Nonequilibrium plasmons and transport properties of a double-junction quantum wire

We study theoretically the current-voltage characteristics, shot noise, and full counting statistics of a quantum wire double barrier structure. We model each wire segment by a spinless Luttinger liquid. Within the sequential tunneling approach, we describe the system’s dynamics using a master eq...

Full description

Saved in:
Bibliographic Details
Published in:Физика низких температур
Date:2006
Main Authors: Kim, J.U., Mahn-Soo Choi, Krive, I.V., Kinaret, J.M.
Format: Article
Language:English
Published: Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України 2006
Subjects:
Online Access:https://nasplib.isofts.kiev.ua/handle/123456789/120865
Tags: Add Tag
No Tags, Be the first to tag this record!
Journal Title:Digital Library of Periodicals of National Academy of Sciences of Ukraine
Cite this:Nonequilibrium plasmons and transport properties of a double-junction quantum wire / J.U. Kim, Mahn-Soo Choi, I.V. Krive, J.M. Kinaret // Физика низких температур. — 2006. — Т. 32, № 12. — С. 1522–1544. — Бібліогр.: 78 назв. — англ.

Institution

Digital Library of Periodicals of National Academy of Sciences of Ukraine
id nasplib_isofts_kiev_ua-123456789-120865
record_format dspace
spelling Kim, J.U.
Mahn-Soo Choi
Krive, I.V.
Kinaret, J.M.
2017-06-13T06:01:02Z
2017-06-13T06:01:02Z
2006
Nonequilibrium plasmons and transport properties of a double-junction quantum wire / J.U. Kim, Mahn-Soo Choi, I.V. Krive, J.M. Kinaret // Физика низких температур. — 2006. — Т. 32, № 12. — С. 1522–1544. — Бібліогр.: 78 назв. — англ.
0132-6414
PACS: 71.10.Pm, 72.70.+m, 73.23.Hk, 73.63.–b
https://nasplib.isofts.kiev.ua/handle/123456789/120865
We study theoretically the current-voltage characteristics, shot noise, and full counting statistics of a quantum wire double barrier structure. We model each wire segment by a spinless Luttinger liquid. Within the sequential tunneling approach, we describe the system’s dynamics using a master equation. We show that at finite bias the nonequilibrium distribution of plasmons in the central wire segment leads to increased average current, enhanced shot noise, and full counting statistics corresponding to a super-Poissonian process. These effects are particularly pronounced in the strong interaction regime, while in the noninteracting case we recover results obtained earlier using detailed balance arguments.
This work has been supported by the Swedish Foundation for Strategic Research through the CARAMEL consortium, STINT, the SKORE-A program, the eSSC at Postech, and the SK-Fund. J.U. Kim acknowledges partial financial support from Stiftelsen Fru Mary von Sydows, fodd Wijk, donationsfond. I.V. Krive gratefully acknowledges the hospitality of the Department of Applied Physies at Chalmers University of Technology.
en
Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України
Физика низких температур
Низкоразмерные и неупорядоченные системы
Nonequilibrium plasmons and transport properties of a double-junction quantum wire
Article
published earlier
institution Digital Library of Periodicals of National Academy of Sciences of Ukraine
collection DSpace DC
title Nonequilibrium plasmons and transport properties of a double-junction quantum wire
spellingShingle Nonequilibrium plasmons and transport properties of a double-junction quantum wire
Kim, J.U.
Mahn-Soo Choi
Krive, I.V.
Kinaret, J.M.
Низкоразмерные и неупорядоченные системы
title_short Nonequilibrium plasmons and transport properties of a double-junction quantum wire
title_full Nonequilibrium plasmons and transport properties of a double-junction quantum wire
title_fullStr Nonequilibrium plasmons and transport properties of a double-junction quantum wire
title_full_unstemmed Nonequilibrium plasmons and transport properties of a double-junction quantum wire
title_sort nonequilibrium plasmons and transport properties of a double-junction quantum wire
author Kim, J.U.
Mahn-Soo Choi
Krive, I.V.
Kinaret, J.M.
author_facet Kim, J.U.
Mahn-Soo Choi
Krive, I.V.
Kinaret, J.M.
topic Низкоразмерные и неупорядоченные системы
topic_facet Низкоразмерные и неупорядоченные системы
publishDate 2006
language English
container_title Физика низких температур
publisher Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України
format Article
description We study theoretically the current-voltage characteristics, shot noise, and full counting statistics of a quantum wire double barrier structure. We model each wire segment by a spinless Luttinger liquid. Within the sequential tunneling approach, we describe the system’s dynamics using a master equation. We show that at finite bias the nonequilibrium distribution of plasmons in the central wire segment leads to increased average current, enhanced shot noise, and full counting statistics corresponding to a super-Poissonian process. These effects are particularly pronounced in the strong interaction regime, while in the noninteracting case we recover results obtained earlier using detailed balance arguments.
issn 0132-6414
url https://nasplib.isofts.kiev.ua/handle/123456789/120865
citation_txt Nonequilibrium plasmons and transport properties of a double-junction quantum wire / J.U. Kim, Mahn-Soo Choi, I.V. Krive, J.M. Kinaret // Физика низких температур. — 2006. — Т. 32, № 12. — С. 1522–1544. — Бібліогр.: 78 назв. — англ.
work_keys_str_mv AT kimju nonequilibriumplasmonsandtransportpropertiesofadoublejunctionquantumwire
AT mahnsoochoi nonequilibriumplasmonsandtransportpropertiesofadoublejunctionquantumwire
AT kriveiv nonequilibriumplasmonsandtransportpropertiesofadoublejunctionquantumwire
AT kinaretjm nonequilibriumplasmonsandtransportpropertiesofadoublejunctionquantumwire
first_indexed 2025-11-26T23:38:51Z
last_indexed 2025-11-26T23:38:51Z
_version_ 1850781668269359104
fulltext Fizika Nizkikh Temperatur, 2006, v. 32, No. 12, p. 1522–1544 Nonequilibrium plasmons and transport properties of a double-junction quantum wire Jaeuk U. Kim1, Mahn-Soo Choi2, Ilya V. Krive3,4, and Jari M. Kinaret3 1Department of Physics, G�teborg University, SE-412 96 G�teborg, Sweden 2Department of Physics, Korea University, Seoul 136-701, Korea 3Department of Applied Physics, Chalmers University of Technology, SE-412 96 G�teborg, Sweden 4B. Verkin Institute for Low Temperature Physics and Engineering, 47 Lenin Ave., Kharkov 61103, Ukraine Received January 3, 2006, revised January 22, 2006 We study theoretically the current-voltage characteristics, shot noise, and full counting statis- tics of a quantum wire double barrier structure. We model each wire segment by a spinless Luttinger liquid. Within the sequential tunneling approach, we describe the system’s dynamics using a master equation. We show that at finite bias the nonequilibrium distribution of plasmons in the central wire segment leads to increased average current, enhanced shot noise, and full count- ing statistics corresponding to a super-Poissonian process. These effects are particularly pro- nounced in the strong interaction regime, while in the noninteracting case we recover results ob- tained earlier using detailed balance arguments. PACS: 71.10.Pm, 72.70.+m, 73.23.Hk, 73.63.–b Keywords: quantum wire, Luttinger liquid, shot noise, full counting statistics. 1. Introduction The recent discovery of novel one-dimensional (1D) conductors with non-Fermi liquid behaviors has inspired extensive research activities both in theory and experiment. The generic behavior of electrons in 1D conductors is well described by the Luttinger liquid (LL) theory, a generalization of the Tomo- naga–Luttinger (TL) model [1–3]. Luttinger liquids are clearly distinguished from Fermi liquids by many interesting characteristics. The most important exam- ples among others would be (a) the bosonic nature of the elementary excitations (i.e., the collective density fluctuations) [4], (b) the power-law behavior of cor- relations with interaction dependent exponents [5–8], and (c) the spin-charge separation [9]. All these cha- racteristics cast direct impacts on transport properties of an 1D system of interacting electrons, including the shot noise and the full counting statistics as we discuss in this paper. In this paper we will consider a particular struc- ture, namely, the single-electron transistor (SET), made of 1D quantum wires (QWs). A conventional SET with noninteracting electrodes itself is one of the most extensively studied devices in recent years in various contexts [10]. The SET of interacting 1D QWs has attracted renewed interest as a tunable de- vice to test the LL theory and thereby improve the un- derstanding of the 1D interacting systems. In parti- cular, the recent experimental reports on the temperature dependence of the resonance level width in a SET of semiconductor QWs [11] and on the tem- perature dependence in a SWNT SET [12] stirred a controversy, with the former consistent with the con- ventional sequential tunneling picture [13,14] and the latter supporting the correlated sequential tunneling picture [12,15]. The issue motivated the more recent theoretical works based on a dynamical quantum Monte Carlo method [16] and on a function renor- malization group method [17–20], and still remain controversial. Another interesting issue on the SET structure of 1D QWs is the effects of the plasmon modes in the central QW [21–24]. It was suggested that the plasmon excitations in the central QW leads to a © Jaeuk U. Kim, Mahn-Soo Choi, Ilya V. Krive, and Jari M. Kinaret, 2006 power-law behavior of the differential conductance with sharp peaks at resonances to the plasmon modes [21]. The plasmon excitations were also ascribed to the shot noise characteristics reflecting the strong LL correlations in the device [22]. Note that in these both works, they assumed fast relaxation of plasmon exci- tations. We hereafter refer to this approach as «equi- librium plasmons». However, it is not clear, espe- cially, in the presence of strong bias (eV k TB�� ) how well the assumption of equilibrium plasmons can be justified. This question is important since the plasmon modes excited by electron tunneling events can influ- ence the subsequent tunnelings of electrons and hence act as an additional source of fluctuations in current. In other words, while the average conductance may not be affected significantly, the effects of the «nonequilibrium plasmons» can be substantial on shot noise characteristics [25–31]. Indeed, a recent work suggest that the nonlinear distribution of the plasmon excitations itself is of considerable interest [23] and it can affect the transport properties through the de- vices, especially the shot noise, significantly [23,24]. Moreover, even if the precise mechanism of plasmon relaxation in nanoscale structures is not well known and it is difficult to estimate its rate, recent computer simulations on carbon nanotubes indicate that the plasmon life time could be of the order of a picosec- ond, much longer than those in three-dimensional structures [32]. It is therefore valuable to investigate systemati- cally the effects of nonequilibrium distribution of plasmon excitations with finite relaxation rate on the transport properties through a SET of 1D QWs. In this work we will focus on the shot noise (SN) charac- teristics and full counting statistics (FCS), which are more sensitive to the nonequilibrium plasmon excita- tions than average current-voltage characteristics. We found that the SN characteristics and the FCS both indicate that the fluctuations in the current through the devices is highly super-Poissonian. We ascribe this effects to the additional conduction channels via exci- tation of the plasmon modes. To demonstrate this rig- orously, we present both analytic expression of simpli- fied approximate models and numerical results for the full model system. Interestingly, the enhancement of the noise and hence the super-Poissonian character of the FCS is more severe in the strong interaction limit. Further more, the sensitive dependence of the super- Poissonian shot noise on the nonequilibrium plasmon excitations and their relaxation may provide a useful tool to investigate the plasmon relaxation phenomena in 1D QWs. The paper is organized as following: In Sec. 2 we introduce our model for a 1D QW SET and briefly ex- amine the basic properties of the tunneling rates within the golden rule approximation. We also intro- duce the master equation approach to be used through- out the paper, and discuss the possible experimental realizations. Before going to the main parts of the pa- per, in Sec. 3, we first review the results of the previ- ous work [23], namely, the nonequilibrium distribu- tion of the plasmons in the central QW in the limit of vanishing plasmon relaxation. This property will be useful to understand the results in the subsequent sec- tions. We then proceed to investigate the consequence of the nonequilibrium plasmons in the context of the transport properties. We first consider average current in Sec. 4, and then discuss shot noise in Sec. 5. Finally, we investigate full counting statistics in Sec. 6. Section 7 concludes the paper. 2. Formalism The electric transport of a double barrier structure in the (incoherent) sequential tunneling regime can be described by the master equation [14,33,34] � � � � � � � � � � t P N n t N n N n P N n t N( , { }, ) [ ( , { } , { }) ( , { }, ) ( , {� � n N n P N n t nN � � �� �� } , { }) ( , { }, )] { } , (1) where P N n t( , { }, ) is the probability that at time t there are N (excess) electrons and { } ( , , ..., , ...)n n n nm� 1 2 plasmon excitations (i.e., collective charge excitations), that is, nm plasmons in the mode m on the quantum dot. The transitions occur via single-electron tunneling through the left (L) or right (R) junctions (see Fig. 1). The total transition rates � in master equation (1) are sums of the two transition rates �L and �R where �L/R N n N n( , { } , { })� � � is the transition rate from a quantum state ( , { })N n� � to another quantum state ( , { })N n via electron tunneling through L/R-junction. Master equation (1) implies that, with known transi- tion rates, the occupation probabilities P N n t( , { }, ) can be obtained by solving a set of linear first order differen- tial equations with the probability conservation P N n t N n ( , { }, ) ,{ } � � 1. Nonequilibrium plasmons and transport properties of a double-junction quantum wire Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 1523 In the long time limit the system converges to a steady-state with probability distribution lim ( , { }, ) t P N n t �� � P N nst ( , { }), irrespective of the initial preparation of the system. To calculate the transition rates, we start from the Hamiltonian of the system. The reservoir temperature is assumed zero (T � 0), unless it is stated explicitly. 2.1. Model and Hamiltonian The system we consider is a 1D quantum wire SET. Schematic description of the system is that a finite wire segment, which we call a quantum dot, is weakly coupled to two long wires as depicted in Fig. 1. The chemical potential of the quantum dot is controlled by the gate voltage (VG) via a capacitively coupled gate electrode. In the low-energy regime, physical proper- ties of the metallic conductors are well described by linearized dispersion relations near the Fermi points, which allows us to adopt the Tomonaga–Luttinger Hamiltonian for each wire segment. We model the system with two semi-infinite LL leads and a finite LL for the central segment. The leads are adiabatically connected to reservoirs which keep them in internal equilibria. The chemical potentials of the leads are controlled by source-drain voltage (V), and the wires are weakly coupled so that the single-electron tunnel- ing is the dominant charge transport mechanism, i.e., we are interested in the sequential tunneling regime. Rigorously speaking, the voltage drop between the two leads (V) deviates from the voltage drop between the left and right reservoirs (sayU) if electron trans- port is activated [35,36]. However, as long as the tun- neling amplitudes through the junctions (barriers) are weak so that the Fermi golden rule approach is appro- priate, we estimate V U� . The total Hamiltonian of the system is then given by the sum of the bosonized LL Hamiltonian � � � �H H H HL D R0 � accounting for three isolated wire segments labeled by � � ( , , )L D R , and the tunnel- ing Hamiltonian �HT accounting for single-electron hops through the junctions L and R at XL and XR, re- spectively: � � � .H H HT� 0 (2) Using standard bosonization technique, the Hamil- tonian describing each wire segment can be expressed in terms of creation and annihilation operators for col- lective excitations ( �†b and �b) [4,37]. For the semi-in- finite leads, it reads � � � , † ,H mb b L R M m m m� � �� � � � � ��� � � � 1 1 for = , , (3) where the index labels the M transport sectors of the conductor and m the wave-like collective excitations on each transport sector. The effects of the Coulomb inter- action in 1D wire are characterized by the Luttinger pa- rameter g�: g � 1 for noninteracting Fermi gas and 0 1� �g for the repulsive interactions (g �� 1 in the strong interaction limit). Accordingly, the velocities of the collective excitations are also renormalized as v v /gF� �� . The energy of an elementary excitation in sector is given by � �� �� �v /L where L is the length of the wire and � the Planck constant. For instance, if the wire has a single transport channel (usually referred to as spinless electrons), e.g., a wire with one transport channel under a strong magnetic field, the system’s dy- namics is determined by collective charge excitations (plasmons) alone ( �� and M � 1). If, however, the spin degrees of freedom survive, the wire has two trans- port sectors (M � 2); plasmons ( �� ) and spin-waves ( �� ) [37]. If the system has two transport channels with electrons carrying spin (M � 4), as is the case with SWNTs, the transport sectors are total-charge-plasmons ( �� ), relative-charge-plasmons ( �� � ), total-spin- waves ( ��� , and relative-spin-waves ( �� � ) [38,39]. For the short central segment, the zero-mode need to be accounted for as well, which yields � � � , † ,H mb bD m m m� � � � � �� � � � 1 � � � �� �� � � � �� � �2 2 2 2 Mg N N Mg N EG r( � ) � . (4) In the second line of Eq. (4), which represents the zero-mode energy of the quantum dot, the operator �N� measures the ground-state charge, i.e., with no excita- tions, in the -sector. The zero-mode energy systemati- cally incorporates Coulomb interaction in terms of the 1524 Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 Jaeuk U. Kim, Mahn-Soo Choi, Ilya V. Krive, and Jari M. Kinaret 3 quantum wires connected to reservoirs Fig. 1. Model system. Two long wires are adiabatically connected to reservoirs and a short wire is weakly coupled to the two leads. Tunneling resistances at junction points XL and XR are RL/R, and the junction capacitances are considered equal C CL R� . Quantum dot is capacitively coupled to the gate electrode. Luttinger parameter g� in the QD. To refer to the zero-mode energy later in this paper, we define the «charging energy» EC as the minimum energy cost to add an excess electron to the QD in the off-Coulomb blockade regime, i.e., E E E /gC D D N /G � � ��[ ( , ) ( , )] .2 0 1 0 1 2 �� � (5) Note that this is twice the conventional definition, and includes the effects of finite level spacing. Hence, the charging energy is smallest in the noninteracting limit (g� � 1) and becomes the governing energy scale in the strong interaction limit (g� �� 1). The origin of the charging energy in conventional quantum dots is the long range nature of the Coulomb interaction. In the theory of Luttinger liquid, the long range inter- action can easily be incorporated microscopically through the interaction strength g� . For the effect of the finite-range interaction across a tunneling junc- tion, for instance see Refs. 40, 41. Note that charge and spin are decoupled in Lut- tinger liquids, which implies the electric forces affect the (total) charge sector only; due to intrinsic e e� in- teractions, g� � 1 but g� �� � 1, and the gate voltage shifts the band bottom of the (total) charge sector as seen by the dimensionless gate voltage parameter NG in Eq. (4). As will be shown shortly, the transport properties of the L/R-leads are determined by the LL interaction parameter g� and the number of the trans- port sector M. In this work, we consider each wire segment has the same interaction strength for the (to- tal) charge sector, g g L� �� ( ) g gR D � � ( ) ( )� . Accord- ingly, the energy scales in the quantum are written by � � �� � �� �( )D F Dv /g L� and � � �� �0 � �� �v /LF D. We consider the ground-state energy in the QD is the same as those in the leads, by choosing the refer- ence energy Er in Eq. (4) equals the minimum value of the zero-mode energy, E N Mg N Mg M Mr G G� � �� � � � � � � � min , ( ) ( )� � �� � � � 2 2 0 2 1 2 1 2 , (6) where min( , )x x� denotes the smaller of x and �x , and the gate charge NG is in the range N MG � [ , ]0 . The zero-mode energy in the QD, E Mg N N M N EG r0 2 0 2 2 2 � � � � �� �� � � � � ( ) (7) yields degenerate ground states for N � 0 and N � 1 excess electrons when N M g /G � � [( ) ]1 1 22 � . Here we replaced N� by the number of the total excess electrons N since N N� � , and N� are all either even or odd integers, simultaneously; in the case of the SWNTs with N N i s i s� � , , excess electrons, where i � 1 2, is the channel index and s � � �, is the spin index of conduction electrons (M = 4), N N� � , N N Ni i i� � � �( ), , , N N Ns s s�� � ��( ), ,1 2 , N N Ni i i i�� � � �� ( ) ( ), ,1 . From now on we consider only one spin-polarized (or spinless) channel unless otherwise stated — our focus is on the role of Coulomb interactions, and the additional channels only lead to more complicated ex- citation spectra without any qualitative change in the physics we address below. A physical realization of the single-channel case may be obtained, e.g., by ex- posing the quantum wire to a large magnetic field. For the system with high tunneling barriers, the electron transport is determined by the bare electron hops at the tunneling barriers. The tunneling events in the DB structure are described by the Hamiltonian � [ � ( ) � ( ) .]† , H t X XT D L R � � � � � � � � � � h.c , (8) where � ( )†� � �X and � ( )�� �X are the electron creation and annihilation operators at the edges of the wires near the junctions at XL and XR. As mentioned ear- lier, the electron field operators �� and � †� are related to the plasmon creation and annihilation operators �b and �†b by the standard bosonization formulae. Differ- ent boundary conditions yield different relations be- tween electron field operators and plasmon operators. Exact solutions for the periodic boundary condition have been known for decades [3,37] but the open boundary conditions which are apt for our system of consideration has been investigated only recently (see, for example, Refs. 38, 42–44). The dc bias voltage V V VL R� between L and R leads is incorporated into the phase factor of the tunneling matrix elements t t ieV t/� � �� �� | | exp( ) by a time-dependent unitary transformation [10]. Here VL/R � VC/CR/L is voltage drop across the L/R- junction where C C C / C CL R L R� ( ) is the effective total capacitance of the double junction, and the bare tunneling matrix amplitudes | |tL/R are assumed to be energy independent. Experimentally, the tunneling matrix amplitude is sensitive to the junction properties while the capacitance is not. For simplicity, the capaci- tances are thus assumed to be symmetric C CL R� throughout this work. By junction asymmetry we mean the asymmetry in (bare) squared tunneling amplitudes | |tL/R 2. The parameter R t / tL R� | | | |2 2 is used to de- scribe junction asymmetry; R � 1 for symmetric junc- tions and R �� 1 for highly asymmetric junctions. It is known that, at low energy scales in the quantum wires with the electron density away from Nonequilibrium plasmons and transport properties of a double-junction quantum wire Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 1525 half-filling, the processes of backward and Umklapp scattering, whose processes generate momentum trans- fer across the Fermi sea (� 2kF), can be safely ignored in the middle of ideal 1D conductors [37], including armchair SWNTs [38]. The Hamiltonian (2) does not include the backward and Umklapp scattering (except at the tunneling barriers) and therefore it is valid away from half-electron-filling. We find that, in the regime where electron spin does not play a role, the addition of a transport chan- nel does not change essential physics present in a sin- gle transport channel. Therefore, we primarily focus attention to a QW of single transport channel with spinless electrons and will comment on the effects due to multiple channel generalization, if needed. 2.2. Electron transition rates The occupation probability of the quantum states in the SET system changes via electron tunneling events across L/R-junctions. In the single-electron tunneling regime, the bare tunneling amplitudes | |tL/R are small compared to the characteristic energy scales of the system and the electron tunneling is the source of small perturbation of three isolated LLs. In this regime, we calculate transition rates �L/R be- tween eigenstates of the unperturbed Hamiltonian H0 to the lowest nonvanishing order in the tunneling am- plitudes | |tL/R . In this golden rule approximation, we integrate out lead degrees of freedom since the leads are in internal equilibria, and the transition rates are given as a function of the state variables and the ener- gies of the QD only [21,23], �L/R N n N n( , { } , { })� � � � � �2 2� � � � | | ( ) ({ }, { })t W n nL/R L/R D . (9) In Eq. (9)WL/R is the change in the Gibbs free energy associated with the tunneling across the L/R-junction, W E N n E N n N N eVL/R D D L/R� � � � ��( , { }) ( , { }) ( )� . (10) Here E N n N n H N nD D( , { }) , { }| � | , { }� ! is the energy of the eigenstate | , { }N n ! of the dot and L/R correspond to � / . For the QD with only one transport channel with spinless electrons only, E N n mn N N g ED p m m G D r( , { }) ( ) ,� �" # $ $ % & ' ' � � � �� 1 2 2 (11) where � ��p � accounting that we consider only charge plasmons and n n b b nm m m m m� !| � � |† is the num- ber of plasmons in the mode m. Note that the excita- tion spectrum (11) in the QD consists of charge excitations which are formed by changing electron number N in the zero-mode, and plasmon excitations which are neutral excitations. In this work, we sharply distinguish those two different excitations. The function �( )x in Eq. (9) is responsible for the plasmon excitations on the leads, and given by (see, e.g., Ref. 14) � � � �( ) ( , ) ( , )†� ! � � � (1 2 0 � � � � �dt X X ti te � � � � � � � � � � � � � � � � � 1 2 1 2 1 2 2 1 1 � � � ) * )� � � � �v g v i F / g F + � ( ) , , , , , , 2 2 1 e �� * / �( ) , (12) where ) � 1/k TB is the inverse temperature in the leads, + is a short wavelength cutoff, and �( )z is the gamma function. The exponent * � � ( )g /M1 1 is a characteristic power law exponent indicating interac- tion strength of the leads with M transport sectors (hence, in our case, M � 1). At g � 1 (noninteracting case), the exponent * � 0 and it grows as g - 0 (strong interaction). The decrease of the exponent * with increasing M implies that the effective interac- tion decreases due to multi-channel effect, and the Luttinger liquid eventually crosses over to a Fermi liquid in the many transport channel limit [45,46]. For the noninteracting electron gas, the spectral density � �( ) is the TDOS multiplied by the Fermi–Dirac distribution function fFD( )� � [ exp( )]1 1 )� ; � �( ) � � f / vFD F( ) ( )� �� for g � 1. At zero temperature, � �( ) is proportional to a power of energy, lim ( ) ( ) (| | ) ( )T Fv / � � � 0 1 1 � � � � � � * � . � � � , (13) where � � +� �v / gF / g1 1( ) is a high energy cut-off. At zero temperature � �( ) is the TDOS for the negative energies and zero otherwise (as it should be), imposed by the unit step function .( )�� . The function � D in Eq. (9) accounts for the plasmon transition amplitudes in the QD, and is given by � /D N N Dn n N n X N n({ }, { }) | , { }| � ( )| , { } |, †� � � � ! � �1 2� � � � !� /N N DN n X N n, | , { }| � ( )| , { } |1 2� � , (14) where we used that the zero-mode overlap is unity for N N� � 0 1 and vanishes otherwise. The overlap integrals between plasmon modes are, although straightforward, quite tedious to calculate, and we refer to Appendix A for the details. The resulting overlap of the plasmon states can be written as a function of the mode occupations nm , 1526 Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 Jaeuk U. Kim, Mahn-Soo Choi, Ilya V. Krive, and Jari M. Kinaret | { }| � ( )| { } | | { }| � ( )| { } |† � ! � � ! �n X n n X nD D� �� � 2 2 � � � �� � � �� � 1 L L n n D D D� 1 � + ({ }, { }) (15) with 1({ }, { }) ! ! | | ( ) ( ) n n gm n n L m m m n nm nm � � � � �� � � �� � � � �2 � � 1 1 m m mn n gm( ) | | , � � � � �� � � �� " # $ % & ' 1 2 (16) where n n nm m m ( ) ( , )� � �min and n n nm m m ( ) ( , )� � �max , * D Dg� � ( )1 1 for the QD with one transport sector, and L xa b ( ) are Laguerre polynomials. Additional tran- sport sectors would appear as multiplicative factors of the same form as 1({ }, { })n n� and result in a reduction of the exponent * D. Notice that in the low energy scale only the first few occupations nm and n m� in the product differ from zero, participating to the transi- tion rate (9) with nontrivial contributions. 2.3. Plasmon relaxation process in the quantum dot In general, plasmons on the dot are excited by tun- neling events and have a highly nonequilibrium distri- bution. The coupling of the system to the environment such as external circuit or background charge in the substrate leads to relaxation towards the equilibrium. While the precise form of the relaxation rate, �p , de- pends on the details of the relaxation mechanism, the physical properties of our concern do not depend on the details. Here we take a phenomenological model where the plasmons are coupled to a bath of harmonic oscillators by H g b b amn m n m n plasmon bath h. c. � �� ( )† , � � � . (17) In Eq. (17) a� and a� † are bosonic operators describ- ing the oscillator bath and gmn � is the coupling con- stants. We will assume an Ohmic form of the bath spectral density function J gmn mn p( ) | | ( ) ,3 / 3 3 � 3� � �� � �� 2 0� (18) where 3� is the frequency of the oscillator cor- responding to a� , � p is a dimensionless constant characterizing the bath spectral density, and �0 1 � � � 2 2 2v L t tF D L R(| | | | ) is the natural time scale of the system. Within the rotating-wave approximation, the plasmon transition rate due to the harmonic oscil- lator bath is given by � �p p p p W n n W / p ({ } { })� � � � � � �0 1e (19) with W n np p m m m� � ��� ( ) , where � ��p � is the plasmon energy. Note that these phenomenological rates obey detailed balance and, therefore, at low temperatures only processes that reduce the total plasmon energy occur with apprecia- ble rates. 2.4. Matrix formulation For later convenience, we introduce a matrix nota- tion for the transition rates �, with the matrix ele- ments defined by [ � ( )] ( , { } , { }) ,{ },{ }� � � � � � 0 � �N N n N nn n � 1 (20) i.e, the element ({ }, { })�n n of the matrix block � ( )� � � N is the transition rate ��( , { } , { })N n N n0 � �1 . Simi- larly, [ � ] [ � � ]{ },{ } { },{ } { },{ } { } � � �� � � 0 n n n n n n n � � � � �� � �/ , (21) and [ � ( )] ({ }, { }){ },{ }�p n n pN n n� � � � � ��� �� �/{ },{ } { } ({ }, { })n n p n n n� . (22) Master equation (1) can now be conveniently ex- pressed as d dt P t P t| ( ) � | ( )! � � !� (23) with � � ( � � � ) , � � � � �� � �� � �p L R � � � � 0 , where | ( )P t ! is the column vector (not to be confused with the «ket» in quantum mechanics) with elements given by ! �N n P t P N n t, { }| ( ) ( , { }, ). Therefore, the time evolution of the probability vector satisfies | ( ) exp( � )| ( )P t t P! � � !� 0 . (24) In the long-time limit, the system reaches a steady state | ( )P 4 !. The ensemble averages of the matrices �� � � can then be defined by ! � !� ��� ( ) , { }| � | ( ) ,{ } � �� �t N n P t N n . (25) We will construct other statistical quantities such as av- erage current and noise power density based on Eq. (25). Nonequilibrium plasmons and transport properties of a double-junction quantum wire Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 1527 3. Steady-state probability distribution of nonequilibrium plasmons By solving the master equation (23) numerically (without plasmon relaxation), in Ref. 23, we obtained the occupation probabilities of the plasmonic many- body excitations as a function of the bias voltage and the interaction strength. We found that in the weak to noninteracting regime, * � 0 or g /� �1 1 1( )* for the wire with one transport sector, the nonequilibrium probability of plasmon excitations is a complicated function of the detailed configuration of state occupa- tions { } ( , , ..., , ...)n n n nm� 1 2 . In contrast, the nonequilibrium occupation proba- bility in the strong interaction regime with (nearly) symmetric tunneling barriers depends only on the to- tal energy of the states, and follows a universal form irrespective of electron charge N in the QD. In the leading order approximation, it is given by P Z eV p ( )( ) exp ( ) log0 1 3 1 2 � * � � � � � � � � � � � � � , (26) where Z is a normalization constant. Notice that � is the total energy, including zero-mode and plasmon contributions. The distribution has a universal form which depends on the bias voltage and the interaction strength. The detailed derivation is in Appendix B. This analytic form is valid for the not too low ener- gies ( , )� eV � 3�p and in the strongly interacting re- gime * � 1. More accurate approximation formula (Eq. (B.13)) is derived in Appendix B. For symmetric junctions, the occupation probabili- ties fall on a single curve, well approximated by the an- alytic formulas Eqs. (26) and (B.13), as seen in the in- sets in Fig. 2, where P( )� is depicted as a function of the state energies for g � 0 2. (a) and g � 0 5. (b), with parameters R � 1 for the inset and R � 100 for the main figures (eV p� 6� , N /G � 1 2). For the asymmetric junctions, the line splits into several branches, one for each electric charge N, see the figure. However, as seen in Fig. 2,a, if the interaction is strong enough (g � 0.3 for R � 100 and eV p� 6� ), each branch is, independ- ently, well described by Eq. (26) or (B.13). For weaker interactions, g � 0.3 for R � 100 and eV p� 6� , the analytic approximation is considerably less accurate as shown in Fig. 2,b. Even in the case of weaker inter- actions, however, the logarithms of the plasmon occu- pation probabilities continue to be nearly linear in � but with a slope that deviates from that seen for symmetric junctions. 4. Average current In terms of the tunneling current matrices across the junction L/R � ( � � )IL/R L/R L/Re� �� � � � , (27) the average current I t tL/R L/R( ) � ( )� !I through L/R-junction is I t N n P tL/R N n L/R( ) , { }| � | ( ) ,{ } � !� I . (28) The total external current I t t( ) �( )� !I , which in- cludes the displacement currents associated with charging and discharging the capacitors at the left and right tunnel junctions, is then conveniently writ- ten as 1528 Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 Jaeuk U. Kim, Mahn-Soo Choi, Ilya V. Krive, and Jari M. Kinaret a 100 10–5 10–10 10–15 10–20 10–25 P( )� 010 –1010 –20 10 0 2 4 6 8 10 0 5 10 � �/ p N = 0 N = 1 N = 2 N = 3 P(0) P (1) 100 10–5 10–10 10–15 P( )� 0 2 4 6 8 10 � �/ p N = 0 N = 1 N = 2 N = 3 P(0) P (1) 100 10–5 10 –10 0 5 10 Fig. 2. The occupation occupation probability P( )� as a function of the mode energy � �/ p. The energy � is abbrevia- tion for E N nD( ,{ }), the bias eV p� 6� , the asymmetry pa- rameter R � 100, and N /G � 1 2 (T � 0). Two analytic ap- proximations, Eq. (26) (blue curve) and Eq. (B.13) (cyan curve), are fitted to the probability distribution of the charge mode N � 1 (blue circle). The interaction parameter is g � 02. (a) and 05. (b). In the inset the case of symmetric junctions (R � 1) is plotted with the same conditions. I t C/C I t L R ( ) ( ) ( ), , � � � � � � (29) where C C CL R � 1 1 1. As the system reaches steady-state in the long-time limit, the charge current is conserved throughout the system, I I� 4 �( ) IL( )4 � � 4IR( ). One consequence of nonequilibrium plasmons is the increase in current as shown in Fig. 3, where the aver- age current is shown as a function of the bias for dif- ferent interaction strengths. The currents are normal- ized by I I eV Ec C� �( )2 with no plasmon relaxation (� p � 0) for each interaction strength g, and we see that the current enhancement is substantial in the strong interaction regime (g � 0.5), while there is ef- fectively no enhancement in noninteracting limit g � 1 (the two black lines are indistinguishable in the fig- ure). In the weak interaction limit the current in- creases in discrete steps as new transport channels become energetically allowed, while at stronger inter- actions the steps are smeared to power laws with expo- nents that depend on the number of the plasmon states involved in the transport processes. Including the spin sector results in additional peaks in the average current voltage characteristic that can be controlled by the transverse magnetic field [47,48]. The current-voltage characteristics show that, in the noninteraction limit, the nonequilibrium approach predicts similar behavior for the average current as the detailed balance approach which assumes thermal equilibrium in the QD. In contrast, in the strong interaction regime, nonequilibrium effects give rise to an enhancement of the particle current. Experimentally, however, the current enhancement may be difficult to attribute to plasmon distribution as the current levels depend on barrier transparencies and plasmon relaxation rates, and neither of them can be easily tuned. We now turn to another experimental probe, the shot noise, which is more sensitive to nonequilibrium effects. 5. Current noise Noise, defined by S d t t t t i( ) lim [ �( )�( ) �( ) ]3 ��� ! � ! �� � � (2 2e I I I , (30) describes the fluctuations in the current through a conductor [49,50]. Thermal fluctuations and the discreteness of the electron charge are two fundamen- tal sources of the noise among others which are sys- tem specific [30,31]. Thermal (equilibrium) noise is not very informative since it does not provide more information than the equilibrium conductance of the system. In contrast, shot noise is a consequence of the discreteness of charge and the stochastic nature of transport. It thus can provide further insight beyond average current since it is a sensitive function of the correlation mechanism [29], internal excitations [51–53], and the statistics of the charge carriers [54–57]. Here we will particularly focus on the roles of the nonequilibrium plasmon excitations in the SETs of 1D QWs [21–24]. Within the master equation approach, the correlation functions K t t t�� � �� �� �� !( ) lim � ( ) � ( ) I I in Eq. (30) can be deduced from the master equation (23). In the matrix notation they can be written as [27,28] K e N n N n �� � �� �� � �( ) , { }| ( )� exp( � � ,{ } [ �2 . I I� � 4 !� � � .( )� exp( � )� ( ) ( � � ) | ( ) ,] / /I I� � �� � �� � � P (31) where .( )x is the unit step function. In principle, there is no well-known justification of the master equation approach, which ignores the quantum coher- ence effects, for the shot noise. Actually, influence of quantum coherence on shot noise is an intriguing is- sue [31]. However, we note that both quantum me- chanical approaches [58,59] and semiclassical deriva- tions based on a master equation approach predict identical shot noise results [26,60,61], implying that the shot noise is not sensitive to the quantum coher- ence in double-barrier structures. Master equation ap- proach was used by many authors for shot noise in SET devices as well [26–28,60,62,63], and some pre- Nonequilibrium plasmons and transport properties of a double-junction quantum wire Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 1529 0 0.5 1.0 eV/2E c 4 3 2 1 cI/I g = 0.3 g = 0.3, �p g = 0.5 g = 0.5, �p g = 0.7 g = 0.7, �p g = 1.0 g = 1.0, �p Fig. 3. Average current I I/ c as a function of the bias volt- age eV/ EC2 and LL interaction parameter g for R � 100 (highly asymmetric junctions) with no plasmon relaxation (�p � 0, solid lines) or with fast plasmon relaxation (�0 410� , dashed lines). The bias voltage is normalized by the charging energy 2EC and current is normalized by the current at eV EC� 2 with no plasmon relaxation for each g. Other parameters are NG � 1/2, T � 0. dictions of them [27] have been experimentally con- firmed [64]. On this ground, here we adopt the mas- ter equation approach and leave the precise test of the justification open to either experimental or further theoretical test. To investigate the correlation effects, the noise power customarily compared to the Poisson value S eIPoisson � 2 . The Fano factor is defined as the ratio of the actual noise power and the Poisson value, F S eI � ( )0 2 . (32) Since thermal noise ( ( ))S k TG VB� �4 0 is not par- ticularly interesting, we focus on the zero frequency shot noise, in the low bias voltage regime where the Coulomb blockade governs the electric transport; T � 0 and eV � 2EC. 5.1. Qualitative discussions As we discuss below in detail, we have found in the absence of substantial plasmon relaxation a giant en- hancement of the shot noise beyond Poissonian limit (F � 1) over wide ranges of bias and gate voltages; i.e., the statistics of the charge transport through the device is highly super-Poissonian. In the opposite limit of fast plasmon relaxation rate, the shot noise is reduced below the Poissonian limit (still exhibiting features specific to LL correlations) in accordance with the previous work [22]. Before going directly into details, it will be useful to provide a possible physical interpretation of the result. We ascribe the giant super-Poissonian noise to the opening of additional conduction channels via the plasmon modes. In the parameter ranges where the gi- ant super-Poissonian noise is observed, there are a con- siderable amount of plasmon excitations [23]. It means that the charges can have more than one possi- ble paths (or, equivalently, more than two local states in the central island involved in the transport) from left to right leads, making use of different plasmon modes. Similar effects have been reported in conven- tional SET devices [29] and single-electron shuttles [65]. It is a rather general feature as long as the multi- ple transport channels are incoherent and have differ- ent tunneling rates. An interesting difference between our results and the previous results [29,65] is that the super-Poissonian noise is observed even in the sequen- tial tunneling regime whereas in Ref. 29 it was ob- served in the incoherent cotunneling regime and in [65] due to the mechanical instability of the elec- tron-mediating shuttle. Notice that the fast plasmon relaxation prevents the additional conduction channel opening, and in the presence of fast plasmon relaxation, the device is qualitatively the same as the conventional SET. The noise is thus reduced to the Poissonian or weakly sub-Poissonian noise. To justify our interpretation, in the following two subsections we compare two parameter regimes with only a few states involved in the transport, which are analytically tractable. When only two charge states (with no plasmon excitations) are involved (Sec. 5.2), the transport mechanism is qualitatively the same as the usual sequential tunneling in a conventional SET. Therefore, one cannot expect an enhancement of noise beyond the Poissonian limit. As the bias voltage in- creases, there can be one plasmon excitation associated with the lowest charging level (Sec. 5.3). In this case, through the three-state approximation, we will expli- citly demonstrate that the super-Poissonian noise arises due to fluctuations induced by additional conduction channels. Detailed analysis of the full model system is provided in subsequent subsections. 5.2. Two-state model (eV p� � ) The electron transport involving only two lowest energy states in the quantum dot are well studied by many authors (see, for instance, Ref. 31). Neverthe- less, for later reference we begin the discussion of shot noise with two-state process, which provides a reason- able approximation for eV p� � . At biases such that eV p� � and sufficiently low temperatures, the two lowest states | , | ,N n1 0 0! � ! and | ,10! dominate the transport process and the rate matrix is given by � ,� � � � " # $ $ % & ' ' � � � � � � (33) where the matrix elements are � � � ��L( , , )10 0 0 and � � ��R( , , )0 0 10 . With the current matrices defined by Eq. (27) � ,IL � " # $ % & '� 0 0 0� � ,IR � " # $ % & ' 0 0 0 � the noise power is obtained straightforwardly by Eqs. (30) and (31) using the steady state probability | ( )P P P 4 ! � " # $ % & ' � " # $ $ % & ' '� � 00 10 1 � � � � . (34) The Fano factor (32) takes a simple form F P P / / 2 00 2 10 2 2 2 1 1 � � � � ( ) ( ) � � � � , (35) where 1530 Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 Jaeuk U. Kim, Mahn-Soo Choi, Ilya V. Krive, and Jari M. Kinaret � � / / � � � � � � �� � � �� 1 2 2 0 0R eV/ E eV/ E and /E0 is the shift of the bottom of the zero mode energy induced by the gate voltage, / / � /E N /g N N /G p G G0 1 2� � �( ) , . (36) Note that Eq. (36) is valid for | |/E eV/0 25 , other- wise I � 0 and S � 0 due to Coulomb blockade. We see from Eq. (35) that the Fano factor is minimized for � � � �/ 1 and maximized for ( )� � � � �/ 1 0, with the bounds 1 2 12/ F5 � . At the gate charge N /G � 1 2, it is determined only by the junction asymmetry parameter R: F R / R2 2 21 1� ( ) ( ) . Note that when only two states are involved in the current carrying process (ground state to ground state transitions), the Fano factor cannot exceed the Pois- son value F � 1. As a consequence of the power law dependence of the transition rates on the transfer energy (13), the Fano factor is a function of the bias voltage, and the interaction strengts, it varies between the minimum and maximum values F E R R eV E eV / / 2 0 1 1 0 1 2 1 1 2 1 2 � � � � 0 6 7 88 9 8 8 at at / / � � (37) (cf. Eq. (5) in Ref. 22). Figure 4 depicts the Fano factor as N / g eV/ R R G p / / dip � � 1 2 2 1 1 1 1 ( )� � � . (38) The Fano factor independently of the interaction strength crosses F R / R2 2 21 1� ( ) ( ) at N /G � 1 2, and it approaches maximum F2 1� at N /G � 01 2 0 g eV/ p( )2� . 5.3. Three-state model (eV � 2�p) The two-state model is applicable for bias voltages below eV Epth � �2 0( | | )� / . Above this threshold voltage, three or more states are involved in the trans- port. For electron transport involving three lowest en- ergy states in the quantum dot | , | , , | ,N n1 0 0 10! � ! !, and | ,11! with nm � 0 for m : 2, the noise power can be cal- culated exactly if the the contribution from the (back- ward) transitions against the bias is negligible, as is typically the case at zero temperature. In practice, however, the backward transitions are not completely blocked for the bias above the threshold voltage of the plasmon excitations, even at zero temperature: once the bias voltage reaches the threshold to initiate plasmon excitations, the high-energy plasmons in the QD above the Fermi energies of the leads are also par- tially populated, opening the possibility of backward transitions. A qualitatively new feature that can be studied in the three-state model as compared to the two-state model is plasmon relaxation: the system with a con- stant total charge may undergo transitions between different plasmon configurations. We will show in this subsection that the analytic solution of the Fano factor of the three-state process yields an excellent agreement with the low bias nu- merical results in the strong interaction regime, while it shows small discrepancy in the weak interaction re- gime (due to non-negligible contribution from the high-energy plasmons). We will also show that within the three state model the Fano factor may exceed the Poisson value. Analytic results By allowing plasmon relaxation, the rate matrix in- volving three lowest energy states | , | , ,| , ,N n1 0 0 10! � ! ! and | ,11! is given by �� � � � � � � � � � � � 0 1 0 1 0 0 1 10 � � � � � � � � � " # $ $ $ $ % & ' ' ' p p ' , (39) with the matrix elements � i L i� � �� ( , , ),1 0 0 � i � � � ��R i i( , , ), ,0 0 1 01 and � p introduced in Eq. (19). Current matrices defined by Eq. (27) are [� ] , [� ]{ , } { , }I IL L21 0 31 1� �� �� � , [� ] , [� ] ,{ , } { , }I IR R12 0 13 1� � � � with [� ]{ , }I� i j � 0 for other set of i j, , ,� 1 2 3, where � � L R, . The noise power is obtained straightfor- Nonequilibrium plasmons and transport properties of a double-junction quantum wire Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 1531 0 0.2 0.4 0.6 0.8 NG 1.5 1.0 0.5 F g = 0.2 g = 0.3 g = 0.5 g = 0.7 g = 1.0 1 – 2 – 3 – 5 – 4 – 1234 5 Fig. 4. Fano factor F S / eI� ( )0 2 as a function of the gate charge NG and LL interaction parameter g for R = 100 (highly asymmetric junctions) at eV p� � (T � 0). wardly by Eqs. (30) and (31) using the steady state probability | ( ) ( ) (P P P P Z p p4 ! � " # $ $ $ % & ' ' ' � � 00 10 11 0 1 0 1 1 � � � � � � ) " # $ $ $ $ % & ' ' ' ' � � � � � � 1 1 0 p , (40) with normalization constant Z � � � � � � � � �0 1 1 0 0 1 � �( )� � � �0 0 1 p . Using the average current I � � e P P( )10 0 11 1� � , the Fano factor F S / eI� ( )0 2 is given by F P P P P Z3 00 2 10 2 11 2 11 0 0 2 1 � ; � � � ; � " # $ $ % & ' ' � � � ( ) ( ) .� � � � � � � � �0 2 1 2 0 1 0 1 0 1 p (41) Compared to the Fano factor (35) in the two-state process, complication arises already in the three-state process due to the last term in Eq. (41) which results from the coupling of P11 and the rates which cannot be expressed by the components of the probability vector. In order to have | , | , , | , ,N n! � ! !0 0 10 and | ,11! as the relevant states, we assume � �0 0 � � or more explicitly N NG G: dip which is introduced in Eq. (38). In the opposite situation ( )� �0 0 � � , the relevant states are | , | , , | , ,N n1 0 0 01! � ! ! and | ,10!, and the above description is still valid with the exchange of electron number N � <0 1 and the corresponding notations � �i i � < , i � 0 1, . To see the implications of Eq. (41), we plot the Fano factor in Fig. 5, with respect to the gate charge NG for symmetric junctions at eV p� 2� (T � 0), with no plasmon relaxation (� p � 0). Two main features are seen in Fig. 5. Firstly, the shot noise is enhanced over the Poisson limit (F � 1) in the strong interaction regime, g � 0.5, for a range of parameters with gate charges away from N /G � 1 2. As discussed above, in the low bias regime eV Ep� �2 0( | | )� / at zero temperature, no plasmons are excited and the electric charges are transported via only the two-state process following the Fano factor (35) which results in the sub-Poissonian shot noise ( )1 2 1/ F5 5 . Once the bias reaches the threshold eV Epth � �2 0( | | )� / , it initiates plasmon excitations which enhance the shot noise over the Poisson limit. This feature is discussed in more detail below. Secondly, in the weak interaction regime (g � 0.5) a small discrepancy between the analytic result (41) (dashed line) and the numerical result (solid line) is found. It results from the partially populated states of the high-energy plasmons over the bias due to nonvanishing transition rates. On the other hand, a simple three-state approximation shows excellent agreement in the strong interaction regime (g � 0 3. in the Fig. 5), indicating negligible contribution of the high-energy plasmons (E N n eV/D( , { }) � 2) to the charge transport mechanism. This is due to the power law suppression of the transition rates (13) as a func- tion of the transfer energy (10). Limiting cases To verify the role of nonequilibrium plasmons as the cause of the shot noise enhancement, we consider two limiting cases of Eq. (41): � p � 0 and � �p i�� � . In the limit of no plasmon relaxation (� p � 0), the Fano factor (41) of the three-state process is simpli- fied as F P P P P3 0 10 10 11 111 2 1 1( ) [( ) ( ) ]� � � � 2 10 11 0 2 1 2 0 1 P P ( ) ( ) , � � � � (42) with steady-state probability | ( )P P P P Z 4 ! � " # $ $ $ % & ' ' ' � " # $ � � 00 10 11 0 1 0 1 1 0 1 � � � � � � $ $ $ % & ' ' ' ' , (43) where the new normalization constant is Z � � � �0 1 � � � � �1 0 0 1 . The three-state approximation is most accurate in the low bias regime 2 0( | | )� /p E� � eV EC�� 2 (/E0 is defined in Eq. (36)) and for gate voltages away from N /G � 1 2, i.e., for 1 2/ NG� � 1 (or 0 � N /G � 1 2 with the exchange of indices regarding particle number N � <0 1) . In this regime, �0 � and/or �1 dominate 1532 Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 Jaeuk U. Kim, Mahn-Soo Choi, Ilya V. Krive, and Jari M. Kinaret 2.0 1.5 1.0 0.5 0 F NG0.2 0.5 0.8 g = 0.3, N g = 0.3, A g = 0.5, N g = 0.5, A g = 1.0, N g = 1.0, A g = 0.3 0.5 1.0 Fig. 5. Fano factor F S / eI� ( )0 2 as a function of the gate charge NG for symmetric junctions (R � 1) at voltage eV p� 2� (T � 0) with no plasmon relaxation. Numerical results (solid lines) versus analytic results with three states, Eq. (41) (dashed lines) for g � 03. , 05. , and 10. . over the other rates, ( , ) ( , )� � � �0 1 0 1 � ��� , which results in 1 � P P P10 0 11 0�� �( , ) . The Fano factor (42) now is reduced to F P P3 0 10 11 0 2 1 2 0 1 1 0 1 2 1 2( ) ( ) ( ) .� � � � � � � � � � (44) Notice that while Eq. (42) is an exact solution for the three-state process with no plasmon relaxation, Eq. (44) is a good approximation only sufficiently far from N /G � 1 2. In this range, Eq. (44) explicitly shows that the opening of new charge transport chan- nels accompanied by the plasmon excitations causes the enhancement of the shot noise (over the Poisson limit). Recently, super-Poissonian shot noises, i.e., the Fano factor F � 1, were found in several different situations in quantum systems. Sukhorukov et al. [29] studied the noise of the co-tunneling current through one or several quantum dots coupled by tunneling junctions, in the Coulomb blockade regime, and showed that strong in- elastic co-tunneling could induce super-Poissonian shot noise due to switching between quantum states carrying currents of different strengths. Thielmann et al. [66] showed similar super-Poissonian effect in a single-level quantum dot due to spin-flip co-tunneling processes, with a sensitive dependence on the coupling strength. The electron spin in a quantum dot in the Coulomb blockade regime can generate super-Poissonian shot noise also at high frequencies [67]. The shot noise en- hancement over the Poisson limit can be observed by studying the resonant tunneling through localized states in a tunnel-barrier, resulting from Coulomb interaction between the localized states [68]. In nano-electro-me- chanical systems, in the semiclassical limit, the Fano factor exceeds the Poisson limit at the shuttle threshold [65,69]. Commonly, the super-Poissonian shot noise is accompanied by internal instability or a multi-channel process in the course of electrical transport. In the limit of fast plasmon relaxation, on the other hand, � �p i�� � and effectively no plasmon is excited, | ( )P P P P 4 ! � " # $ $ $ % & ' ' ' � � � � � 00 10 11 0 1 0 0 0 1 1 0� � � � � � " # $ $ $ $ % & ' ' ' ' . (45) Consequently, the Fano factor (41) is given by F P3 00 21 2 2 1 2 1 2 1( ) ,� � �� � � � � � � " #$ % &' . (46) The maximum Fano factor F � 1 is reached if one of the rates �0 � or �0 dominates, while the minimum value F /� 1 2 requires that � � �0 0 1 � �� , i.e., that the total tunneling-in and tunneling-out rates are equal; more explicitly, 1 1 2 2 1 1 2 0 0 0R E eV/ E g eV/ E p� � � �� � � �� � � � � � � � � � � � / / � / � � 1. (47) 5.4. Numerical results The limiting cases of no plasmon relaxation (� p � 0) and a fast plasmon relaxation (� p � 104) are summarized in Fig. 6,a and b, respectively. In the fig- ure the Fano factor is plotted as a function of gate voltage and interaction parameter g in the strong in- teraction regime 0 3 0 6. .5 5g at eV p� 2� for R � 100 (strongly asymmetric junctions). As shown in Fig. 6,a, the Fano factor is enhanced above the Poisson limit (F � 1) for a range of gate charges not very close to N /G � 1 2, especially in the strong interaction regime, as expected from the three-state model. The shot noise enhancement is lost in the presence of a fast plasmon relaxation process, in agreement with analytic arguments, as seen in Fig. 6,b when F is bounded by 1 2 1/ F5 5 . Hence, slowly re- Nonequilibrium plasmons and transport properties of a double-junction quantum wire Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 1533 F g NG 2.5 2.0 1.5 1.0 0.5 0.7 0.6 0.5 0.4 0.3 0.2 0.3 0.4 0.5 0.6 F 1.0 0.9 0.8 0.7 0.6 0.5 NG g 0.7 0.6 0.5 0.4 0.3 0.2 0.6 0.5 0.4 0.3 b Fig. 6. Fano factor F S / eI� ( )0 2 as a function of the gate charge NG and LL interaction parameter g for R � 100 (highly asymmetric junctions) at eV p� 2� (T � 0), with no plasmon relaxation (�p � 0) (a) and with fast plasmon relaxation (�p � 104) (b). laxing plasmon excitations enhance shot noise, and this enhancement is most pronounced in the strongly interacting regime. In the limit of fast plasmon relaxation F exhibits a minimum value F /� 1 2 at positions consistent with predictions of the three-state model: the voltage polar- ity and ratio of tunneling matrix elements at the two junctions is such that total tunneling-in and tunnel- ing-out rates are roughly equal for small values of NG . If plasmon relaxation is slow, F still has minima at ap- proximately same values of NG but the minimal value of the Fano factor is considerably larger due to the presence of several transport channels. 5.5. Interplay between several charge states and plasmon excitations near eV EC� 2 So far, we have investigated the role of nonequi- librium plasmons as the cause of the shot noise en- hancement and focused on a voltage range when only two charge states are significantly involved in trans- port. The question naturally follows what is the conse- quence of the involvement of several charge states. Do they enhance shot noise, too? To answer this question, we first consider a toy model in which the plasmon excitations are absent dur- ing the single-charge transport. In the three-N-state re- gime where the relevant states are | | , |N! � � ! !1 0 and |1! with no plasmon excitations at all. At zero tempera- ture, the rate matrix in this regime is given by �� � � � � � " # $ $ $ $ % & ' ' ' ' � � � � � � � � � � � � 1 0 1 0 0 1 0 1 0 0 , (48) where the matrix elements are �i � � �L i i( , { } , { }), �1 0 0 � �i R i i � � �( , { } , ),1 0 0 i � �1 0 1, , . Repeating the procedure in Sec. 5.3, we arrive at a Fano factor that has a similar form as Eq. (42), F P P P PN3 1 1 1 11 2 1 1� � � � [( ) ( ) ] � � � � � � � � � � 2 1 1 1 1 1 1 P P � � � � , (49) with the steady-state probability vector | ( )P P P P Z 4 ! � " # $ $ $ % & ' ' ' � " # $ $ � � � 1 0 0 1 1 1 1 0 1 � � � � � � $ $ % & ' ' ' ' , (50) where Z � � � �� � � � � �0 1 1 1 1 0 . Despite the formal similarity of Eq. (49) with Eq. (42), its implication is quite different. In terms of the transition rates, F N3 reads F Z N3 2 1 0 1 0 1 0 11 2 � � � � � � � [ ( ( ) )� � � � � � � � � � � � � � � � �1 0 0 1 1 0 1( ( ) )] . (51) Since � � � ��1 0 and � �1 0 � in the three-N-state re- gime, the Fano factor F N3 is sub-Poissonian, i.e., F N3 1� , consistent with the conventional equilibrium descriptions [22,27]. We conclude that while plasmon excitations may enhance the shot noise over the Poisson limit, the involvement of several charge states and the ensuing correlations, in contrast, do not alter the sub-Pois- sonian nature of the Fano factor in the low-energy re- gime. This qualitative difference is due to the fact that certain transition rates between different charge states vanish identically (in the absence of co-tunneling): it is impossible for the system to move directly from a state with N � �1 to N � 1 or vice versa. Therefore, we expect that for bias voltages near eV EC p� �2 2� , when both plasmon excitations and several charge excitations are relevant, the Fano fac- tor will exhibit complicated nonmonotonic behavior. Exact solution is not tractable in this regime since it involves too many states. Instead, we calculate the shot noise numerically, with results depicted in Fig. 7, where the zero temperature Fano factor is shown as a function of the bias eV and LL interaction parameter g for R � 100 at N /G � 1 2, (a) with no plasmon relaxation ( )� p � 0 and (b) with fast plasmon relaxation (� p � 104). In the bias regime up to the charging energy eV EC5 2 , the Fano factor increases monotonically due to nonequilibrium plasmons. On the other hand, additional charge states contribute at eV EC: 2 which tends to suppress the Fano factor. As a consequence of this competition, the Fano factor reaches its peak at eV EC� 2 and is followed by a steep decrease as shown in Fig. 7,a. Note the significant enhancement of the Fano factor in the strong interaction regime, which is due in part to the power law dependence of the transition rates with exponent * � �( )1 1/g as dis- cussed earlier, and in part to more plasmon states be- ing involved for smaller g since E /gC p� � . The latter reason also accounts for the fact that the Fano factor begins to rise at a lower apparent bias for smaller g: the bias voltage is normalized by EC so that plasmon excitation is possible for lower values of eV/EC for stronger interactions. In the case of fast plasmon relaxation, the rich structure of the Fano factor due to nonequilibrium 1534 Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 Jaeuk U. Kim, Mahn-Soo Choi, Ilya V. Krive, and Jari M. Kinaret plasmons is absent as shown in Fig. 7,b, in agreement with the discussions in previous subsections. The only remaining structure is a sharp dip around eV EC� 2 that can be attributed to the involvement of addi- tional charge states at eV EC: 2 . Not only the mini- mum value of the Fano factor is a function of the in- teraction strength but also the bias voltage at which it occurs depends on g. The minimum Fano factor occurs at higher bias voltage, and the dip tends to be deeper with increasing interaction strength. Note for g � 0.5, the Fano factor did not reach its minimum still at larg- est voltages plotted (eV/ EC2 12� . ). The Fano factor at very low voltages eV p� 2� for N /G � 1 2 is F R / R( ) ( ) ( )0 2 21 1� (see Eq. (35)), regardless of the plasmon relaxation mechanism. As shown in Fig. 7,b, the Fano factor is bounded above by this value in the case of fast plasmon relaxation. Notice that in the case of no plasmon relaxation (Fig. 7,a), the dips in the Fano factor at eV EC� 2 reach below F ( )0 . See Fig. 1,a in Ref. 24 for more detail. Since both the minimum value of F and the voltage at which it occurs are determined by a competition be- tween charge excitations and plasmonic excitations, they cannot be accurately predicted by any of the sim- ple analytic models discussed above. 6. Full counting statistics Since shot noise, that is a current-current correla- tion, is more informative than the average current, we expect even more information with higher-order cur- rents or charge correlations. The method of counting statistics, which was introduced to mesoscopic physics by Levitov and Lesovik [70] followed by Muzy- kantskii and Khmelnitskii [71] and Lee et al. [72], shows that all orders of charge correlation functions can be obtained as a function related to the probabil- ity distribution of transported electrons for a given time interval. This powerful approach is known as full counting statistics (FCS). The first experimental study of the third cumulant of the voltage fluctuations in a tunnel junction was carried out by Reulet et al. [73]. The experiment indicates that the higher cumulants are more sensitive to the coupling of the system to the electromagnetic environment. See also Refs. 74, 75 for the theoretical discussions on the third cumulant in a tunnel-barrier. We will now carry out a FCS analysis of transport through a double barrier quantum wire system. The analysis will provide a more complete characterization of the transport properties of the system than either average current or shot noise, and shed further light on the role of nonequilibrium versus equilibrium plasmon distribution in this structure. Let P M( , ) be the probability that M electrons have tunneled across the right junction to the right lead during the time . We note that P M( , ) � � � � �� ��lim ( , , { }, ; , , { , ,{ },{ } t P M M N n t M N n M N nN n 0 0 0 0 0 0 0 � }, ),t where P M M N n t M N n t( , , { }, ; , , { }, )0 0 0 0 is cal- led joint probability since it is the probability that, up to time t, M0 electrons have passed across the right junction and N0 electrons are confined in the QD with { }n0 plasmon excitations, and that M M0 electrons have passed R-junction with ( , { })N n excita- tions in the QD up to time t . The master equation for the joint probability can easily be constructed from Eq. (23) by noting that M M- 0 1 as N N- � 1 via only R-junction hopping To obtain P M( , ) , it is convenient to define the characteristic function conjugate to the joint proba- bility as g N n( , , { }, )= � Nonequilibrium plasmons and transport properties of a double-junction quantum wire Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 1535 F 7 5 3 1 0 1 2eV/E C 0.4 0.6 0.8 1.0 g a F 2.0 2.2 2.4eV/E C g 0.4 0.6 0.8 1.0 1.0 0.9 0.8 0.7 0.6 0.5 b Fig. 7. Fano factor F S / eI� ( )0 2 as a function of the bias eV and LL interaction parameter g for R � 100 (highly asymmetric junctions) at N /G � 1 2 (T � 0): with no plas- mon relaxation (�p � 0) (a) and with fast plasmon relax- ation (�p � 104) (b). � � � �� � �lim ( , , { }, ; , , { , ,{ } t i M M M N n P M M N n t M N ne � �0 0 0 0 0 0 0 }, ).t The characteristic function satisfies the master equation � � ! � � ! = = = | ( , ) �( )| ( , )g g� with the initial condition | ( , ) | ( )g P= � ! � 4 !0 . The =-dependent �� in Eq. (52) is related to the previously defined transition rate matrices through �( ) � ( � � � )� � � �= � � � � �p L L L 0 � �� ( � � � ) .� � �R R i R i0 e e� � (53) The characteristic function G( , )= conjugate to P M( , ) is now given by G N n g N n ( , ) , { }| ( , ) ,{ } = = � !� , or G N n P N n ( , ) , { }| exp [ �( ) ]| ( ) . ,{ } = = � � 4 !� � (54) Finally, the probability P M( , ) is obtained by P M d G dz i G z z i M M ( , ) ( , ) ( , ) = � = � � �� �( ( �2 2 0 2 2 1 e � (55) with z i /� e � 2, where the contour runs counterclock- wise along the unit circle and we have used the symme- try property G z G z( , ) ( , ) � � for the second equality. Taylor expansion of the logarithm of the character- istic function in i= defines the cumulants or irreducible correlators > k( ): ln ( , ) ( ) ! ( )G i k k k k= = > � � � � 1 . (56) The cumulants have a direct polynomial relation with the moments n n P nk k n ( ) ( , ) � � . The first two cumulants are the mean and the variance, and the third cumulant characterizes the asymmetry (or skew- ness) of the P M( , ) distribution and is given by > / 3 3 3( ) ( ) ( ( ) ( ))� � �n n n . (57) In this section, we investigate FCS mainly in the context of the probability P M( , ) that M electrons have passed through the right junction during the time . Since the average current and the shot noise are proportional to the average number of the tunnel- ing electrons !M and the width of the distribution of P M( , ) , respectively, we focus on the new aspects that are not covered by the study of the average cur- rent or shot noise. In order to get FCS in general cases we integrate the master equation (52) numerically (see Sec. 6.3). In the low-bias regime, however, some analytic calcu- lations can be made. We will show through the fol- lowing subsections that for symmetric junctions in the low-bias regime (2 2�p CeV E� � ), and irrespective of the junction symmetry in the very low bias regime (eV p� 2� ), P M( , ) is given by the residue at z � 0 alone, P M M d dz G z M M z ( , ) ( )! ( , ) . � , , , � 1 2 2 2 0 (58) Through this section we assume that the gate charge is N /G � 1 2, unless stated explicitly not so. We will now follow the outline of the previous sec- tion and start by considering two analytically tracta- ble cases before proceeding with the full numerical re- sults. 6.1. Two-state process; eV p� � For the very low bias eV p� 2� at zero temperature, no plasmons are excited and electrons are carried by transitions between two states ( , ) ( , ) ( , )N n � <0 0 10 . In this simplest case, the rate matrix �( )� = in Eq. (52) is determined by only two participating transition rates � � � ��L( , , )10 0 0 and � � ��R( , , )0 0 10 , �( ) ( ) ( ) � = / / / / � � � � � � " # $ $ % & ' ' � � � � 0 0 0 0 e i (59) with �0 2 2� � �� � ( ) , ( )� � / � �/ / . Substituting the steady-state probability Eq. (34) and the transition rate matrix (59) to Eq. (54), one finds G z f z f z f z 2 2 2 20 0 2 4 1 1( , ) ( ) [( ( )) ( ) � �� � e e � � � � ( ( )) ],( )1 2 2 0 2f z f ze � � (60) where f z / z2 0 2 2( ) ( )� �� � �� , with � �� /� � �0 2 2 � � � � and � �2 2 2� / �/ . Now, it is straightforward to calculate the cu- mulants. In the long time limit ��� � ( ) 1, for instance, in terms of the average current I2 � � � � e /� � � �( ) and the Fano factor F2 in (35), the three lowest cumulants are given by > 1 2( ) ,� I > 2 2 2( ) ,� eI F > 3 2 2 2 23 1 2 1 4( ) [ ( ) ]� � e I F / / , (61) where the electron charge (�e) is revived. These are in agreement with the phase-coherent quantum-mechani- cal results [76]. It is convenient to discuss the asym- 1536 Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 Jaeuk U. Kim, Mahn-Soo Choi, Ilya V. Krive, and Jari M. Kinaret metry (skewness) by the ratio A /e� > >3 2 1 ( ) - 4 , noticing the Fano factor F /e� �� lim ( ( ) � > � > 2 1 . The factor A F / /2 2 23 1 2 1 4� � ( ) is positive definite (positive skewness) and bounded by 1 4 12/ A5 5 . It is interesting to notice that A2 is a monotonic func- tion of F2 and has its minimum A /2 1 4� for the mini- mum F /2 1 2� and maximum A2 1� for the maximum F2 1� . Notice A � 1 for a Poissonian and A � 0 for a Gaussian distribution. Therefore, the dependence of A2 on the gate charge NG is similar to that of the Fano factor F2 (see Fig. 4) with dips at NG dip in Eq. (38). Together with Eq. (37), it implies that the ef- fective shot noise and the asymmetry of the probabil- ity distribution per unit charge transfer have their re- spective minimum values at the gate charge NG dip which depends on the tunnel-junction asymmetry and the interaction strength of the leads. The integral in Eq. (55) is along the contour de- picted in Fig. 8. Notice that the contributions from the part along the branch cuts are zero and we are left with the multiple poles at z � 0. By residue theorem, the two-state probability P M2( , ) is given by Eq. (58). The exact expression of P M2( , ) is cumbersome. For symmetric tunneling barriers with N /G � 1 2 ( , )/ � �� � �� 0 0� , however, Eq. (60) is reduced to G z z z zs z z 2 2 20 0 0 4 1 1( )( , ) [( ) ( ) ] � � �� � � e e e � � � . (62) Accordingly, P Ms 2 ( )( , ) is concisely given by P M M M s M M 2 0 2 1 0 2 0 1 2 2 1 2 ( )( , ) ( ) ( )! ( ) ( )! �� � " # $ e � � � $ % & ' ' : �1 2 2 1 10 2 1( ) ( )! for � M M M , (63) with P /s 2 00 1 20( )( , ) ( ) �� e � � , in agreement with Eq. (24) of Ref. 76. While this distribution resembles a sum of three Poisson distributions, it is not exactly Poissonian. For a highly asymmetric junctions R �� 1 ( )� �� �� , the first term in Eq. (60) dominates the dynamics of G z2( , ) and its derivatives, and the char- acteristic function is approximated by G z za 2 0 2 2( )( , ) exp ( ( ) ) � � /� � � � . (64) Now, the solution of P Ma 2 ( )( , ) is calculated by this equation and Eq. (58). The leading order approxima- tion in � � �/ leads to the Poisson distribution, P M M a M 2 ( )( , ) ( ) ! . � � �� � e (65) For a single tunneling-barrier, the charges are trans- ported by the Poisson process (F � 1) [31]. Therefore, we recover the Poisson distribution in the limit of strongly asymmetric junctions and in the regime of the two-state process, in which electrons see effec- tively single tunnel-barrier. For the intermediate barrier asymmetry the proba- bility P M2( , ) of a two-state process is given by a dis- tribution between Eq. (63) (for symmetric junctions) and the Poissonian (65) (for the most asymmetric junctions). 6.2. Four-state process; 2�p � eV � 2EC Since we focus the FCS analysis on the case N /G � 1 2, the next simplest case to study is a four- state-model rather than the three-state-model dis- cussed in the connection of the shot noise in the previ- ous section. In the bias regime where the transport is governed by Coulomb blockade (eV EC� 2 ) and yet the plas- mons play an important role (eV p: 2� ), it is a fairly good approximation to include only the four states with N � 0 1, and n1 0 1� , (nm � 0 for m : 2). For general asymmetric cases, the rate matrix �( )� = in Eq. (52) is determined by ten participating transition rates (four rates from each junction, and two relax- ation rates). The resulting eigenvalues of | ( , )g = ! are the solutions of a quartic equation, which is in general very laborious to present analytically. Nonequilibrium plasmons and transport properties of a double-junction quantum wire Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 1537 C2 C1 – /2 + /2 + /2 – /2 C0 C 3 C4 Fig. 8. Contour of Eq. (55). The arguments of f z2( ) � � �( )� � ��/ z0 2 2 are / / / /2 2 2 2, , , along the branch cuts C C C1 2 3, , , and C4, respectively. For symmetric junctions (R � 1) with no plasmon relaxation, however, the rate matrix is simplified to �( )� = � � � � � � � � � � � � � � � 00 10 00 2 01 2 01 11 10 2 11 2 00 0 0 z z z z � � � " # $ $ $ $ $ % & ' ' ' ' ' � � � � � � � 01 00 10 10 11 01 11 0 0 (66) with the matrix elements � ij L i j� � �� ( , , )1 0 � ��R i j( , , )0 1 . The steady-state probability is then given by | ( ) | ( , ) ( ) ( )P g zs 4 01 10 01 10 10 0 1 2 4 ! � � ! � " # $ $ $ � � � � � � �� $ % & ' ' ' ' . (67) Solving Eq. (54) with this probability and the rate matrix (66) is laborious but straightforward and one finds G z /( , ) ( ) � � � � �� ; � � �e 00 10 01 11 2 ; 6 7 8 98 ? @ 8 A8 - � " # $ $ % & ' ' G z z z z zI ( , ) ( ) { } 1 8 2 , (68) with G z A z f zI z z f z /( , ) ( ) ( ) ( ( )) � � �� � � �� � � �� 6 7 � �e 00 11 4 2 4 1 9 ; ; � � � � �� � � �� ? @ A - �1 2 4 01 10 4 4 A z f z z f z f z ( ) ( ) ( ) { ( ) ( )} � � , (69) where A z( ) and f z4( ) are given by A z z( ) ( )� � � � � �� � � � � � �00 10 01 11 00 11 012 , f z z a b4 00 11 2 10 01 2 24( ) ( ) ( )� � � � � � � � (70) with dimensionless parameters a b, given by a � � � � � � ( )( ) ( ) , � � � � � � � � � � 00 11 00 10 01 11 00 11 2 10 014 b � � � � � 4 4 10 01 00 10 01 11 00 11 2 10 01 � � � � � � � � � � ( ) ( ) . Integral of G zI ( , ) along the contour | |z � 1 contains two branch points at z a ibc � 0 , however, the inte- gral along the branch cuts cancel out due to the sym- metry under [ ( ) ( )]f z f z4 4- � . Therefore, the contri- bution from the branch cuts due to G zI ( , ) and G zI ( , )� is zero to the probability P Ms 4 ( )( , ) , and it is given by the residues only at z � 0, i.e., by Eq. (58). The explicit expression of P Ms 4 ( )( , ) is cumbersome. The probability distribution P M2( , ) for the two- state process deviates from Eq. (63) as a function of the asymmetry parameter R and reaches Poissonian in the case of strongly asymmetric junctions. In a similar manner, P Ms 4 ( )( , ) deviates from P Ms 2 ( )( , ) as a function of ratio of the transition rates � ij . 6.3. Numerical results It is worth mentioning that for strongly asymmetric junctions P M( , ) is Poissonian in the very low bias regime (eV p� 2� ), as seen from Eq. (65). It exhibits a crossover at eV p� 2� : P M( , ) deviates from Pois- son distribution for 2 2�p CeV E� � while it is Poissonian for eV p� 2� (at T � 0), as shown by the shot noise calculation. Voltage dependence The analytic results presented above are useful in interpreting the numerical results in Fig. 9, where probability P M( , ) for symmetric junctions (R � 1) and R � 100, in the case of LL parameter g � 0 5. with no plasmon relaxation (� p � 0), is shown as a function of eV and M, that is the number of transported elec- trons to the right lead during such that during this time ! � �M Ic 10 electrons have passed to the right lead at eV EC� 2 . The peak position of the distribution of P M( , ) is roughly linearly proportional to the average particle flow !M , and the width is proportional to the shot noise but in a nonlinear manner. In a rough estimate, therefore, the ratio of the peak width to the peak posi- tion is proportional to the Fano factor. Two features are shown in the Fig. 9. First, the average particle flow (the peak position) runs with different slope when the bias voltage crosses new energy levels, i.e., at eV p� 2� and eV EC� 2 , that is consistent with the I V� study (compare Fig. 9,b with Fig. 3). Notice E /gC p� �� � 2�p for g � 0 5. . Second, the width of the distribution increases with increasing voltage, with different char- acteristics categorized by eV p� 2� and eV EC� 2 . Es- pecially in the bias regime eV EC� 2 in which several charge states participate to the charge transport, for the highly asymmetric case, the peak runs very fast while its width does not show noticeable increase. It causes the dramatic peak structure in the Fano factor around eV EC� 2 as discussed in Sec. 5. The deviation of the distribution of probability P M( , ) due to the nonequilibrium plasmons from its low voltage (equilibrium) counterpart is shown in the insets. Notice in the low-bias regime eV p� 2� , it fol- lows Eq. (63) for the symmetric case (Fig. 9,a, inset I), and the Poissonian distribution (65) for the highly 1538 Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 Jaeuk U. Kim, Mahn-Soo Choi, Ilya V. Krive, and Jari M. Kinaret asymmetric case (Fig. 9,b, inset I). The deviation is already noticeable at eV Ep C� �3 15� . for R � 1 (in- set II in Fig. 9,a), while it deviates strongly around eV EC� 2 for R � 100 (inset III in Fig. 9,a). Interaction strength dependence We have concluded in Sec. 5 that shot noise shows most dramatic behavior around eV EC� 2 due to inter- play between the nonequilibrium plasmons and the charge excitations. To see its consequence in FCS, we plot in Fig. 10 the probability P M( , ) as a function of the particle number M and the interaction parameter g for �� � �10 2 0e/I eV EC p( , ) with no plasmon re- laxation and with fast plasmon relaxation. The main message of Fig. 10,a is that the shot noise enhancement, i.e., the broadening of the distribution curve, is significant in the strong interaction regime with gradual increase with decreasing g. Fast plasmon relaxation consequently suppresses the average current and the shot noise dramatically as shown in Fig. 10,b implying the Fano factor enhancement is lost. Effec- tively, the probability distribution of P M( , ) for dif- ferent interaction parameters maps on each other al- most identically if the time duration is chosen such that !M equals for all g. 7. Conclusions We have studied different transport properties of a Luttinger-liquid single-electron transistor including average current, shot noise, and full counting statis- tics, within the conventional sequential tunneling approach. At finite bias voltages, the occupation probabilities of the many-body states on the central segment is found to follow a highly nonequilibrium distribution. The energy is transferred between the leads and the quantum dot by the tunneling electrons, and the elec- tronic energy is dispersed into the plasmonic collective Nonequilibrium plasmons and transport properties of a double-junction quantum wire Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 1539 40 30 20 10 0 M 0.5 1.0 1.5 2.0 eV/E C 0.8 0.6 0.4 0.2 0 a 1.0 0 0.5 1.0 1.5 2.0 eV/E C 40 30 20 10 M 0.8 0.6 0.4 0.2 0 b 1.0 Fig. 9. Probability P M( , )� with no plasmon relaxation ��p � 0) during the time (� � 10/Ic, where Ic is the particle current at eV EC� 2 with no plasmon relaxation (�p � 0): for symmetric junctions (R � 1) (a) and for a highly asym- metric junctions (R � 100) (b). Here g � 0 5. , N /G � 1 2, and T � 0. Inset shows cross-sectional image of P M( , )� (blue solid line) as a function of M and the reference dis- tribution function (a) Eq. (63) (magenta dashed line) and (b) the Poisson distribution Eq. (65) (red dashed), at eV EC� 05. (I), eV EC� 15. (II), and eV EC� 2 (III). 0 5 10 15 20 M P(M, )� 0.20 0.15 0.10 0.05 0.4 0.6 0.8 1.0 g a 0 5 10 15 20 M 0.4 0.6 0.8 1.0 g P(M, )� 0.20 0.15 0.10 0.05 b Fig. 10. The probability P M( , )� that M electrons have passed through the right junction during the time � � 10 0/I , where I0 is the particle current with no plasmon relaxation (�p � 0); with no plasmon relaxation ��p � 0) (a) and with fast plasmon relaxation (�p � 104) (b). Here eV E RC� �2 100, , N /G � 1 2 and T � 0. excitations after the tunneling event. In the case of nearly symmetric barriers, the distribution of the oc- cupation probabilities of the nonequilibrium plasmons shows impressive contrast depending on the interac- tion strength: In the weakly interacting regime, it is a complicated function of the many-body occupation configuration, while in the strongly interacting re- gime, the occupation probabilities are determined al- most entirely by the state energies and the bias volt- age, and follow a universal distribution resembling Gibbs (equilibrium) distribution. This feature in the strong interaction regime fades out with the increasing asymmetry of the tunnel-barriers. We have studied the consequences of these non- equilibrium plasmons on the average current, shot noise, and counting statistics. Most importantly, we find that the average current is increased, shot noise is enhanced beyond the Poisson limit, and full counting statistics deviates strongly from the Poisson distribu- tion. These nonequilibrium effects are pronounced es- pecially in the strong interaction regime, i.e. g � 0.5. The overall transport properties are determined by a balance between phenomena associated with nonequi- librium plasmon distribution that tend to increase noise, and involvement of several charge states and the ensuing correlations that tend to decrease noise. The result of this competition is, for instance, a nonmonotonic voltage dependence of the Fano factor. At the lowest voltages when charge can be trans- ported through the system, the plasmon excitations are suppressed, and the Fano factor is determined by charge oscillations between the two lowest zero-modes. Charge correlations are maximized when the tunnel- ing-in and tunneling-out rates are equal, which for symmetric junctions occurs at gate charge N /G � 1 2. At these gate charges the Fano factor acquires its low- est value which at low voltages is given by a half of the Poisson value, known as 1/2 suppression, as only two states are involved in the transport, at somewhat larger voltages increases beyond the Poisson limit as plasmon excitations are allowed, and at even higher voltages ex- hibits a local minimum when additional charge states are important. If the nonequilibrium plasmon effects are suppressed, e.g., by fast plasmon relaxation, only the charge excitations and the ensuing correlation ef- fects survive, and the Fano factor is reduced below its low-voltage value. The nonequilibrium plasmon effects are also suppressed in the noninteracting limit. Acknowledgments This work has been supported by the Swedish Founda- tion for Strategic Research through the CARAMEL con- sortium, STINT, the SKORE-A program, the eSSC at Postech, and the SK-Fund. J.U. Kim acknowledges par- tial financial support from Stiftelsen Fru Mary von Sydows, f�dd Wijk, donationsfond. I.V. Krive gratefully acknowledges the hospitality of the Department of Ap- plied Physies at Chalmers University of Technology. Appendix A: The transition amplitudes in the quantum dot In this appendix, we derive the transition ampli- tudes, Eqs. (15) and (16). As shown in Eq. (14), the zero-mode overlap of the QD transition amplitude is either unity or zero. Therefore, we focus on the over- lap of the plasmon states. It is enough to consider | { }| ( )| { } | , ,† � ! �n X n X X X D L R� � � 2 due to the symmetry between matrix elements of tun- neling-in and tunneling-out transitions, as in Eq. (15), | { }| ( )| { } | [| { }| ( )| { } |† † � ! � � ! �� �n X n n X nr r � � � 2 2B � ! � ! � { }| ( )| { } { }| ( )| { } ]†n X n n X nr rB B� � � � !2 2| { }| ( )| { } |†n X nrB � , (A.1) where r � �( ) denotes the right (left)-moving com- ponent, and the cross terms of oppositely moving components cancel out due to fermionic anti-commu- tation relations. The transition amplitudes at XL is identical to that at XR. For simplicity, we consider the case at XL � 0 only. The overlap elements of the many-body occupations � { }| , , ..., ,... |n n n nm1 2 and | { }n� ! � � � � � !| , ,..., ,...n n nm1 2 are | { }| ( )| { } | | | | |† � � ! � � ! � � 2n x n n nr m m m mB � C0 1 2 2 1 2 + , (A.2) where C Dm m m mb b� exp[ ( )]† with Dm i/ gmM� � is the bosonized field operator at an edge of the wire with open boundary conditions (see for instance Ref. 44). Here g is the interaction parameter, m is the mode index (and the integer momentum of it), and M is the number of transport sectors; if M � 1, the contribu- tions of the different sectors must be multiplied. The operators bm and bm † denote plasmon annihilation and creation and + is a high energy cut-off. Using the Baker–Haussdorf formula C D � � � m m m m a aa a m m m m� � exp[ ( )]† † e e e 2 2 , (A.3) and the harmonic oscillator states | ( ) ! | , | | ! , † n a n n a n n n ! � ! � 0 0 (A.4) 1540 Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 Jaeuk U. Kim, Mahn-Soo Choi, Ilya V. Krive, and Jari M. Kinaret one can show that, if n n5 �, | | | | | | ! ! ! ( )! | | ( ) � ! � � � � � � � � � � ; � n n n n n n n n n C D�2 2 22 e ; � � �E( , ; | | )n n n1 1 2 2D , (A.5) where E( , ; )x x z� is a degenerate hypergeometric func- tion defined by [77] E( , ; ) ! ( )! ( )! ( )! ( )! x x z z x x x x � � � � � � � � " � � � � � � � � 0 1 1 1 1# $ $ % & ' ' . (A.6) If n n�5 , the indices n and n� are exchanged in Eq. (A.5). The function E( , ; )x x z� is a solution of the equation z x z xz z� � � � � �2 0E E E( ) . By solving this differential equation with the proper normalization constant, one obtains E( , ; | | ) !( )! ! (| || |n n n n n n n Ln n n � � � � � � � � �1 1 2 22 D D�e ), where L ya b ( ) is the Laguerre polynomials [77]. In terms of the Laguerre polynomials, therefore, the transition amplitude (A.5) is written by | | | | ( ) ! ! / | | ( ) ( ) � ! � ; � � � n n gmM n n m m m gmM n n m mm m C 2 1e ; � � �� � � �� " # $ % & '� � L gmMn n n m m m ( ) | | 1 2 , (A.7) where n n n( ) ( , )� � �min and n n n( ) ( , )� � �max . We introduce a high frequency cut-off m k L /c F D~ � to cure the vanishing contribution due to e 1/gmM , 1 2 1 2 1 4 1 � � � + + e � 2 - � � �� � � �� / mg m m D c D c L L , (A.8) where the exponent is * � � ( )g /M1 1 . We arrive at the desired form of the on-dot transition matrix elements, | { }| ( { } |† )|� ! � � � �� � � �� ;n X n L La D D � + � 2 1 � � ; � � �� � � �� � � � � � �22 � 1 1 g mM n n L n n m m m n n m m m�� | | ( ) ( ) |! ! ( ) m mn g mM � � �� � � �� " # $ % & ' | , 1 2 � (A.9) Appendix B: Universal occupation probability In this appendix, we derive the universal distribu- tion of the occupation probability, which becomes Eq. (26) in the leading order approximation. Since the occupation probability of the plasmon many-body states is a function of the state energy in the strong interaction regime, we introduce the dimensionless energy n mn m m� � of the state with { } ( , ,..., ,...)n n n nm� 1 2 plasmon occupations. Exclud- ing the zero mode energy, therefore, the energy of the state { }n is given by E n E n nD D p({ }) ( )� � � with the state degeneracy D n( ), i.e., the number of many-body states n satisfying n mn m m� � , asymptotically follow- ing the Hardy–Ramanujan formula [78] D n / nn/( ) ( ).� e � 2 3 4 3 (B.1) We denote n sd by the corresponding dimensionless bias voltage eV n sd p� � . We obtain an analytic approximation to the occu- pation probability P n( ) at zero temperature by setting the on-dot transition elements in (15) to unity and considering the scattering-in and scattering-out pro- cesses for a particular many-body state { }n . The total scattering rates at zero temperature are given by a simple power-law Eq. (13), � .( ) ( )( )n m n m n / n m n /sd sd� � � � � � 2 2 � � � �.( )( )n m n / n m n /sd sd2 2 � � � � � �.( )( )n m n / n m n /sd sd2 2 � , (B.2) where the constant factor in Eq. (13) is set to unity. The master equation now reads � � � �t P n P m D m m n P n D m n m m ( ) [ ( ) ( ) ( ) ( ) ( ) ( )]� � . (B.3) To solve this master equation, we assume an ansatz of a power-law P m P n qn m n( ) ( ) .� (B.4) In the steady-state, master equation (B.3) in terms of this ansatz becomes ( ) ( )m n n / D m q m m sd n m n i � � � � � 2 � � � � � � � � �( ) ( ),n m n / D m m n n / sd sd 0 2 2 � (B.5) in which the sum in the LHS runs from m ni � �max( ,0 � n /sd 2), where max( , )x x� gives larger of x and x�. Nonequilibrium plasmons and transport properties of a double-junction quantum wire Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 1541 Using a saddle point integral approximation e e ef x f x f x x x dx dx( ) ( ) ( )( ) ,( (� �� 0 0 0 21 2 if ( ) , ( )� � �� �f x f x0 00 0 , (B.6) we solve Eq. (B.5) to obtain an equation for lnqn and find that for a large n, exp( ) exp( ( )),z /z n F n� (B.7) where z q C C n /sd � � | | , , ln � � * 1 2 and F n( ) is a slowly varying function of n for n �� 1, F n C n C n n sd( ) ln .� " # $ % & ' � � � � � � � 2 3 6 1 4 1 1 � � *� � (B.8) We assume an ansatz for the solution of z in Eq. (B.7), z n / F n n / F n K� � ( ) ( ) ( ) ( )ln ln2 2 F (B.9) and find a constant K which minimizes the correction term F. Putting this ansatz into Eq. (B.7) and solving the equation for F, we find at K � 0 8. , that the correc- tion term F is negligibly small (F � 0 01. ). Noting f n n / F n K( ) ( ) ( )� �ln 2 is almost li- near function in the regime of our interest (3 � n � 15), we linearize it around a valuen n� 0 (for instance,n0 9� ), f n f n n n f n( ) ( )( ) ( ),� � � 0 0 0 and solve Eq. (B.7) by ansatz (B.9) with above linearized form; � � � � � ln ln ( ) ( ) ( ) ( ) . q C n F n f n n f n f n nn � 1 2 0 0 0 0 (B.10) Apply � �m nP m qln ln[ ( )] [ ] to Eq. (B.10), and solve the integral equation for ln [ ( )]P n , ln ln[ ( )] [( ) ( )P n C dn n / F n n � � (� 2 � � �f n n f n f n n( ) ( ) ( ) ]0 0 0 0 . (B.11) The leading order approximation P n( )( )0 to the probability P n( ) of the average occupation from this integral results in Eq. (26) P n Z n Zn n n n n sd sd( )( ) ,0 1 3 2 1 1 3 2 1 � � � �� � e log (B.12) where Z is the partition function. A more accurate approximation P n( )( )1 can be de- rived by solving the integral Eq. (B.11) to a higher degree of precision, which yields P Z n n n Csd C n/ C ( )1 1 2 2 6 � � � � � � � � � �� � � �� � � �� � � �� � � � � n C n F n Kexp ( ) ln � � �" #$ % &' � � � � � � ;� 0 3 2 3 4 ; � � � � � � � � � � �� � � �� exp 4 6 2 2 2 2 � n n n / n n / n nsd sd n/n sd sd � � � � � � � � � � � � � � � � � � � � � �( )� 1 2 2 2 2 / sdn n , (B.13) where n0 9� is used and f n f n( ) ( )0 2 0� with minor correction is utilized for formal simplicity. Note the first term with normalization constant Z approaches P n( )( )0 in Eq. (B.12) as n/n sd - 0, noticing C /nsd� *� 2 1( ) . The integral equation (B.11) can be solved even without approximating on f n( ), with the expense of more cumbersome appearance of P n( ). 1. S. Tomonaga, Prog. Theor. Phys. 5, 544 (1950). 2. J.M. Luttinger, J. Math. Phys. 4, 1154 (1963). 3. F.D.M. Haldane, J. Phys. C14, 2585 (1981). 4. A.O. Gogolin, A.A. Nersesyan, and A.M. Tsvelik, Bosonization and Strongly Correlated Systems, Cam- bridge University Press, Cambridge (1998). 5. A.M. Chang, L.N. Pfeiffer, and K.W. West, Phys. Rev. Lett. 77, 2538 (1996). 6. S.J. Tans, M.H. Devoret, H. Dai, A. Thess, R.E. Smalley, L.J. Geerligs, and C. Dekker, Nature (Lon- don) 386, 474 (1997). 7. M. Bockrath, D.H. Cobden, J. Lu, A.G. Rinzler, R.E. Smalley, L. Balents, and P.L. McEuen, Nature (London) 397, 598 (1999). 8. Z. Yao, H.W.C. Postma, L. Balents, and C. Dekker, Nature (London) 402, 273 (1999). 9. T. Lorenz, M. Hofmann, M. Gr�ninger, A. Freimuth, G.S. Uhrig, M. Dumm, and M. Dressel, Nature (Lon- don) 418, 614 (2002). 10. G.L. Ingold and Y.V. Nazarov, in Single Charge Tun- neling, Coulomb Blockade Phenomena in Nanostructures, H. Grabert and M.H. Devoret (eds.), Plenum Press, New York (1992), chap. 2. 11. O.M. Auslaender, A. Yacoby, R. De Picciotto, K.W. Baldwin, L.N. Pfeiffer, and K.W. West, Phys. Rev. Lett. 84, 1764 (2000). 1542 Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 Jaeuk U. Kim, Mahn-Soo Choi, Ilya V. Krive, and Jari M. Kinaret 12. H.W.C. Postma, T. Teepen, Z. Yao, M. Grifoni, and C. Dekker, Science 293, 76 (2001). 13. C.L. Kane and M.P.A. Fisher, Phys. Rev. B46, 7268 (1992). 14. A. Furusaki, Phys. Rev. B57, 7141 (1998). 15. M. Thorwart, M. Grifoni, G. Cuniberti, H.W.C. Postma, and C. Dekker, Phys. Rev. Lett. 89, 196402 (2002). 16. S. H�gle and R. Egger, Europhys. Lett. 66, 565 (2004). 17. Y.V. Nazarov and L.I. Glazman, Phys. Rev. Lett. 91, 126804 (2003). 18. D.G. Polyakov and I.V. Gornyi, Phys. Rev. B68, 035421 (2003). 19. V. Meden, T. Enss, S. Andergassen, W. Metzner, and K. Schoenhammer, cond-mat/0403655 (2004). 20. T. Enss, V. Meden, S. Andergassen, X. Barnabe- Theriault, W. Metzner, and K. Sch�nhammer, cond- mat/0411310 (2004). 21. A. Braggio, M. Grifoni, M. Sassetti, and F. Napoli, Europhys. Lett. 50, 236 (2000). 22. A. Braggio, R. Fazio, and M. Sassetti, Phys. Rev. B67, 233308 (2003). 23. J.U. Kim, I.V. Krive, and J.M. Kinaret, Phys. Rev. Lett. 90, 176401 (2003). 24. J.U. Kim, J.M. Kinaret, and M.S. Choi, J. Phys.: Condens. Matter 17, 3815 (2005). 25. J.W. Wilkins, S. Hershfield, J.H. Davies, P. Hyld- gaard, and C.J. Stanton, Phys. Scripta T42, 115 (1992). 26. J.H. Davies, P. Hyldgaard, S. Hershfield, and J.W. Wilkins, Phys. Rev. B46, 9620 (1992). 27. S. Hershfield, J.H. Davies, P. Hyldgaard, C.J. Stanton, and J.W. Wilkins, Phys. Rev. B47, 1967 (1993). 28. A.N. Korotkov, Phys. Rev. B49, 10381 (1994). 29. E.V. Sukhorukov, G. Burkard, and D. Loss, Phys. Rev. B63, 125315 (2001). 30. M.J.M. De Jong and C.W.J. Beenakker, in NATO ASI Series Vol. 345, L. Sohn, L. Kouwenhoven, and G. Sch�n (eds.), Kluwer Academic Publishers, Dord- recht (1997), p. 225. 31. Y.M. Blanter and M. B�ttiker, Phys. Rep. 336, 1 (2000). 32. D. Tom�nek, private communication (2004). 33. I.O. Kulik and R.I. Shekhter, Sov. Phys. JETP 41, 308 (1975). 34. L.I. Glazman and R.I. Shekhter, J. Phys.: Condens. Matter 1, 5811 (1989). 35. R. Egger and H. Grabert, Phys. Rev. Lett. 77, 538 (1996); ibid. 80, 2255(E) (1998). 36. R. Egger and H. Grabert, Phys. Rev. B58, 10761 (1998). 37. J. Voit, Rep. Prog. Phys. 58, 1116 (1995). 38. C. Kane, L. Balents, and M.P.A. Fisher, Phys. Rev. Lett. 79, 5086 (1997). 39. R. Egger and A.O. Gogolin, Phys. Rev. Lett. 79, 5082 (1997). 40. M. Sassetti, G. Cuniberti, and B. Kramer, Solid State Commun. 101, 915 (1997). 41. M. Sassetti and B. Kramer, Phys. Rev. B55, 9306 (1997). 42. M. Fabrizio and A.O. Gogolin, Phys. Rev. B51, 17827 (1995). 43. S. Eggert, H. Johannesson, and A.E. Mattsson, Phys. Rev. Lett. 76, 1505 (1996). 44. A.E. Mattsson, S. Eggert, and H. Johannesson, Phys. Rev. B56, 15615 (1997). 45. K.A. Matveev and L.I. Glazman, Phys. Rev. Lett. 70, 990 (1993). 46. K.A. Matveev and L.I. Glazman, Physica B189, 266 (1993). 47. A. Braggio, M. Sassetti, and B. Kramer, Phys. Rev. Lett. 87, 146820 (2001). 48. F. Cavaliere, A. Braggio, J.T. Stockburger, M. Sas- setti, and B. Kramer, Phys. Rev. Lett. 93, 036803 (2004). 49. S. Kogan, Electronic Noise and Fluctuations in Solids, Cambridge University Press, Cambridge (1996). 50. C. Beenakker and C. Sch�nenberger, Phys. Today 56, 37 (2003). 51. R. Deblock, E. Onac, L. Gurevich, and L.P. Kouwen- hoven, Science 301, 203 (2003). 52. M.S. Choi, F. Plastina, and R. Fazio, Phys. Rev. Lett. 87, 116601 (2001). 53. M.S. Choi, F. Plastina, and R. Fazio, Phys. Rev. B67, 045105 (2003). 54. C.L. Kane and M.P.A. Fisher, Phys. Rev. Lett. 72, 724 (1994). 55. L. Saminadayar, D.C. Glattli, Y. Jin, and B. Etienne, Phys. Rev. Lett. 79, 2526 (1997). 56. R. De Picciotto, M. Reznikov, M. Heiblum, V. Uman- sky, G. Bunin, and D. Mahalu, Nature (London) 389, 162 (1997). 57. E. Comforti, Y.C. Chung, M. Heiblum, V. Umansky, and D. Mahalu, Nature (London) 416, 515 (2002). 58. L.Y. Chen and C.S. Ting, Phys. Rev. B43, 4534 (1991). 59. O.L. Bo and Y. Galperin, J. Phys.: Condens. Matter 8, 3033 (1996). 60. L.Y. Chen and C.S. Ting, Phys. Rev. B46, 4714 (1992). 61. L.Y. Chen, Phys. Rev. B48, 4914 (1993). 62. A.N. Korotkov, D.V. Averin, K.K. Likharev, and S.A. Vasenko, in Single Electron Tunneling and Mesoscopic Devices, H. Koch and H. L�bbrig (eds.), Springer, Berlin (1992), p. 45. 63. U. Hanke, Y.M. Galperin, and K.A. Chao, Phys. Rev. B50, 1595 (1994). 64. H. Birk, M.J.M. de Jong, and C. Sch�nenberger, Phys. Rev. Lett. 75, 1610 (1995). 65. A. Isacsson and T. Nord, Europhys. Lett. 66, 708 (2004). 66. A. Thielmann, M.H. Hettler, J. K�nig, and G. Sch�n, cond-mat/0501534 (2005). 67. H.A. Engel and D. Loss, Phys. Rev. Lett. 93, 136602 (2004). Nonequilibrium plasmons and transport properties of a double-junction quantum wire Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 1543 68. S.S. Safonov, A.K. Savchenko, D.A. Bagrets, O.N. Jouravlev, Y.V. Nazarov, E.H. Linfield, and D.A. Ritchie, Phys. Rev. Lett. 91, 136801 (2003). 69. T. Novotny, A. Donarini, C. Flindt, and A.P. Jauho, Phys. Rev. Lett. 92, 248302 (2004). 70. L.S. Levitov and G.B. Lesovik, JETP Lett. 58, 233 (1993). 71. B.A. Muzykantskii and D.E. Khmelnitskii, Phys. Rev. B50, 3982 (1994). 72. H. Lee, L.S. Levitov, and A.Y. Yakovets, Phys. Rev. B51, 4079 (1995). 73. B. Reulet, J. Senzier, and D.E. Prober, Phys. Rev. Lett. 91, 196601 (2003). 74. L.S. Levitov and G.B. Lesovik, Phys. Rev. B70, 115305 (2004), also at cond-mat/0111057. 75. C.W.J. Beenakker, M. Kindermann, and Y.V. Naza- rov, Phys. Rev. Lett. 90, 176802 (2003). 76. M.J.M. de Jong, Phys. Rev. B54, 8144 (1996). 77. I.S. Gradshteyn and I.M. Ryzhik, Table of Integrals, Series, and Products, Academic Press, London (1980). 78. G.H. Hardy and S. Ramanujan, Proc. London Math. Soc. 17, 75 (1918). 1544 Fizika Nizkikh Temperatur, 2006, v. 32, No. 12 Jaeuk U. Kim, Mahn-Soo Choi, Ilya V. Krive, and Jari M. Kinaret