominator such that itcan describe the Coulomb blockade effect as mentionedabove. To completely determine the dot Green’s func-tion, the diagonal and off-diagonal dot occupations needto be self-consistently calculated using

〈d†nsdms〉 =1

2πi

∫dǫ[G<

dσ,dσ(ǫ)]n,m . (33)

Finally, from Eqs. (28) and (14), we obtain the expres-sion for the energy change rate at the reservoir

QR(ω) = −(QD +QT ) = −ieωTr{∫

(~ω + 2ǫ− ǫdσ)A(ǫ, ~ω)[V (~ω)− U(~ω)]}. (34)

Importantly, neither QR(ω) nor QD(ω) have a well de-fined parity when the AC frequency is reversed, andtherefore, within linear response theory, these two magni-tudes do not represent physical quantities [see Sec. III].Based on these observations, the reservoir and dot fre-quency dependent heat current expressions must be thusmodified to exhibit a proper parity property when theAC frequency is reversed. We find the expressions

JR = QR(ω) +1

2QT (ω), JD = QD(ω) +

1

2QT (ω) .(35)

satisfy the parity property. Remarkably, the choice of thefactor 1

2QT (ω) is unique in order to ensure a well definedparity property in both JR, and JD.Formally, these expressions agree with their non-

interacting counterpart for the time dependent heatcurrents.27 However, it should be noted that our theo-retical analysis (i) goes beyond low AC frequencies incontrast with Ref. [27] in which the heat rate was ob-tained up to second order in the AC frequency using theFloquet theory, (ii) includes the effect of Coulomb block-ade, and (iii) is applicable to multi-orbital conductors incontrast to previous time heat formulations.Remarkably, our AC heat formula contains photon-

assisted tunneling events only possible for sufficientlyhigh AC frequencies. Note that a similar definition forthe time-dependent heat, but applicable to a spin chainmodel (using a tight-binding model), was proposed inRef. [33]. In such work, the heat flux connecting twosites, i and i ± 1, was incorporated to the general heatflow expression with a 1/2 factor in close analogy toEq. (35). However, caution is needed when this com-parison is made. The results for a chain of sites are not

immediately generalized to our continuum model by justkeeping the factor one-half in front of the tunneling en-ergy flow.After these considerations, the formulations for the the

dot and reservoir frequency dependent heat currents inthe linear response regime read

JD = −JR = ieωTr{∫

(~ω

2+ ǫ

)A(ǫ, ~ω)

× [V (~ω)− U(~ω)]}. (36)

Importantly, this formula is the central finding of ourwork.

D. Electrothermal admittance

In the presence of a time-dependent driving force, itis quite general to characterize the transport using theconcept of admittance. The complex electrothermal ad-mittance is defined as

M(ω) =JR(ω)

V (~ω). (37)

Notice that Eq. (37) which can be obtained from Eq. (36)contains the unknown function U(~ω). Therefore, for acomplete characterization of the complex electrothermaladmittance, we need first to determine the internal po-tential U(~ω). For such purpose, we note that the dis-placement current ID can be featured by a capacitance Cin a simple model [here, we consider the situation shownin Fig. 1 (a)]

ID(ω) = −iωCU(~ω). (38)

Due to current conservation, the displacement currentis equal to the tunneling current IT 17 for a quantumcapacitor

IT (ω) = g(ω)[V (~ω)− U(~ω)], (39)

where [see Ref. [17] for its explicit derivation]

g(ω) = ie2ωTr

{∫dǫ

2πA(ǫ, ~ω)

}. (40)

The internal potential is obtained when we impose cur-rent conservation ID = IT , then U(~ω) reads

U(~ω) =g(ω)V (~ω)

−iωC + g(ω). (41)

Inserting Eq. (41) into Eq. (36) completely characterizesthe linear response of the heat current to a time depen-dent voltage, i.e., the electrothermal admittance

M(ω) = m(ω)−iωC

[−iωC + g(ω)], (42)

Page 6: Time dependent heat flow in interacting quantum conductors

6

with

m(ω) = ieωTr

{∫dǫ

(~ω

2+ ǫ

)A(ǫ, ~ω)

}. (43)

Similarly, the thermoelectrical admittance is defined as

L(ω) =IT (ω)

T (ω), (44)

with T (ω) being the Fourier transform of a time modu-lated temperature T (t). Remarkably, the thermoelectri-cal and the electrothermal admittances are reciprocallyrelated due to the microreversibility principle

M(ω) = TL(ω). (45)

Notice that microreversibility principle only holds at lin-ear order in V (t), and U(t). Finally, a second order ex-pansion in the AC frequency for m(ω)

m(ω) = −iωCM + ω2C2MRM , (46)

allows us to obtain the RC electrothermal parameters.For comparison, we also calculate the electrical admit-tance defined by

G(ω) =IT

V (~ω)= g(ω)

−iωC

[−iωC + g(ω)]. (47)

The corresponding second order expansion of g(ω) in fre-quency

g(ω) = −iωCG + ω2C2GRG , (48)

always yields positive RC parameters, i.e., IT is alwaysdelayed with respect to V (ω). This is in clear contrastto the electrothermal admittance case in which both CM

and RM can be either positive or negative. The heatflow response is elapsed or delayed with respect to theelectrical signal depending on the system parameters.Note however that the product of both quantities, theRC time, is kept always positive as expected. Similarresults were obtained in Ref. [26].

III. RESULTS

Previously, we have derived the formal expressions forthe reservoir heat flow and the corresponding electrother-mal admittance. Now, we apply these formulas to thecase of an interacting multi-orbital conductor. However,before addressing the multi-orbital case, for illustrationpurpose we first investigate a single orbital interactingquantum capacitor under the influence of an AC poten-tial. Later on, we will analyze a prototype of multi-orbital conductor, a carbon nanotube quantum dot at-tached to a single reservoir. In both cases, we show re-sults for the electrical and electrothermal transport.

0 10E

F/Γ

0.25

0.3

0.35

0.4

0.45

0.5

RG

(U

nits

of

h/e2 )

2EC

FIG. 3: Charge relaxation resistance for a single orbital quan-tum capacitor versus the Fermi energy. Parameters: ǫd = 0,∆Z = 0.5Γ, 2EC = 10Γ, and kBT = 0.04Γ.

-4 -3 -2 -1 0 1 2 3 4hω/Γ

0

0.5

1

ℜe

g(ω

) (U

nits

of

e2 /h)

-4 -3 -2 -1 0 1 2 3 4hω/Γ

-0.5

0

0.5

ℑm

g(ω

) (U

nits

of

e2 /h)

(a) (b)

FIG. 4: (a) Real and (b) imaginary parts of g(ω) for a singleorbital quantum capacitor as a function of the AC frequency.Photon-assisted excitations occur at the transition rate ener-gies ±~ω = |EF − ǫdσ| when kBT ≪ Γ. Parameters: ǫd = 0,∆Z = 0.625Γ, 2EC = 10Γ, EF = 0, and kBT = 0.04Γ.

A. Single orbital quantum capacitor

We consider an interacting quantum dot with just oneorbital contacted to a single reservoir. The reservoir hasan electrical potential that oscillates in time. In accor-dance with our previous theoretical considerations, theprefactor g(ω) of the electrical admittance G(ω) can beexpanded in powers of the AC frequency up to secondorder as shown in Eq. (48). A very well known result

Page 7: Time dependent heat flow in interacting quantum conductors

7

-4 -3 -2 -1 0 1 2 3 4hω/Γ

-0.2

-0.1

0

0.1

ℜe

m(ω

) (U

nits

of

e/h)

-4 -3 -2 -1 0 1 2 3 4hω/Γ

-0.3

-0.15

0

0.15

0.3

ℑm

m(ω

) (U

nits

of

e/h)

(a) (b)

FIG. 5: (a) Real and (b) imaginary parts of m(ω) for a singleorbital quantum capacitor versus the AC frequency. Photon-assisted excitations occur at the resonant conditions. Param-eters: ǫd = 0, ∆Z = 0.625Γ, 2EC = 10Γ, EF = 0, andkBT = 0.04Γ.

establishes that RG = h/(2Qe2) becomes universal be-ing Q the number of transport channels. Our exampleas shown in Fig. 3 considers the presence of a Zeemanfield to explicitly break spin degeneracy. RG takes thevalue of h/2e2 when just one of the two spin-resolvedlevels (ǫd↑/↓ = EF ± ∆Z/2) participates in transport,

whereas it becomes h/4e2 when the opposite spin channelalso contributes. Clearly, the two main peaks observedin RG are separated roughly by 2EC the charging energy(2EC ≫ ∆Z). Our results for RG indicate the presence ofcharging effects and the Coulomb blockade phenomenon.More importantly, our major interest resides in the be-

havior of the electrical and electrothermal admittances atarbitrary AC frequencies. Figure 4 shows the real [Fig. 4(a)] and imaginary [Fig. 4 (b)] parts of the prefactorg(ω) in the expression for the electrical admittance G(ω)[see Eq. (47)]. We observe that ℜeg(ω) = ℜeg(−ω) andℑmg(ω) = −ℑmg(−ω). The factor −iωC/(−iωC+g(ω))in Eq. (47) then does not change the parity property ofG(ω) with respect to ω such that for simplicity we con-sider g(ω). Excitations occur when the AC frequencymatches with the resonant condition ±~ω ≈ |ǫdσ − EF |.We recall that ǫdσ is the spin-dependent dot energy levelrenormalized by electron-electron interactions accordingto

ǫdσ = ǫd + σ∆Z + 2EC〈d†σdσ〉 . (49)

For the parameters used in Fig. 4, we obtain ǫdσ ≈0.3Γ, 2.1Γ which agrees with the resonant behavior foundin g(ω) with peaks at ±~ω ≈ 0.3Γ, and ±~ω ≈ 2.1Γ.Similar features are observed in the electrothermal ad-

mittance shown in Fig. 5. Interestingly, the imaginarypart of m(ω) for ω > 0 takes either positive or negativevalues by tunning the AC frequency, which indicates thattime heat current can be either delayed or elapsed with

respect to the AC electrical time-dependent signal.Now, we discuss the parity property of the response

functions g(ω) and m(ω) [thus G(ω) and M(ω)]. Wewrite g(ω) in the form

g(ω) = ℜeg(ω) + iℑmg(ω) , (50)

and express the real/imaginary part as

ℜeg(ω) =1

2[g(ω) + g∗(ω)] =

1

2

∫ ∞

−∞

dt eiωt [g(t) + g(−t)] ,

(51)

ℑmg(ω) =1

2i[g(ω)− g∗(ω)] =

1

2i

∫ ∞

−∞

dt eiωt [g(t)− g(−t)] .

(52)Here, we used the fact that the response function g(t)must be real to have a real expectation value for thecurrent IT (t). From Eq. (52), it is quite easy to showthat

ℑmg(−ω) =1

2i

∫ ∞

−∞

dt eiωt [g(−t)− g(t)] , (53)

which implies ℑmg(ω) = −ℑmg(−ω). Using a simi-lar line of reasoning, we can also prove that ℜeg(ω) =ℜeg(−ω). This argument also works for m(ω). It is wor-thy to notice that this parity property comes from thefact that we have included the contribution due to thetunnel Hamiltonian in our definition for the heat flowas shown in Eq. (35). Furthermore, this parity argu-ment goes beyond the simple site partitioning schemeexplained in Ref. [33].

B. Multi-orbital quantum capacitor

We now investigate a single reservoir carbon nanotubequantum dot as an example of a multi-orbital interact-ing conductor. We regard the nanotube quantum dot asa localized single particle level described by two quan-tum numbers, the orbital quantum number τ associatedto clockwise (τ = +1) and anti-clockwise (τ = −1) semi-classical orbits along the nanowire circumference (relatedwith the K-valley degeneracy in graphene) and the spindegree of freedom σ. In the presence of magnetic fieldalong the nanotube axis, the dot energy level splits inthe spin sector by the Zeeman field ∆Z and in the or-bital sector by an amount ∆orb that depends on the nan-otube radius.31 Besides, due to the nanowire curvature anon-negligible spin-orbit interaction is present, yielding aKramers splitting of magnitude ∆so. All together yieldsthe following carbon nanotube dot energy level:

ǫdστ = ǫd + σ∆Z + τ∆orb + στ∆so. (54)

In the following, we show results that correspond to therealistic parameters:31 ǫd = 0, ∆Z = 0.625Γ, ∆orb =5∆Z (orbital magnetic moments can be 5 − 20 times

Page 8: Time dependent heat flow in interacting quantum conductors

8

larger than its spin counterpart31), and ∆so = 0.5∆Z.The charging energy is 2EC = 2Γ which lies in the stronginteracting regime. We restrict ourselves to the low tem-perature regime kBT = 0.05Γ.

-4 -2 0 2 4 6 8 10E

F/Γ

2

4

6

CG

(Uni

ts o

f e2 /Γ

)

-4 -2 0 2 4 6 8 10E

F/Γ

-0.02

0

0.02

CM

(Uni

ts o

f e)

-4 -2 0 2 4 6 8 10E

F/Γ

0.25

0.5

RG

(h/

e2 )

-4 -2 0 2 4 6 8 10E

F/Γ

-4

0

RM

(h/e

Γ)

(a)

(c)

(b)

(d)

FIG. 6: (a) Electrical capacitance CG and (c) electrical re-laxation resistance RG versus the Fermi energy EF . (b) Elec-trothermal capacitance CM and (d) electrothermal relaxationresistance RM versus the Fermi energy EF . Nanotube pa-rameters: ǫd = 0, ∆Z = 0.625Γ, ∆orb = 5∆Z , ∆so = 0.5∆Z ,2EC = 2Γ, and kBT = 0.05Γ.

First, we show our results for the electrical and elec-trothermal RC parameters in Fig. 6. In the low tem-perature limit (kBT ≪ Γ) the RC parameters exhibituniversal values.12–14 We recall that in the low frequencyregime the RC parameters characterize the electrical andelectrothermal admittances [see Eqs. (46) and (48)]. CG

and RG are displayed in Fig. 6 (a) and (c) when EF isvaried. The electrical capacitance CG shows oscillationswhich peaks at the positions located roughly at

ǫd −∆Z −∆orb +∆so ≈ −3.4Γ, (55)

ǫd +∆Z −∆orb −∆so + 2EC ≈ −0.8Γ,

ǫd −∆Z +∆orb −∆so + 2(2EC) ≈ 6.2Γ,

ǫd +∆Z +∆orb +∆so + 3(2EC) ≈ 10Γ.

As expected, at each nanotube level position, RG takesthe value of h/2e2, while in the middle of two consecu-tive resonances (when two resonances contribute to RG)it diminishes to half of this value. Whereas RG and CG

display always positive values, the electrothermal capaci-tance CM and resistanceRM can become positive or neg-ative when EF is tuned. This is shown in Fig. 6 (b) and(d). Indeed, CM changes sign whenever the Fermi en-ergy matches with any of the nanotube resonances. Heatcurrent becomes delayed or elapsed with respect to theAC signal depending on the Fermi energy position. Theelectrothermal resistance RM modifies its sign not onlyat the points when EF coincides with the nanotube res-onances [see Eq. (55)] but also when the electron-holesymmetry point occurs, just at the midpoint between

two consecutive resonance points. The sign inversion inRM happens when the derivative of the carbon nanotubequantum dot density of states vanishes.26 Remarkably,both CM and RM diverge around the resonance points,behavior that is washed out by enhancing temperature.(Similar results were obtained for the weak interactinglimit, when EC ≪ Γ see Ref. [26] for details).

-12 -8 -4 0 4 8 12hω/Γ

0

0.2

0.4

ℜe

g(ω

) (U

nits

of

e2 /h)

-12 -8 -4 0 4 8 12hω/Γ

-0.3

0

0.3

ℑm

g(ω

) (U

nits

of

e2 /h)

(a) (b)

FIG. 7: (a) Real and (b) imaginary parts of g(ω) versus theAC frequency ω. Parameters: ǫd = 0, ∆Z = 0.625Γ, ∆orb =5∆Z , ∆so = 0.5∆Z , 2EC = 2Γ, EF = 0, and kBT = 0.05Γ.

We now discuss the case of arbitrary AC frequenciesand analyze the prefactor g(ω) of the electrical admit-tance in Fig. 7. The real and imaginary parts of g(ω)versus the AC frequency are depicted in Fig. 7 (a) and(b). As in the single orbital quantum capacitor, weobserve that ℑmg(ω) has odd parity with respect toω, while ℜeg(ω) is an even function of ω as a conse-quence of being response functions of a real perturbat-ing force. ℑmg(ω) accounts for the dissipative part ofthe electrical conduction with resonances roughly locatedat ±~ω ≈ |EF − ǫdστ |. For the HF approximation,these resonances coincide with the dot level positions thatare renormalized by interactions according to Eq. (10).These renormalized level positions (when EF = 0) are atǫdστ ≈ −1.4Γ, −0.8Γ, 6.1Γ, 8.0Γ leading to the observedresonances in g(ω). The resonant behavior of g(ω) re-flects the photon-assisted tunneling processes in whichtransport through the nanotube occurs by absorpting oremitting single photons. Furthermore, these resonancesare broaden mainly by Γ at very low temperatures. Theℜeg(ω) corresponds to the reactive part of the electricalconduction and has also a similar resonant structure.We turn into the analysis of the electrothermal admit-

tance M(ω) which is given in Eq. (42). As before, it isconvenient to examine m(ω) defined in Eq. (43) and wethus plot it as a function of the AC frequency in Fig. 8.First, we observem(ω) displays the same parity propertyas g(ω) which can be explained in the same way. Fur-thermore, the real and imaginary parts of m(ω) are alsocharacterized by features located at ±~ω ≈ |EF − ǫdστ |.

Page 9: Time dependent heat flow in interacting quantum conductors

9

-12 -8 -4 0 4 8 12hω/Γ

-0.1

0

0.1

0.2

0.3

0.4

0.5ℜ

e m

(ω) (

Uni

ts o

f e/

h)

-12 -8 -4 0 4 8 12hω/Γ

-0.3

0

0.3

ℑm

m(ω

) (U

nits

of

e/h)

(a) (b)

FIG. 8: (a) Real and (b) imaginary parts of m(ω) versus theAC frequency ω. Parameters: ǫd = 0, ∆Z = 0.625Γ, ∆orb =5∆Z , ∆so = 0.5∆Z , 2EC = 2Γ, EF = 0, and kBT = 0.05Γ.

Remarkably, the observed peaks indicate single photonabsorption and emission processes for the transport ofheat. It is worthy to emphasize that our approach isable to capture such photon-assisted processes in con-trast with other calculations restricted to AC frequenciessmaller or similar to the tunnel coupling Γ: the othercalculations assume the AC frequency is small such thatonly a second order expansion of m(ω) in powers of ω(characterized by the pairs CM, RM) can be justified.However, since we do not have such a restriction for theapplied AC frequency, our results clearly exhibit photon-assisted processes for the heat transport.Remarkably, m(ω) (either its real or imaginary part)

takes positive or negative values depending on the ACfrequency regime. We observe ℜem(ω) < 0 for low andmoderate ω, whereas for large AC frequencies its sign is

reversed. We stress the importance of such result for thefunctionality of quantum circuits, in which one could ma-nipulate the sign of the heat flow spectrum by properlytunning the AC frequency.

IV. CONCLUSIONS

In closing, we have investigated the heat current spec-trum in the linear response regime for an interacting con-ductor coupled to a single reservoir and modulated byan electrical AC signal. Our results, valid for arbitraryAC frequencies, show that the heat current expressionfor the reservoir needs to consider the heat stored or re-laxed at the barrier. We illustrate our findings with twoprototype of interacting conductors, namely, a single or-bital quantum dot and a multi-orbital conductor–a car-bon nanotube quantum dot coupled to a single reservoir.We deal with the strong interacting limit where Coulombblockade phenomena applies. We highlight that (i) elec-trical and heat transport displays photon-assisted trans-port features, and (ii) the electrothermal admittance canbe positive or negative and the sign can be chosen by ad-justing properly the AC frequency. This is an importantissue for engineering nanoelectronic circuits with optimalheat dissipation performances.

V. ACKNOWLEDGEMENTS

We acknowledge J. Splettstoesser and D. Sanchez forcritical reading and useful discussions. Work supportedby MINECO Grant No. FIS2011-23526. We acknowledgefinancial support from the German Ministry of Innova-tion NRW.

1 S. Jezouin, F. D. Parmentier, A. Anthore, U. Gennser, A.Cavanna, Y. Jin, and F. Pierre, Science 342, 601 (2013).C. W. Chang, D. Okawa, A. Majumdar, A. Zettl, Science,314, 1121 (2006). Y. Kim, W. Jeong, K. Kim, W. Lee, andP. Reddy, Nat. Nanotechnol. 9, 881 (2014).

2 R. Maynard, and E. Akkermans, Phys. Rev. B 32, 5440(1985). L. D. Hicks and M. S. Dresselhaus, Phys. Rev. B47, 16631 (1993). L. E., Bell, Science 321, 1457 (2008). Y.Dubi and M. Di Ventra, Rev. Mod. Phys. 83, 131 (2011).P. Rodgers, Nature Nanotech. 3, 76 (2008). J.-P. Brantut,C. Grenier, J. Meineke, D. Stadler, S. Krinner, C. Kollath,T. Esslinger, and A. Georges, Science 342, 713 (2013). Y.Kim, W. Jeong, K. Kim, W. Lee, and P. Reddy, NatureNanotech. 10, 1038 (2014).

3 M. Buttiker, A. Pretre, and H. Thomas, Phys. Lett. A 180,364 (1993).

4 M. Buttiker, A. Pretre, and H. Thomas, Phys. Rev. Lett.70, 4114 (1993).

5 Y. Dovzhenko, J. Stehlik, K. D. Petersson, J. R. Petta, H.

Lu, and A. C. Gossard Phys. Rev. B 84, 161302(R) (2011).G. Cao, H-O. Li, T. Tu, L. Wang, C. Zhou, M. Xiao, G-C.Guo, H-W. Jiang, and G-P. Guo, Nat. Commun. 4, 1401(2013).

6 W. Lu, J. Zhongqing, L. Pfeiffer, K. W. West and A. J.Rimberg, Nature 423, 422 (2003).

7 S. Gustavsson, R. Leturcq, T. Ihn, K. Ensslin and A. C.Gossard, J. Appl. Phys. 105, 122401 (2009).

8 T. Jullien, P. Roulleau, B. Roche, A. Cavanna, Y. Jin, andD. C. Glattli, Nature 514 603, (2014).

9 J. Gabelli, G. Feve, J.-M. Berroir, B. Placais, A. Cavanna,B. Etienne, Y. Jin, and D.C. Glattli, Science 313, 49(2006).

10 J. Gabelli, G. Feve, T. Kontos, J.-M. Berroir, B. Placais,D.C. Glattli, B. Etienne, Y. Jin, and M. Buttiker, Phys.Rev. Lett. 98, 166806 (2007).

11 M. R.Delbecq, V. Schmitt, F. D.Parmentier, N. Roch, J.J.Viennot, G. Feve, B. Huard, C. Mora, A. Cottet, and T.Kontos, Phys. Rev. Lett. 107, 256804 (2011).

Page 10: Time dependent heat flow in interacting quantum conductors

10

12 Y. Fu and S.C. Dudley, Phys. Rev. Lett. 70, 65 (1993).13 T. Christen and M. Buttiker Phys. Rev. Lett. 77, 143

(1996).14 T. Christen and M. Buttiker Phys. Rev. B 53, 2064 (1996).15 A. Pretre, H. Thomas, and M. Buttiker Phys. Rev. B 54,

8130 (1996).16 M. Buttiker and T. Christien in Mesoscopic Electron

Transport, edited by L. Kouwenhoven, G. Schon, andL. Sohn, NATO ASI Series E, Kluwer Ac. Publ., Dordrecht(1996).

17 S.E. Nigg, R. Lopez, and M. Buttiker, Phys. Rev. Lett.97, 206804 (2006).

18 C. Mora and K. Le Hur, Nature Physics 6 697 (2010).19 M. Lee, R. Lopez, Mahn-Soo Choi, T. Jonckheere, and T.

Martin Phys. Rev. B 83, 201304(R) (2011).20 J. Gabelli, G. Feve, J.-M. Berroir, and B. Placais, Rep.

Prog. Phys. 75, 126504 (2012).21 M. Hashisaka, K. Washio, H. Kamata, K. Muraki, and

T. Fujisawa, Phys. Rev. B 85, 155424 (2012).22 S.J. Chorley, J. Wabnig, Z.V. Penfold-Fitch, K.D. Peters-

son, J. Frake, C.G. Smith, and M.R. Buitelaar, Phys. Rev.Lett. 108, 036802 (2012).

23 J. Basset, H. Bouchiat, and R. Deblock, Phys. Rev. B 85,085435 (2012).

24 T. Frey, P.J. Leek, M. Beck, J. Faist, A. Wallraff, K. En-sslin, T. Ihn, and M. Buttiker, Phys. Rev. B 86, 115303

(2012).25 L. Arrachea, M. Moskalets, and L. Martin-Moreno, Phys.

Rev. B 75, 245420 (2007). M. Moskalets and M. Buttiker,Phys. Rev. B 80, 081302(R) (2009).

26 J. S. Lim, R. Lopez, and D. Sanchez Phys. Rev. B 88,

201304(R) (2013).27 M. F. Ludovico, J-S. Lim, M. Moskalets, L. Arrachea, and

D. Sanchez Phys. Rev. B 89, 161306(R) (2014).28 F. Giazotto and M. J. Martınez-Perez, Nature (London)

492, 401 (2012). B. Sothmann, R. Sanchez, A. N. Jordan,and M. Buttiker, New J. Phys. 15, 095021 (2013). J. V.Koski, T. Sagawa, O.-P. Saira, Y. Yoon, A. Kutvonen, P.Solinas, M. Mttnen, T. Ala-Nissila, and J. P. Pekola, Nat.Phys. 9, 644 (2013).

29 A.P. Jauho, N.S. Wingreen, and Y. Meir, Phys. Rev. B 50,5528 (1994).

30 M. Bockrath, D. H. Cobden, P. L. McEuen, N. G. Chopra,A. Zettl, A. Thess, and R. E. Smalley, Science 275, 1922(1997).

31 E. D. Minot, Y. Yaish, V. Sazonova, and P. L. McEuen,Nature (London) 428, 536 (2004).

32 P. Brune, C. Bruder, and H. Schoeller, Phys. Rev. B 56,4730 (1997).

33 Lian-Ao Wu, Dvira Segal, J. Phys. A: Math. and Theo.42, 025302 (2009).