cond-mat0303397/DDD.tex
1: %% LyX 1.2 created this file.  For more info, see http://www.lyx.org/.
2: %% Do not edit unless you really know what you are doing.
3: \documentclass[twocolumn,english,aps,prb,showpacs]{revtex4}
4: \usepackage[T1]{fontenc}
5: \usepackage[latin1]{inputenc}
6: \usepackage{graphicx}
7: \usepackage{amssymb}
8: 
9: \makeatletter
10: 
11: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% LyX specific LaTeX commands.
12: \providecommand{\LyX}{L\kern-.1667em\lower.25em\hbox{Y}\kern-.125emX\@}
13: %% Bold symbol macro for standard LaTeX users
14: \newcommand{\boldsymbol}[1]{\mbox{\boldmath $#1$}}
15: 
16: 
17: \usepackage{babel}
18: \makeatother
19: \begin{document}
20: 
21: \newcommand{\kt}[1]{\left|#1\right\rangle }
22: 
23: 
24: 
25: \newcommand{\br}[1]{\left\langle #1\right|}
26: 
27: 
28: 
29: \newcommand{\ph}{\hat{\phi }}
30: 
31: 
32: 
33: \newcommand{\pd}{\hat{\psi }^{\dagger }}
34: 
35: \newcommand{\ps}{\hat{\psi }}
36: 
37: 
38: 
39: \title{Dephasing in sequential tunneling through a double-dot interferometer
40: {\small }\\
41: }
42: 
43: 
44: \author{Florian Marquardt and C. Bruder}
45: 
46: 
47: \affiliation{Departement Physik und Astronomie, Universität Basel, Klingelbergstr.
48: 82, CH-4056 Basel}
49: 
50: 
51: \date{March 19th, 2003}
52: 
53: 
54: \email{Florian.Marquardt@unibas.ch}
55: 
56: \begin{abstract}
57: We analyze dephasing in a model system where electrons tunnel sequentially
58: through a symmetric interference setup consisting of two single-level
59: quantum dots. Depending on the phase difference between the two tunneling
60: paths, this may result in perfect destructive interference. However,
61: if the dots are coupled to a bath, it may act as a which-way detector,
62: leading to partial suppression of the phase-coherence and the reappearance
63: of a finite tunneling current. In our approach, the tunneling is treated
64: in leading order whereas coupling to the bath is kept to all orders
65: (using $P(E)$ theory). We discuss the influence of different bath
66: spectra on the visibility of the interference pattern, including the
67: distinction between {}``mere renormalization effects'' and {}``true
68: dephasing''.
69: \end{abstract}
70: 
71: \pacs{73.23.Hk, 71.38.-k, 03.65.Yz}
72: 
73: \maketitle
74: 
75: \section{Introduction}
76: 
77: The destruction of quantum-mechanical phase coherence due to coupling
78: of a system to an irreversible bath is a subject important not only
79: because of its connection to fundamental issues (the quantum measurement
80: process and the quantum-classical transition) but also because of
81: its role in the suppression of phenomena resulting from quantum interference
82: effects, such as those studied in mesoscopic physics (including Aharonov-Bohm
83: interference, weak localization and universal conductance fluctuations).
84: Recently, the field of mesoscopic physics in particular has seen a
85: revival of interest in these questions, due to surprising experimental
86: findings\cite{mohantywebb} concerning a possible saturation of the
87: weak-localization dephasing rate at low temperatures, that have not
88: yet been explained convincingly. Apart from investigations dealing
89: directly with the problem of weak localization in a disordered system
90: of interacting electrons, several toy models have been analyzed\cite{cedraschi,buettrevb,GZ_PB,OurABring,NagaevBuettikerHO,Guinea,GSZ_modelsLowTDeph,ImryInexistenceZPDP}
91: to answer the question whether decoherence at zero temperature is
92: possible at all, contrary to the expectations based on perturbation
93: theory. One of the difficulties faced by models involving discrete
94: levels consists in the fact that destruction of phase coherence for
95: a superposition of excited states of finite excitation energy is perfectly
96: possible even at zero temperature, due to spontaneous emission of
97: energy into the bath. It is only in the zero-frequency limit of the
98: linear response in a system with a continuous spectrum (relevant for
99: weak-localization and other equilibrium transport experiments) that
100: perturbation theory suggests in general a vanishing dephasing rate,
101: because then the perturbation does not supply energy to the system,
102: such that at $T=0$ the system is not able to leave a trace in the
103: bath, which is considered to be the prerequisite for decoherence.
104: 
105: Some questions of interest concerning dephasing, especially in connection
106: with mesoscopic systems and low temperatures, are the following ones:
107: How reliable is the simple classical picture of a phase being randomized
108: by fluctuating external noise\cite{sai}? In particular, what is the
109: meaning of the zero-point fluctuations of the bath in this picture,
110: as opposed to the thermal fluctuations dominating at frequencies lower
111: than the temperature? When do the former lead to {}``mere renormalization
112: effects'' and how is it possible to distinguish these from {}``true''
113: dephasing? Under which circumstances is the suppression of off-diagonal
114: terms in the reduced system density matrix itself already a good indicator
115: of dephasing? How reliable are simple arguments based on Golden Rule
116: and energy conservation, related to the connection between dephasing
117: and the trace left in the bath by the particle ({}``which-way''
118: detection)? When does perturbation theory fail qualitatively, what
119: is the influence of non-Markoffian behaviour? How does the dephasing
120: rate depend on the energy supplied by an external perturbation (frequencies
121: excited in linear response, bias voltage applied in a transport measurement)?
122: What is the influence of the Pauli principle in a system of degenerate
123: fermions? How strong are the qualitative differences in behaviour
124: resulting from different bath spectra?
125: 
126: In this work, we will present a model that is able to give insights
127: into most of these questions. 
128: 
129: %
130: \begin{figure}
131: \begin{center}\includegraphics[  height=5cm]{geometry2new.eps}\end{center}
132: 
133: 
134: \caption{\label{DDdoubleslitsetup}The double-dot {}``double-slit'' setup,
135: with a fixed phase difference $\varphi $ between the two paths and
136: under the influence of a fluctuating environment.}
137: \end{figure}
138: 
139: 
140: Our model represents a kind of mesoscopic double-slit setup. It consists
141: of two single-level quantum dots which are tunnel-coupled to two leads,
142: with a possible phase difference between the two interfering paths
143: (see Fig. \ref{DDdoubleslitsetup}). Due to destructive interference
144: (at $\varphi =\pi $), the tunneling current may be suppressed completely,
145: provided the two dot-levels are degenerate and the setup is symmetric
146: in the two interfering paths. Coupling the dots to a bath may partly
147: destroy the phase coherence and re-enable the electrons to go through
148: the setup. For a symmetric setup, with equal coupling strength between
149: the bath and each of the two dots, mere renormalization effects will
150: not be able to lift the destructive interference in this way. Thus,
151: a finite tunneling current may be taken as a genuine sign of dephasing.
152: This criterion for dephasing has been employed before in a model of
153: dephasing due to spin-flip transitions in first-order tunneling transport
154: through one or two dots\cite{koeniggefen}, as well as for cotunneling
155: through an Aharonov-Bohm ring coupled to a fluctuating magnetic flux\cite{OurABring}. 
156: 
157: The influence of phonons on sequential tunneling through two quantum
158: dots \emph{in series} has been studied experimentally in Ref.~\onlinecite{KouwenhouwenDD}.
159: There, inelastic transitions induced by piezoelectric coupling to
160: acoustic phonons in GaAs have been essential for obtaining a finite
161: current through the two off-resonant dot levels. This kind of setup
162: has been analyzed theoretically in Refs.~\onlinecite{StoofNazarov,brune,ziegler,BrandesKramer,aguado,Debald,KeilSchoellerRTRG}.
163: On the other hand, we will be analyzing tunneling through two dots
164: placed \emph{in parallel}. Early theoretical investigations of this
165: problem (without a fluctuating environment) include Refs.~\onlinecite{akera,shahbazyan}.
166: Recently, a parallel-dot tunneling setup has been realized experimentally
167: in Ref.~\onlinecite{KotthausDD}, with an emphasis on spectroscopy of the
168: {}``molecular states'' of the double-dot system (with inter-dot
169: tunneling present). In our model of an interference setup, we choose
170: to describe a situation without tunneling between the dots (but with
171: Coulomb-repulsion). In addition, we want to concentrate on interference
172: effects in the orbital motion and therefore consider the case of spin-polarized
173: transport. This model - in the absence of a fluctuating environment
174: - has been investigated previously in Ref.~\onlinecite{BoeseDD}. Other
175: recent theoretical works concerning tunneling through dots in a parallel
176: geometry have mostly investigated spin and Kondo physics \cite{LossShenjaDD,mahnsoo,BoeseHofstetterSchoeller},
177: but also dephasing by spin-flip transitions\cite{koeniggefen}. Some
178: works have treated the influence of phonons in tunneling interference
179: structures\cite{HauleBonca,key-2}, but no systematic discussion of
180: dephasing and the visibility of the interference pattern has been
181: given. Some while ago, dephasing by \emph{nonequilibrium} current
182: noise has been investigated experimentally\cite{key-1} and theoretically\cite{aleinerwhichpath}
183: in a setup with a single quantum-dot placed into one arm of an Aharonov-Bohm
184: interferometer.
185: 
186: Our analysis of dephasing in sequential tunneling through a double-dot
187: will take into account the system-bath coupling exactly, while we
188: treat the tunnel-coupling only in leading order. The presence of the
189: Fermi sea in the leads introduces some aspects related to the Pauli
190: principle and to the behaviour of systems with a continuous spectrum
191: that cannot be analyzed in simpler models of dephasing in discrete
192: systems coupled to a bath.
193: 
194: The work is organized as follows: After setting up the model (Sec.
195: \ref{sec:The-model}), we will present a qualitative discussion of
196: its main features (\ref{sec:Qualitative-discussion}). In particular,
197: we will discuss the relation between entanglement, dephasing and renormalization
198: effects. Subsequently, we derive a general formula for the tunneling
199: decay rate of an electron that has been placed on the two dots in
200: a symmetric superposition of states (Sec. \ref{sec:Decay-rate-and}).
201: This is done by building on the concepts of the $P(E)$ theory of
202: tunneling in a dissipative environment \cite{ingoldhabil,SchoenPE}.
203: Following this, we will evaluate the dependence of the tunneling rate
204: on the bias voltage and the bath spectra (Sec. \ref{sec:Evaluation-for-different}).
205: We will interpret the results in terms of {}``renormalization effects''
206: and {}``true dephasing'' (Sec. \ref{sec:Discussion-of-the}). Building
207: on these sections, we will finally derive a master equation for the
208: case of weak tunnel coupling (Sec. \ref{DDseqTunnelingSection}),
209: which allows us to calculate the sequential tunneling current as a
210: function of bias voltage, temperature, and phase difference (Sec.
211: \ref{sec:Evaluation-of-the}).
212: 
213: The most important results derived in this work are the following:
214: Equation (\ref{DDGammabasic}) is the general expression for the phase-dependent
215: tunneling decay rate in presence of the fluctuating environment. It
216: forms the basic input for the master equation (Eqs. (\ref{DDrhoppeq})-(\ref{DDrhopmEq})),
217: that describes sequential tunneling through the double-dot, where
218: the resulting current can be obtained from Eq. (\ref{DDcurrentexpression}).
219: The visibility of the interference pattern, which is defined by the
220: phase-dependence of the current, is given in Eq. (\ref{DDvisibilityI}).
221: It is connected with the visibility obtained from the phase-dependence
222: of the tunneling rate itself (Eqs. (\ref{eq:visDef}), (\ref{eq:visGamm})). 
223: 
224: 
225: \section{The model}
226: 
227: \label{sec:The-model}We consider a Hamiltonian describing two degenerate
228: single-level quantum dots, with respective single-particle states
229: $\left|+\right\rangle $ and $\left|-\right\rangle $ (spin is excluded
230: for simplicity, since we are interested in dephasing of the electronic
231: motion). Each of them is tunnel-coupled to two electrodes (with the
232: same strength for both dots), but involving a possible phase difference
233: between the tunnel amplitudes (see Fig. \ref{DDdoubleslitsetup}).
234: In addition, the potential difference between the two dots is given
235: by a fluctuating field $\hat{F}$, whose dynamics is derived from
236: a linear bath. It represents the fluctuations due to phonons or Nyquist
237: noise. The system-bath coupling strength is taken to be the same for
238: both dots, while the sign is opposite, such that the bath can distinguish
239: between an electron being on dot $\left|+\right\rangle $ or $\left|-\right\rangle $:
240: 
241: \begin{eqnarray}
242: \hat{H} & = & \epsilon (\hat{n}_{+}+\hat{n}_{-})+\hat{F}(\hat{n}_{+}-\hat{n}_{-})+U\hat{n}_{+}\hat{n}_{-}+\nonumber \\
243:  &  & \hat{H}_{L}+\hat{H}_{R}+\hat{H}_{B}+\hat{V}\label{DDhamiltonian}
244: \end{eqnarray}
245: 
246: 
247: Here $\hat{n}_{\pm }$ are the particle numbers on the two dots (equal
248: to $0$ or $1$). The bath Hamiltonian $\hat{H}_{B}$ describes a
249: set of uncoupled harmonic oscillators. It governs the dynamics of
250: the fluctuating potential $\hat{F}$, which is assumed to be linear
251: in the oscillator coordinates. The coupling between electron and bath
252: is of the form of the {}``independent boson model''\cite{mahan}.
253: For the case of exactly one electron on the double-dot, and in the
254: absence of tunneling, it corresponds to a spin-boson model with {}``diagonal
255: coupling''. In this model, no transition between different levels
256: is brought about by the bath, such that pure dephasing results. $U$
257: denotes the Coulomb repulsion energy, which we will take to be so
258: large that double-occupancy is forbidden. Note that the degeneracy
259: of the two dot-levels is important in the following: It is necessary
260: to ensure complete destructive interference at $\varphi =\pi $ (compare
261: also the discussion in Sec. \ref{DDseqTunnelingSection}). 
262: 
263: The terms $\hat{H}_{L}$ and $\hat{H}_{R}$ contain the energies of
264: the electrons in the left and right reservoirs: 
265: 
266: \begin{equation}
267: \hat{H}_{L(R)}=\sum _{k}\epsilon _{k}\hat{a}_{L(R)k}^{\dagger }\hat{a}_{L(R)k}\, .\end{equation}
268: The tunneling between the dots and the leads is described by $\hat{V}=\hat{V}_{L}+\hat{V}_{R}$,
269: with
270: 
271: \begin{equation}
272: \hat{V}_{R}=\sum _{k}t_{k}^{R}\hat{a}_{Rk}^{\dagger }(\hat{d}_{+}+e^{i\varphi }\hat{d}_{-})+h.c.\label{DDtunnelHam}\end{equation}
273: 
274: 
275: for the right junction, and 
276: 
277: \begin{equation}
278: \hat{V}_{L}=\sum _{k}t_{k}^{L}\hat{a}_{Lk}^{\dagger }(\hat{d}_{+}+\hat{d}_{-})+h.c.\label{eq:DDtunnelL}\end{equation}
279: 
280: 
281: for the left junction.
282: 
283: Here $\hat{d}_{\pm }$ are the annihilation operators for the two
284: dots ($\hat{n}_{\pm }=\hat{d}_{\pm }^{\dagger }\hat{d}_{\pm }$) and
285: the phase-factor of $e^{i\varphi }$ controls the interference between
286: tunneling events along either the upper or lower path. The tunneling
287: phase difference might be thought of as arising due to the Aharonov-Bohm
288: phase from a magnetic flux penetrating the region between the quantum
289: dots. 
290: 
291: Note that the tunneling matrix elements $t_{k}^{R(L)}$ are assumed
292: not to depend on the dot state $\kt{+}$ or $\kt{-}$ in our model.
293: This means that the dots are close enough such that they couple to
294: the same point on the lead electrodes, to within less than a Fermi
295: wavelength. Obviously there could be no appreciable interference effect
296: if the dots were separated by some larger distance (in which case
297: the $k$-dependence of matrix elements would be different for the
298: two states). The same idealized assumption underlies several similar
299: models (see, e.g., Refs.~\onlinecite{koeniggefen,BoeseDD,BoeseHofstetterSchoeller}).
300: The effect of an arbitrary dot separation has been discussed in some
301: detail in Ref.~\onlinecite{shahbazyan}. 
302: 
303: The present model, without the bath, has been analyzed previously
304: in Ref.~\onlinecite{BoeseDD} (see also Sec. IV.C of Ref.~\onlinecite{koeniggefen}).
305: There, an orbital type of Kondo effect was found in equilibrium, for
306: $\varphi =\pi $, when the level energy was below the chemical potential.
307: This arises because at $\varphi =\pi $ there are two states of the
308: double-dot that couple only to the left and the right lead, respectively
309: (denoted by $\kt{e}$ and $\kt{o}$ in the following). These degenerate
310: states form the pseudospin responsible for the Kondo effect. However,
311: that mechanism will be irrelevant for our analysis, as we consider
312: the transport situation where the (renormalized) level energy lies
313: between the chemical potentials of the left and the right lead. Therefore,
314: the degeneracy is effectively lifted by the bias voltage (which will
315: be assumed to be much larger than the tunneling rate), and only the
316: state coupling to the left lead would be occupied at $\varphi =\pi $. 
317: 
318: 
319: \section{Qualitative discussion}
320: 
321: \label{sec:Qualitative-discussion}In this and the following three
322: sections, we first analyze the escape of a single electron into the
323: right lead, where the electron is assumed to start out in a symmetric
324: superposition of the two dot levels, which has been formed by an electron
325: tunneling onto the dots from the left lead. In the situation without
326: any bath, this is the state $\left|e\right\rangle \equiv (\left|+\right\rangle +\left|-\right\rangle )/\sqrt{2}$. 
327: 
328: Without dephasing, the tunneling decay out of state $\kt{e}$ is made
329: impossible in the case of perfect destructive interference at $\varphi =\pi $,
330: while maximal constructive interference is present for $\varphi =0$.
331: It should be noted that the attribution of the phase factor to one
332: of the tunnel couplings represents a certain choice of gauge, which
333: affects the wave functions in the following discussion but none of
334: the physically observable quantities that are derived as a result
335: of the master equation in Section \ref{DDseqTunnelingSection}.
336: 
337: For simplicity, we will assume a zero-temperature situation throughout
338: the following qualitative discussion, with a bias $eV>0$ applied
339: between the two dots and the lead in such a way that the electron
340: is allowed to tunnel into the lead (see Fig. \ref{DDevpic}). In addition,
341: since we will describe the tunneling decay to the right, we will only
342: consider the coupling $\hat{V}_{R}$ to the right lead in this section
343: and drop the index $R$ for now.
344: 
345: %
346: \begin{figure}
347: \begin{center}\includegraphics[  height=5cm]{energydiagram.eps}\end{center}
348: 
349: 
350: \caption{\label{DDevpic}The ground state $\kt{\chi _{+}}$ ($\kt{\chi _{-}}$)
351: which the bath assumes in the presence of an electron on dot $\kt{+}$
352: ($\kt{-}$), shown schematically for a single bath oscillator (see
353: main text). After the electron has tunneled into the lead, $\kt{\chi _{-}}$
354: becomes a superposition of excited states (dashed), while the state
355: $\kt{\chi _{0}}$ represents the ground state of the bath in the new
356: potential.}
357: \end{figure}
358: 
359: 
360: Without the bath and for perfect constructive interference ($\varphi =0$),
361: the tunneling decay rate $\Gamma $ will take on its maximum value
362: of $2\Gamma _{0}$, with
363: 
364: \begin{equation}
365: \Gamma _{0}\equiv 2\pi D\left\langle \left|t_{k}\right|^{2}\right\rangle \, ,\label{DDbaretunnelrate}\end{equation}
366: 
367: 
368: where $D$ is the lead density of states at the Fermi energy, $\left\langle \left|t_{k}\right|^{2}\right\rangle $
369: is the angular average of $\left|t_{k}\right|^{2}$ at this energy.
370: The bias voltage $V$ does not enter in this case, as long as it is
371: positive (permitting decay). For $\varphi =\pi $, $\Gamma $ vanishes
372: due to perfect destructive interference. In general, we have:
373: 
374: \begin{equation}
375: \Gamma =\Gamma _{0}(1+\cos \varphi )\, .\label{eq:nobath}\end{equation}
376: 
377: 
378: If the bath is included in the description, the following happens:
379: 
380: First of all, the energy of a single extra electron on any of the
381: two dots will be renormalized from its initial value of $\epsilon $,
382: since the bath relaxes to a ground state of lower energy in presence
383: of the electron. We will assume that the value of $\epsilon $ has
384: been chosen exactly to compensate for this energy change, which is
385: given by $-\int _{0}^{\infty }d\omega \, \left\langle \hat{F}\hat{F}\right\rangle _{\omega }/\omega $
386: (see App. \ref{indepBosonApp}). Then, the energy of an electron on
387: the dot (and the bath in its new ground state) is the same as that
388: of the electron being in the lead, at the Fermi energy of $\epsilon _{F}\equiv 0$
389: (for $V=0$). 
390: 
391: Tunneling of an electron from the dots to the lead will not change
392: the bath state, but it will displace the origin of the harmonic oscillators
393: comprising the bath, since the coupling to $\hat{F}$ is switched
394: off ($\hat{n}_{+}-\hat{n}_{-}$ changes to zero). Therefore, the original
395: ground state of the bath (in presence of the electron) will become
396: a superposition of excited states in the new bath potential (in absence
397: of the electron; see Fig. \ref{DDevpic}). On the other hand, since
398: energy conservation has to be fulfilled with respect to the total
399: energy of the electrons and the bath before and after the tunneling
400: event, only those excited bath states can be reached whose energies
401: are not greater than $eV$, the energy supplied to the electron by
402: the bias voltage. This leads to the Coulomb-blockade type \emph{suppression}
403: of the tunneling rate at low bias voltages, for $\varphi =0$. Physically,
404: this effect is just the same as that described by Franck-Condon overlap
405: integrals evaluated between vibronic states for electronic transitions
406: in molecules. Qualitatively, this effect is independent of the interference
407: setup, since it already occurs for tunneling through a single dot
408: coupled to a bath.
409: 
410: In contrast, for the case of destructive interference ($\varphi =\pi $),
411: the bath may actually \emph{enhance} the tunneling rate from its initial
412: value of $0$, since it partly destroys the phase coherence that is
413: a presupposition for perfect interference. An electron coming from
414: the left lead will form the following entangled state with the bath,
415: instead of the symmetric superposition $\left|e\right\rangle =(\left|+\right\rangle +\left|-\right\rangle )/\sqrt{2}$:
416: 
417: \begin{equation}
418: (\left|+\right\rangle \left|\chi _{+}\right\rangle +\left|-\right\rangle \left|\chi _{-}\right\rangle )/\sqrt{2}\, .\label{DDentangled}\end{equation}
419: 
420: 
421: Here the states $\left|\chi _{\pm }\right\rangle $ denote the respective
422: ground states of the bath for a bath Hamiltonian given by $\hat{H}_{B}\pm \hat{F}$,
423: which are related to each other by a parity transformation (This also
424: means we assume by definition there to be no phase factor between
425: these states; e.g. both may be assumed to have real-valued positive
426: wave functions). Actually, the entangled state considered here will
427: be formed only if the electron is given barely enough energy to enter
428: the double-dot at all (i.e. chemical potential of the left lead infinitesimally
429: larger than the renormalized level position). Otherwise, excited bath
430: states may be created even at this step. These complications will
431: be taken care of in the complete discussion of the sequential tunneling
432: current (Section \ref{DDseqTunnelingSection}). There, it will turn
433: out that the tunneling decay rate derived in the following, based
434: on our physically motivated ansatz (\ref{DDentangled}), is exactly
435: the rate that enters the full master equation. Thus, we proceed with
436: the ansatz (\ref{DDentangled}) for the initial entangled state, in
437: order to calculate the rate for such an electron to tunnel into the
438: right lead.
439: 
440: The bath measures (to some extent) which dot the electron resides
441: on, such that the reduced system density matrix (for the electron
442: on the two dots) becomes mixed and its off-diagonal elements get suppressed
443: by the overlap factor $\left\langle \chi _{+}|\chi _{-}\right\rangle $.
444: Put differently, the phase factor between the two dot states in the
445: wave function of the electron (initially equal to $+1$) becomes uncertain.
446: Therefore, there is a finite probability of 
447: 
448: \begin{equation}
449: P_{o}=(1-\left\langle \chi _{+}|\chi _{-}\right\rangle )/2\label{eq:Podd}\end{equation}
450: 
451: 
452: to find the electron in the antisymmetric (odd) state $\left|o\right\rangle \equiv (\left|+\right\rangle -\left|-\right\rangle )/\sqrt{2}$.
453: At $\varphi =\pi $, where tunneling decay of the symmetric superposition
454: $\kt{e}$ is blocked due to destructive interference, the state $\kt{o}$
455: is allowed to decay into the lead, at the maximal rate of $2\Gamma _{0}$.
456: In this way, the interference-induced blockade of electron tunneling
457: is lifted by dephasing. 
458: 
459: However, this simple picture is true only for large bias voltages,
460: when energy conservation permits any final state of the bath after
461: the tunneling event. If the maximum energy supplied to the electron
462: is limited, the suppression discussed above (for the case of $\varphi =0$)
463: will apply again. In particular, if the bias voltage is turned to
464: zero, energy conservation only allows the state $\left|\chi _{0}\right\rangle $
465: to be reached, which is the ground state of the bath in the absence
466: of any electrons on the dots. Then, the tunneling rate is exactly
467: zero again, despite the fact that the reduced density matrix of the
468: electron may be mixed to a strong extent. The reason is the following:
469: When the overlap of the entangled state (\ref{DDentangled}) with
470: the state $\left|\chi _{0}\right\rangle $ is taken, the two overlap
471: factors $\left\langle \chi _{0}|\chi _{+}\right\rangle $ and $\left\langle \chi _{0}|\chi _{-}\right\rangle $
472: turn out to be the same, \emph{if} the coupling of the bath to the
473: two dots is symmetric (i.e. of equal strength, only of opposite sign),
474: which we have assumed in writing down the Hamiltonian, Eq. (\ref{DDhamiltonian}).
475: Therefore, the electronic state resulting from the projection of (\ref{DDentangled})
476: onto $\kt{\chi _{0}}$ is equal to the symmetric combination, whose
477: decay is forbidden. Thus, the combination of energy conservation and
478: Pauli blocking prevents a finite tunneling rate at zero bias voltage,
479: in spite of the mixed state of the electron coupled to the bath. In
480: this limit the entanglement between electron and bath only leads to
481: renormalization effects (such as the change in tunneling rate), but
482: not to genuine dephasing. If the coupling were asymmetric, then destructive
483: interference could be lost even without dephasing (merely due to renormalization),
484: just as it would be the case for initially asymmetric bare tunnel
485: couplings. That is why the asymmetric case is uninteresting for our
486: purposes of distinguishing renormalization effects from real dephasing. 
487: 
488: However, whether we are indeed able to claim that dephasing actually
489: vanishes in the limit of low bias voltages will depend on the behaviour
490: of the tunneling rate as a function of $V$ and on the comparison
491: of the cases $\varphi =0$ and $\varphi =\pi $. Here, the bath spectrum,
492: and, above all, its low-frequency properties, enter. In order to be
493: able to discuss $\Gamma (V,\varphi )$ quantitatively, we will make
494: use of the concepts of the $P(E)$ theory of tunneling in a dissipative
495: environment.
496: 
497: 
498: \section{Decay rate and connection to $P(E)$ theory}
499: 
500: \label{sec:Decay-rate-and}The tunneling rate $\Gamma $ will be calculated
501: using the standard Fermi Golden Rule, i.e. lowest order perturbation
502: theory in the bare tunneling rate $\Gamma _{0}$, but taking into
503: account exactly the bath coupling. In deriving the formula for $\Gamma $,
504: it turns out to be useful to assume that the bath oscillators do \emph{not}
505: get shifted in the tunneling event (unlike the qualitative considerations
506: from above), but it is rather the bath states which get displaced
507: (in the opposite direction). Obviously, this amounts to the same,
508: as long as we are interested only in overlap integrals of different
509: bath states after the event. To that end, we introduce the displacement
510: operator $\exp (i\hat{\phi })$, which transforms the bath ground
511: state of $\hat{H}_{B}$ into that of $\hat{H}_{B}+\hat{F}$. Here
512: $\hat{\phi }$ is a suitable hermitian operator that is linear in
513: the bosonic variables of the bath. In fact, this amounts to performing
514: the canonical transformation of the independent boson model\cite{mahan},
515: see Appendix \ref{indepBosonApp}. In terms of the two dot states
516: $+$ and $-$, we have $\hat{F}_{+}=\hat{F}$ and $\hat{F}_{-}=-\hat{F}$,
517: as well as $\ph _{+}=\ph $ and $\ph _{-}=-\ph $. The transformation
518: eliminates the system-bath coupling from the Hamiltonian, but gives
519: rise to modified dot operators $\hat{d}_{\pm }'=e^{\pm i\ph }\hat{d}_{\pm }$
520: in the transformed tunnel Hamiltonian $\hat{V}_{R}'$ (see Eq. (\ref{IBdtransform})). 
521: 
522: We will assume the tunnel-coupling to be sufficiently weak, such that
523: we can use lowest-order perturbation theory to calculate the tunneling
524: decay rate:
525: 
526: \begin{equation}
527: \Gamma =2\pi \sum _{f}\left|\left\langle f|\hat{V}'_{R}|i\right\rangle \right|^{2}\delta (E_{f}-E_{i})\, ,\end{equation}
528: 
529: 
530: where the initial state $\left|i\right\rangle $ is given by the configuration
531: involving the electron residing in the symmetric superposition on
532: the dots, the unperturbed Fermi sea in the lead and the bath in its
533: ground state $\left|i_{B}\right\rangle $. The bath ground state has
534: become independent of the position of the electron, due to the above-mentioned
535: transformation. At finite temperatures, an additional thermal average
536: over the initial bath state and the initial state of the electrons
537: in the lead has to be performed. The energies and eigenstates refer
538: to the Hamiltonian without tunnel coupling. Applying the new tunneling
539: Hamiltonian $\hat{V}'_{R}$ to the initial state, we obtain the following
540: expression:
541: 
542: \begin{eqnarray}
543: \Gamma =\pi \sum _{k,f_{B}}\left|t_{k}\right|^{2}(1-f(\epsilon _{k}+eV))\times  &  & \nonumber \\
544: \left|\left\langle f_{B}|e^{+i\hat{\phi }}+e^{i\varphi }e^{-i\hat{\phi }}|i_{B}\right\rangle \right|^{2}\delta (E_{f}^{B}-E_{i}^{B}+\epsilon _{k})\, , &  & \label{DDgammaanfang}
545: \end{eqnarray}
546: 
547: 
548: Here $f(\cdot )$ is the Fermi function (for chemical potential equal
549: to zero), and $E_{f,i}^{B}$ are the energies of the initial and final
550: bath states. The energy supplied to the bath is equal to the energy
551: lost by the electron (given by $-\epsilon _{k}$, since the renormalized
552: dot energy is zero). Following the usual derivation of the $P(E)$
553: theory \cite{ingoldhabil,SchoenPE}, we express the energy-conserving
554: $\delta $ function as an integral over time and also replace the
555: sum over lead states $k$ by an integral over the energy $E=-\epsilon _{k}$
556: supplied to the bath, finally yielding:
557: 
558: \begin{eqnarray}
559: \Gamma =\Gamma _{0}\int _{-\infty }^{+\infty }dE(1-f(-E+eV))\int _{-\infty }^{+\infty }\frac{dt}{2\pi }e^{iEt}\times  &  & \nonumber \\
560: \frac{1}{2}\left\langle (e^{-i\hat{\phi }(t)}+e^{-i\varphi }e^{i\hat{\phi }(t)})(e^{i\hat{\phi }}+e^{i\varphi }e^{-i\hat{\phi }})\right\rangle  &  & \label{DDGammavorher}
561: \end{eqnarray}
562: 
563: 
564: For the case of arbitrary temperature, the brackets denote a thermal
565: average over the initial bath state $\left|i_{B}\right\rangle $.
566: We introduce the definitions:
567: 
568: \begin{equation}
569: P_{(-)}(E)=\frac{1}{2\pi }\int _{-\infty }^{+\infty }dt\, e^{iEt}\, e^{\pm \left\langle \hat{\phi }(t)\hat{\phi }\right\rangle -\left\langle \hat{\phi }^{2}\right\rangle }\, .\label{DDPEDef}\end{equation}
570: 
571: 
572: This permits us to write down our final result for the tunneling decay
573: rate in terms of $P_{(-)}(E)$:
574: 
575: \begin{equation}
576: \Gamma =\Gamma _{0}\int _{-\infty }^{+\infty }dE\, (1-f(-E+eV))\, (P(E)+\cos (\varphi )P_{-}(E))\label{DDGammabasic}\end{equation}
577: 
578: 
579: The formula given here constitutes the basic expression for the decay
580: rate as a function of bias voltage and interference phase $\varphi $.
581: It represents the appropriate modification of Eq. (\ref{eq:nobath})
582: in presence of a bath. 
583: 
584: Note that for the slightly more general case of arbitrarily correlated
585: fluctuating potentials $\hat{F}_{+}$ and $\hat{F}_{-}$ attached
586: to the dots (i.e. an interaction of the form $\hat{F}_{+}\hat{n}_{+}+\hat{F}_{-}\hat{n}_{-}$),
587: the function $P_{-}(E)$ would contain the cross-correlator of the
588: associated phases $\hat{\phi }_{+}$ and $\hat{\phi }_{-}$, while
589: $P(E)$ would depend on the autocorrelator of $\hat{\phi }_{+}$ or
590: $\hat{\phi }_{-}$ (assumed to be the same, for the setup to remain
591: symmetric). In contrast to the model treated here, such an interaction
592: would also involve fluctuations of the sum of energies of the dot-levels.
593: However, they would only add to the renormalization effects mentioned
594: previously and do not contribute to dephasing by themselves, since
595: such fluctuations cannot distinguish between the two interfering paths.
596: 
597: By using the definitions
598: 
599: \begin{equation}
600: \gamma _{(-)}\equiv \Gamma _{0}\int dE\, (1-f(-E+eV))\, P_{(-)}(E)\, ,\label{DDgammaDefinition}\end{equation}
601: 
602: 
603: we can write
604: 
605: \begin{equation}
606: \Gamma =\gamma +\cos (\varphi )\gamma _{-}\, .\end{equation}
607: 
608: 
609: The strength of the dependence of $\Gamma $ on the phase $\varphi $
610: may be taken as a signature of phase coherence in our model. We define
611: the {}``visibility'' of the interference pattern in the usual way,
612: by
613: 
614: \begin{equation}
615: \upsilon \equiv (\Gamma _{max}-\Gamma _{min})/(\Gamma _{max}+\Gamma _{min}),\label{eq:visDef}\end{equation}
616: 
617: 
618: which is equal to the ratio
619: 
620: \begin{equation}
621: \upsilon =\frac{\gamma _{-}}{\gamma }\, .\label{eq:visGamm}\end{equation}
622: 
623: 
624: The visibility $\upsilon $ will be $1$ whenever the destructive
625: interference is perfect, and it is zero if there is no dependence
626: of $\Gamma $ on $\varphi $.
627: 
628: The effects of the bath on the decay rate are encoded in the functions
629: $P(E)$ and $P_{-}(E)$, whose general properties we will discuss
630: now. In the next section, we will evaluate them for different types
631: of bath spectra. 
632: 
633: As usual, the function $P(E)$ describes the probability (density)
634: that an electron will emit the energy $E$ into the bath while tunneling
635: into the lead. It is real, nonnegative and normalized to unity \cite{SchoenPE,ingoldhabil}. 
636: 
637: At large times $\left|t\right|\rightarrow \infty $, the correlation
638: function $\left\langle \hat{\phi }(t)\hat{\phi }\right\rangle $ in
639: the exponent of the integral (\ref{DDPEDef}) will decay towards zero,
640: for a continuous bath spectrum. This means that the integrand of $P(E)$
641: approaches the value of $z\equiv \exp (-\left\langle \hat{\phi }^{2}\right\rangle )$,
642: starting from $1$ at $t=0$. Therefore, $P(E)$ contains a {}``quasiparticle
643: $\delta $ peak'' of strength $z$ at $E=0$, if $z$ does not vanish.
644: It corresponds to the probability $z$ of having no energy transfer
645: at all from the electron to the bath (similar to the recoil-free emission
646: of a $\gamma $ ray by a nucleus inside a crystal, i.e. the Mössbauer
647: effect).
648: 
649: The function $P_{-}(E)$ in front of the $\cos (\varphi )$ term in
650: Eq. (\ref{DDGammabasic}) is different: The integrand of $P_{-}(E)$
651: will increase at large times, towards the value of $z$, starting
652: from $z^{2}$ at $t=0$. The function $P_{-}(E)$ is real-valued (because
653: of $\left\langle \hat{\phi }(t)\hat{\phi }\right\rangle =\left\langle \hat{\phi }\hat{\phi }(t)\right\rangle ^{*}$),
654: but it can become negative. Therefore, it cannot be interpreted as
655: a probability density, in contrast to $P(E)$. Its normalization is
656: given by:
657: 
658: \begin{equation}
659: \int dE\, P_{-}(E)=z^{2}\, .\end{equation}
660: 
661: 
662: If $z$ is nonzero, $P_{-}(E)$ also has a $\delta $ peak at $E=0$,
663: of weight $z$, just as $P(E)$. As a consequence, in the case of
664: destructive interference ($\varphi =\pi $), the tunneling rate $\Gamma $
665: at $V\rightarrow 0,\, T=0$ still vanishes even in the presence of
666: the bath, since the $\delta $ peaks contained in $P(E)$ and $P_{-}(E)$
667: cancel exactly in the integral (\ref{DDGammabasic}). The physical
668: reason for this coherence has been discussed at the end of the previous
669: section. 
670: 
671: In the case of constructive interference ($\varphi =0$), at $T=0$
672: and for $V\rightarrow 0$, the integration over $E$ will only capture
673: the $\delta $ peaks contained in $P_{(-)}(E)$, yielding $\Gamma =2z\Gamma _{0}$.
674: Thus, the tunneling rate is suppressed by the constant factor $z$
675: from its noninteracting value. However, this may be interpreted as
676: a mere renormalization of the effective tunnel coupling, since the
677: visibility $\upsilon $ of the interference pattern is still equal
678: to unity. In order to connect this result to the qualitative discussion
679: from above, we note that the overlap of the two different bath ground
680: states that are adapted to the absence or presence of an electron
681: on dot $\pm $, is given by: 
682: 
683: \begin{equation}
684: \left\langle \chi _{0}|\chi _{\pm }\right\rangle =\left\langle \chi _{0}\left|e^{\pm i\hat{\phi }}\right|\chi _{0}\right\rangle =\exp (-\left\langle \hat{\phi }^{2}\right\rangle /2)=z^{1/2}\, ,\end{equation}
685: 
686: 
687: Therefore, the magnitude squared of this overlap, that determines
688: the probability of tunneling without exciting any bath mode, is equal
689: to $z$.
690: 
691: On the other hand, for sufficiently large bias voltages (much larger
692: than the cutoff frequency of the bath spectrum), the normalization
693: conditions for $P_{(-)}(E)$ yield
694: 
695: \begin{equation}
696: \Gamma =\Gamma _{0}(1+z^{2}\cos (\varphi ))\, .\label{DDgammahighv}\end{equation}
697: 
698: 
699: The visibility is given by $\upsilon =z^{2}$. In this limiting case,
700: where the restrictions due to energy conservation and the Pauli principle
701: are no longer important, the tunneling rate $\Gamma $ at the point
702: $\varphi =\pi $ of destructive interference does not vanish. It takes
703: the value $\Gamma _{0}(1-z^{2})$, which is small if the effects of
704: the bath are weak ($z$ near to $1$) and is equal to one half the
705: ideal maximum value $2\Gamma _{0}$ for a bath that is sufficiently
706: strong to destroy phase coherence completely ($z=0$), leading to
707: an incoherent mixture of symmetric and antisymmetric states on the
708: two dots. In the latter case, the visibility vanishes (even for arbitrary
709: voltages), since then $P_{-}(E)$ is equal to zero, which makes $\Gamma $
710: independent of $\varphi $. This will be true for the Ohmic bath,
711: to be discussed in the next section. 
712: 
713: As explained above, the reduced density matrix of the electron on
714: the dots coupled to the bath predicts a finite probability of $P_{o}=(1-\left\langle \chi _{+}|\chi _{-}\right\rangle )/2$
715: to find the electron in the antisymmetric state if one starts out
716: from the symmetric superposition before coupling it to the bath. The
717: overlap factor of the bath states involved in this probability can
718: be expressed as 
719: 
720: \begin{equation}
721: \left\langle \chi _{+}|\chi _{-}\right\rangle =\left\langle \chi _{0}|(e^{-i\hat{\phi }})^{2}|\chi _{0}\right\rangle =z^{2}\, .\label{DDchipm}\end{equation}
722: 
723: 
724: Comparing with the result $\Gamma (\varphi =\pi )=\Gamma _{0}(1-z^{2})$
725: given above, it may be observed that the decay rate at sufficiently
726: large bias voltages is indeed determined directly by the probability
727: to find the electron in the state whose decay is not forbidden by
728: destructive interference (as has been argued already at the end of
729: the previous section, near Eq. (\ref{eq:Podd})). It is only in this
730: limiting case, where an arbitrary amount of energy is available for
731: excitation of the bath, that the suppression of interference effects
732: in the transport situation is correctly deduced from the electron's
733: reduced density matrix in the presence of the bath. Formally, this
734: holds because the sum over final bath states $f_{B}$ in Eq. (\ref{DDgammaanfang})
735: is not restricted any more and corresponds to the insertion of a complete
736: set of basis states. Thus, one obtains, directly from Eq. (\ref{DDgammaanfang}):
737: 
738: \begin{equation}
739: \Gamma =\frac{\Gamma _{0}}{2}\left\langle \chi _{+}+e^{-i\varphi }\chi _{-}|\chi _{+}+e^{i\varphi }\chi _{-}\right\rangle \, ,\label{eq:highvexplain}\end{equation}
740: 
741: 
742: which reduces to Eq. (\ref{DDgammahighv}) when the overlaps are evaluated,
743: using Eq. (\ref{DDchipm}). Physically, the case of high bias voltage
744: corresponds to a kind of infinitely fast von Neumann projection measurement
745: that determines the state of the electron, revealing the fluctuations
746: due to the bath. In contrast, at low bias voltages (low energy supply),
747: a kind of {}``weak'' measurement is carried out that takes a longer
748: amount of time, such that only the low-frequency fluctuations of the
749: bath are important for dephasing.
750: 
751: 
752: \section{Evaluation for different bath spectra}
753: 
754: \label{sec:Evaluation-for-different}We will restrict the discussion
755: to $T=0$ at first. 
756: 
757: The simplest example for the bath is a single harmonic oscillator
758: of frequency $\omega $. This offers an approximate description of
759: the interaction with optical phonon modes ({}``Einstein model'').
760: In this case, $P(E)$ and $P_{-}(E)$ can be obtained easily by expanding
761: the exponential in a Taylor series and using $\left\langle \hat{\phi }(t)\hat{\phi }\right\rangle =\left\langle \hat{\phi }^{2}\right\rangle \exp (-i\omega t)$,
762: before the integration over time is performed. For $P(E)$, the resulting
763: series of $\delta $ peaks at harmonics of $\omega $ corresponds
764: to all possible processes where the electron emits any number $n$
765: of phonons into the bath while tunneling into the lead. The expression
766: for $P_{-}(E)$ is the same, apart from alternating signs in front
767: of the $\delta $ functions: 
768: 
769: \begin{equation}
770: P_{(-)}(E)=z\sum _{n=0}^{\infty }\frac{\left\langle \pm \hat{\phi }^{2}\right\rangle ^{n}}{n!}\delta (E-n\omega )\, .\end{equation}
771: 
772: 
773: Thus, every process involving the transfer of an even number of quanta
774: to the bath will not ruin the destructive interference at $\varphi =\pi $,
775: since the corresponding contributions from $P(E)$ and $P_{-}(E)$
776: cancel in Eq. (\ref{DDGammabasic}). This is because the coupling
777: between electron and bath is of the type $(\hat{n}_{+}-\hat{n}_{-})\hat{F}$,
778: which gives a different sign of the interaction amplitude for a phonon
779: emission process, depending on the dot. Therefore, the amplitude of
780: emission of an \emph{even} number of phonons will \emph{not} depend
781: on the dot, it is insensitive to the state of the electron, and the
782: amplitudes of the electron tunneling from $\left|+\right\rangle $
783: and $\left|-\right\rangle $ will still interfere destructively.
784: 
785: In contrast, emission processes involving an odd number of quanta
786: introduce a negative sign for an electron starting in state $\left|-\right\rangle $,
787: {}``detecting'' the path (or rather, the initial state) of the electron
788: and interfering \emph{constructively} with the processes from $\left|+\right\rangle $.
789: This lifts the destructive interference and makes $\Gamma \neq 0$
790: at $\varphi =\pi $. However, below the frequency $\omega $ of the
791: oscillator, destructive interference at $\varphi =\pi $ is still
792: perfect since no quantum can be emitted, while the magnitude of $\Gamma $
793: at $\varphi =0$ is renormalized by the factor $z$, as has been discussed
794: above in general for the limiting case $V\rightarrow 0$. The same
795: holds true for any bath with a finite excitation gap, at $T=0$. This
796: is shown in Figs. \ref{DDGamma0} and \ref{DDGammaPi}, to be discussed
797: in the next section.
798: 
799: We now pass on to arbitrary bath spectra. At first, we will cover
800: the case $z\neq 0$ ({}``weak baths''), when we can apply perturbation
801: theory to discuss the behaviour of $P_{(-)}(E)$ at low energy transfers
802: $E$ (and, consequently, that of $\Gamma $ at low voltages). A Taylor-expansion
803: of the exponent in Eq. (\ref{DDPEDef}) yields:
804: 
805: \begin{eqnarray}
806: P_{(-)}(E)=\frac{z}{2\pi }\sum _{n=0}^{\infty }\frac{1}{n!}\int _{-\infty }^{+\infty }dt\, e^{iEt}\left[\pm \left\langle \hat{\phi }(t)\hat{\phi }\right\rangle \right]^{n} &  & \nonumber \\
807: =z\sum _{n=0}^{\infty }\frac{(\pm 1)^{n}}{n!}(\left\langle \hat{\phi }\hat{\phi }\right\rangle _{\omega }*\ldots *\left\langle \hat{\phi }\hat{\phi }\right\rangle _{\omega })(E) &  & \label{DDPfolding}
808: \end{eqnarray}
809: 
810: 
811: The repeated convolution product contains $n$ times the correlator
812: $\left\langle \hat{\phi }\hat{\phi }\right\rangle _{\omega }$, for
813: $n=0$ it is to equal $\delta (E)$, and the negative sign holds for
814: $P_{-}(E)$. 
815: 
816: For the following discussion, we prescribe the spectrum of the fluctuating
817: potential $\hat{F}$ to be a power-law in frequency $\omega $ (at
818: $T=0$), with exponent $s$:
819: 
820: \begin{equation}
821: \left\langle \hat{F}\hat{F}\right\rangle _{\omega }^{T=0}=2\alpha \omega _{c}\left(\frac{\omega }{\omega _{c}}\right)^{s}\theta (\omega _{c}-\omega )\theta (\omega ),\label{DDUUlaw}\end{equation}
822: 
823: 
824: The dimensionless parameter $\alpha $ characterizes the bath strength.
825: In order to be able to rely on perturbation theory, we have to ensure
826: $z>0$. Since $\left\langle \hat{\phi }\hat{\phi }\right\rangle _{\omega }=\left\langle \hat{F}\hat{F}\right\rangle _{\omega }/\omega ^{2}$,
827: the variance of the fluctuating phase, $\left\langle \hat{\phi }^{2}\right\rangle $,
828: will be finite only for $s>1$ (at $T=0$, otherwise $s>2$). In that
829: case, we have $z=\exp (-2\alpha /(s-1))$. This means the perturbative
830: analysis presented above is restricted to a super-Ohmic bath, $s>1$.
831: The case of the Ohmic bath will be discussed separately further below.
832: 
833: After keeping only terms up to second order in the expansion of $P_{(-)}(E)$
834: given in Eq. (\ref{DDPfolding}), we get
835: 
836: \begin{equation}
837: P(E)+P_{-}(E)=z(2\delta (E)+(\left\langle \hat{\phi }\hat{\phi }\right\rangle _{\omega }*\left\langle \hat{\phi }\hat{\phi }\right\rangle _{\omega })(E)+\ldots )\, ,\end{equation}
838: 
839: 
840: for the symmetric combination, that will determine the prefactor of
841: $1+\cos (\varphi )$ in the expression for $\Gamma $, Eq. (\ref{DDGammabasic}),
842: and
843: 
844: \begin{equation}
845: P(E)-P_{-}(E)=2z\left\langle \hat{\phi }\hat{\phi }\right\rangle _{E}+\ldots \end{equation}
846: 
847: 
848: for the antisymmetric combination (determining the prefactor of $1-\cos (\varphi )$).
849: Inserting these into (\ref{DDGammabasic}), using the power law for
850: $\left\langle \hat{\phi }\hat{\phi }\right\rangle _{\omega }=\left\langle \hat{F}\hat{F}\right\rangle _{\omega }/\omega ^{2}$
851: given by (\ref{DDUUlaw}), and performing the energy integrals, we
852: find, for sufficiently low voltages ($2\alpha (eV/\omega _{c})^{s-1}\ll s-1$): 
853: 
854: \begin{eqnarray}
855: \Gamma \approx \frac{\Gamma _{0}}{2}z\{(1+\cos (\varphi ))(1+\frac{\alpha ^{2}C_{s}}{(s-1)}\left(\frac{eV}{\omega _{c}}\right)^{2(s-1)})+ &  & \nonumber \\
856: (1-\cos (\varphi ))\frac{2\alpha }{s-1}\left(\frac{eV}{\omega _{c}}\right)^{s-1}\}\, . &  & \label{DDweakgamma}
857: \end{eqnarray}
858: 
859: 
860: The numerical prefactor $C_{s}$ is defined as $\int _{0}^{1}(y(1-y))^{s-2}dy$. 
861: 
862: From Eq. (\ref{DDweakgamma}), we see that the destructive interference
863: at $\varphi =\pi $ is perfect at $V=0$, but gets lifted when increasing
864: the bias voltage, with a power $V^{s-1}$. In contrast, the decay
865: rate $\Gamma $ at $\varphi =0$ starts out from the constant value
866: of $2z\Gamma _{0}$ and grows as $V^{2(s-1)}$. Therefore, the visibility
867: $\upsilon $ starts out at $1$ for $V=0$ but decreases as:
868: 
869: \begin{equation}
870: \upsilon \approx 1-\frac{4\alpha }{s-1}\left(\frac{eV}{\omega _{c}}\right)^{s-1}\, .\label{viasapprox}\end{equation}
871: 
872: 
873: For $s\downarrow 1$, the range in bias voltage $V$ where these approximate
874: expressions hold shrinks to zero (at constant $\alpha $ and $\omega _{c}$).
875: At $s=1$, i.e. for the Ohmic bath, the probability $z$ of not emitting
876: energy into the bath vanishes completely. As discussed above, this
877: means that there is no $\varphi $-dependence at all in $\Gamma $,
878: and, consequently, the visibility is zero at all bias voltages. Furthermore,
879: the tunneling rate vanishes for $eV\rightarrow 0$, even at $\varphi =0$.
880: This is the well-known Coulomb-blockade type of behaviour for tunneling
881: in presence of Ohmic dissipation \cite{devoret}. At higher bias voltages,
882: the blockade is removed and $\Gamma $ grows towards $\Gamma _{0}$.
883: The growth at low voltages is determined by the power-law behaviour
884: of $P(E)$, which rises as $c\omega _{c}^{-2\alpha }E^{2\alpha -1}$,
885: where the exponent is determined by the bath-strength rather than
886: the exponent $s=1$ of the bath spectrum. The dimensionless prefactor
887: $c$ must be found from the normalization condition for $P(E)$ and
888: depends only on $\alpha $ (and the type of cutoff in the bath spectrum).
889: Therefore, in the case of the Ohmic bath we have, at low $V$ and
890: $T=0$: 
891: 
892: \begin{equation}
893: \Gamma (V)=\Gamma _{0}\frac{c}{2\alpha }\left(\frac{eV}{\omega _{c}}\right)^{2\alpha }\, .\end{equation}
894: 
895: 
896: Finally, we briefly discuss the case of finite temperatures, $T>0$.
897: 
898: In that case, the variance of $\hat{\phi }$ is given by
899: 
900: \begin{equation}
901: \left\langle \hat{\phi }^{2}\right\rangle =\int _{0}^{\infty }d\omega \, \left\langle \hat{\phi }\hat{\phi }\right\rangle _{\omega }^{(T=0)}\coth \left(\frac{\omega }{2T}\right)\, ,\end{equation}
902: 
903: 
904: which yields
905: 
906: \begin{equation}
907: \left\langle \hat{\phi }^{2}\right\rangle \approx \left\langle \hat{\phi }^{2}\right\rangle ^{(T=0)}+4\alpha \left(\frac{T}{\omega _{c}}\right)^{s-1}\int _{0}^{\infty }\frac{y^{s-2}}{e^{y}-1}dy\, .\end{equation}
908: 
909: 
910: The approximation of extending the integral to infinity holds for
911: temperatures much smaller than the bath cutoff $\omega _{c}$. This
912: formula gives the temperature-dependence of the renormalization factor
913: $z=\exp \left(-\left\langle \hat{\phi }^{2}\right\rangle \right)$.
914: The second integral diverges for $s\leq 2$, because $z=0$ for these
915: cases, in contrast to $T=0$ where $z=0$ only for $s\leq 1$. Again,
916: this results in complete absence of the interference effect in the
917: tunneling rate $\Gamma (V,\varphi )$ (because $P_{-}(E)$ vanishes).
918: It may seem surprising that an infinitesimally small temperature can
919: yield such a drastic qualitative change (for $1<s\leq 2$), compared
920: to the zero-temperature case, since the difference should be observable
921: only at very large times $t\gg 1/T$. However, it must be remembered
922: that our analysis is carried out for the limit $\Gamma _{0}\rightarrow 0$,
923: where the average decay time of the given state is inifinitely large.
924: In other words, the limits $T\rightarrow 0$ and $\Gamma _{0}\rightarrow 0$
925: do not commute for such relatively strong baths. At finite $\Gamma _{0}$,
926: the transition from one to the other case should turn out to be smooth,
927: but this goes beyond the present analysis. 
928: 
929: Apart from the change in $z$ with temperature, there are two other
930: important differences to the case $T=0$: First of all, even at $V\rightarrow 0$
931: the electron may emit energy into the bath, due to the thermal smearing
932: of the Fermi surface in the lead (lifting of Pauli blocking). Secondly,
933: it may now also absorb some energy during the tunneling process. Both
934: facts will, in general, lead to a finite tunneling decay rate at $\varphi =\pi ,\, V\rightarrow 0$
935: for any bath, where, at $T=0$, the rate had vanished in any case. 
936: 
937: We can approximate the visibility $\upsilon $ at $V\rightarrow 0$
938: and finite $T$ by using the expansion (\ref{DDPfolding}). Inserting
939: the resulting expressions for $\gamma _{(-)}$ (\ref{DDgammaDefinition})
940: into $\upsilon =\gamma _{-}/\gamma $, we obtain
941: 
942: \begin{equation}
943: \upsilon (T,V\rightarrow 0)\approx 1-4\int d\epsilon \, \left\langle \ph \ph \right\rangle _{\epsilon }f(\epsilon )\, .\end{equation}
944: 
945: 
946: We evaluate the integral for a power-law bath spectrum in the limit
947: $T\ll \omega _{c}$:
948: 
949: \begin{eqnarray}
950: \int d\epsilon \, \left\langle \ph \ph \right\rangle _{\epsilon }f(\epsilon ) & = & \nonumber \\
951: =\int _{0}^{\infty }d\epsilon \, \frac{\left\langle \ph \ph \right\rangle _{\epsilon }^{T=0}}{\sinh (\beta \epsilon )} &  & \nonumber \\
952: \approx 2\alpha \omega _{c}^{1-s}\int _{0}^{\infty }d\epsilon \, \frac{\epsilon ^{s-2}}{\sinh (\beta \epsilon )}\, . &  & 
953: \end{eqnarray}
954: 
955: 
956: This yields:
957: 
958: \begin{equation}
959: 1-\upsilon (T,V\rightarrow 0)\approx 32\, \alpha \, \left(\frac{T}{\omega _{c}}\right)^{s-1}(\frac{1}{2}-2^{-s})\Gamma (s-1)\zeta (s-1)\, ,\label{DDvisfiniteT}\end{equation}
960: 
961: 
962: where $\Gamma $ is the Euler gamma function, and $\zeta $ the Riemann
963: zeta function. Therefore, the decrease of the visibility with increasing
964: temperature $T$ (and $V\rightarrow 0$) is governed by the same power-law
965: as that for increasing bias voltage $V$ at $T=0$, see Eq. (\ref{viasapprox}).
966: 
967: 
968: \section{Discussion of the results}
969: 
970: \label{sec:Discussion-of-the}The following discussion relates to
971: the results obtained for $T=0$, that are plotted in the figures.
972: 
973: %
974: \begin{figure}
975: \begin{center}\includegraphics[  height=7cm]{arrayNew.eps}\end{center}
976: 
977: 
978: \caption{\label{DDbathspectra}The bath spectrum $\left\langle \hat{F}\hat{F}\right\rangle _{E}$
979: (bottom) and the resulting functions $P(E)$ (top) and $P_{-}(E)$
980: (middle), plotted vs. energy $E$, for different baths. Energies are
981: measured in units of the {}``bath cutoff'' $\omega _{c}$. Energy
982: axis is the same in all panels (starting at $E=0$, horizontal tick
983: distance: $1$); vertical tick distance in all panels is $0.5$. \emph{a}:
984: $s=1.5,\, \alpha =0.25$; \emph{b}: {}``acoustic phonons'', $s=3,\, \alpha =1$;
985: \emph{c}: {}``optical phonons'', Bath with gap; \emph{d}: $s=1,\, \alpha =0.25$;
986: \emph{e}: $s=1,\, \alpha =0.75$ (d,e are {}``Ohmic'' baths of different
987: strength, $z=0$)}
988: \end{figure}
989: 
990: 
991: %
992: \begin{figure}
993: \begin{center}\includegraphics[  height=7cm]{Gamm0NEW.eps}\end{center}
994: 
995: 
996: \caption{\label{DDGamma0}Decay rate $\Gamma $ as a function of bias voltage
997: $V$ for the case of \emph{constructive} interference ($\varphi =0$),
998: at $T=0$. Curves correspond to different bath spectra shown in Fig.
999: \ref{DDbathspectra}. Dashed lines correspond to approximation Eq.
1000: (\ref{DDweakgamma}). The initial Coulomb-blockade type suppression
1001: to a value of $\Gamma /2\Gamma _{0}=z$ ($z=0$ for the Ohmic bath
1002: d,e) is lifted with increasing bias voltage, saturating at $\Gamma /2\Gamma _{0}=(1+z^{2})/2$.
1003: Inset depicts energy diagram with definition of bias voltage for this
1004: situation.}
1005: \end{figure}
1006: 
1007: 
1008: %
1009: \begin{figure}
1010: \begin{center}\includegraphics[  height=7cm]{GammaPINEW.eps}\end{center}
1011: 
1012: 
1013: \caption{\label{DDGammaPi}Decay rate $\Gamma (V)$ for the case of \emph{destructive}
1014: interference $(\varphi =\pi )$, at $T=0$. Dashed lines refer to
1015: Eq. (\ref{DDweakgamma}). Due to dephasing, the decay rate becomes
1016: finite at finite voltages, saturating at $\Gamma /2\Gamma _{0}=(1-z^{2})/2$.
1017: For the Ohmic bath (d,e) the dependence is exactly equal to that for
1018: $\varphi =0$ (Fig. \ref{DDGamma0}).}
1019: \end{figure}
1020: %
1021: \begin{figure}
1022: \begin{center}\includegraphics[  height=7cm]{Vis.eps}\end{center}
1023: 
1024: 
1025: \caption{\label{DDVis}Visibility $\upsilon =(\Gamma _{\varphi =0}-\Gamma _{\varphi =\pi })/(\Gamma _{\varphi =0}+\Gamma _{\varphi =\pi })$
1026: as a function of bias voltage $V$ for different bath spectra (see
1027: Fig. \ref{DDbathspectra}). For the Ohmic bath (cases d,e) $\upsilon \equiv 0$.
1028: Dashed lines correspond to Eq. (\ref{viasapprox}). Inset illustrates
1029: change in interference pattern $\Gamma (\varphi )$ upon switching
1030: on the interaction with the bath.}
1031: \end{figure}
1032: 
1033: 
1034: In Fig. \ref{DDbathspectra}, several different types of bath spectra
1035: $\left\langle \hat{F}\hat{F}\right\rangle _{E}$ are shown. Cases
1036: (a),(b),(d) and (e) are power-laws of the form given in Eq. (\ref{DDUUlaw}),
1037: for a cutoff frequency of $\omega _{c}=1$. The last two (d,e) are
1038: of Ohmic type ($s=1,\, z=0$), which corresponds physically to gate
1039: voltage fluctuations due to Nyquist noise. Case (c) represents a bath
1040: with an excitation gap (for example optical phonons), with a spectrum
1041: given by an inverted parabola. In the limit of infinitely small spectral
1042: bandwidth, it would correspond to the single harmonic oscillator (Einstein
1043: mode) discussed above. Case (b), with a bath spectrum rising as $\omega ^{3}$,
1044: corresponds to the experimentally relevant case of piezoelectric coupling
1045: to acoustic phonons, which was determined to be the major inelastic
1046: mechanism in the experiments of Ref.~\onlinecite{KouwenhouwenDD} on double-dots
1047: in GaAs (see Ref.~\onlinecite{BrandesKramer} for a theoretical analysis
1048: deriving this spectrum for wavelengths larger than the dot distance).
1049: The spectra for the first three cases (a,b,c) have been chosen to
1050: give the same renormalization factor, $z=1/e$. The same figure shows
1051: the resulting functions $P(E)$ and $P_{-}(E)$. These have been obtained
1052: using the integral equation described in Refs.~\onlinecite{minnhagen,ingoldhabil}.
1053: We recall that the low-energy behaviour of $P(E)$ is given by $\left\langle \hat{\phi }\hat{\phi }\right\rangle _{E}=\left\langle \hat{F}\hat{F}\right\rangle _{E}/E^{2}$
1054: for the cases with $z\neq 0$, where perturbation theory may be applied.
1055: In case (c), the alternating signs of the different contributions
1056: to $P_{-}(E)$ may be observed, whose physical meaning has been explained
1057: above for the limiting case of the harmonic oscillator.
1058: 
1059: We now briefly mention some numerical estimates for the bath strengths
1060: as they may occur in experimental situations.
1061: 
1062: In GaAs, the lack of inversion symmetry leads to piezoelectric fields
1063: proportional to the lattice deformation, whose effect on electrons
1064: at low frequencies is much stronger than that of the usual deformation
1065: potential (where it is only the \emph{potential} that is proportional
1066: to the deformation). For the piezoelectric coupling\cite{gant} to
1067: acoustic phonons in GaAs, one finds (compare Ref.~\onlinecite{BrandesKramer})
1068: $\left\langle \hat{\phi }\hat{\phi }\right\rangle _{\omega }^{T=0}=W\, \omega /(c_{s}/d)^{2}$
1069: for $\omega \ll c_{s}/d$, where $c_{s}\approx 5\cdot 10^{3}\, m/s$
1070: is an estimate for the average velocity of longitudinal sound waves
1071: in GaAs, and $d$ denotes the distance between the quantum dots. We
1072: obtain $W=const\cdot (eh_{14}/4\pi )^{2}/(\hbar \rho c_{s}^{3})$,
1073: where $eh_{14}=1.4\, eV/nm$ is the single piezo-electric modulus
1074: in the cubic $T_{d}$ structure of $GaAs$ and $\rho =5.3\cdot 10^{3}\, kg/m^{3}$
1075: the mass density. The numerical constant is of order $1$ and accounts
1076: for the details of the sound wave dispersion relation as well as the
1077: orientation of the crystal axes with respect to the vector separating
1078: the quantum dots. Inserting these values, $W$ is found to be on the
1079: order of $0.01$. In order to obtain the renormalization factor $z$,
1080: the spectrum $\left\langle \hat{\phi }\hat{\phi }\right\rangle _{\omega }$
1081: must be integrated over all frequencies (see above), i.e. up to the
1082: cutoff frequency $\omega _{c}$. The effective cutoff frequency $\omega _{c}\propto c_{s}/d_{0}$
1083: is determined by the extent $d_{0}$ of the dot wave functions (for
1084: $d_{0}=100nm$ one obtains $\omega _{c}\sim 50\, GHz$). Given the
1085: present values, and assuming $d_{0}\approx d$, this leads to estimates
1086: for $\int \left\langle \hat{\phi }\hat{\phi }\right\rangle _{\omega }d\omega $
1087: on the order of $0.01$, yielding $z=\exp (-\left\langle \hat{\phi }^{2}\right\rangle )$
1088: near $1$. Note that the distance $d$ between the dots cancels in
1089: the estimate for $z$, as long as the cutoff frequency is assumed
1090: to be given by $\omega _{c}\propto c_{s}/d$. However, as $\omega _{c}$
1091: might be considerably larger than $c_{s}/d$ (if $d_{0}\ll d$), one
1092: could also obtain a $z$ that deviates more strongly from unity.
1093: 
1094: For the Ohmic bath, we may imagine the quantum dots placed inside
1095: a capacitor $C$ connected to a circuit of resistance $R$, such that
1096: the potential difference $2\hat{F}$ between the dots would be given
1097: by the fluctuating voltage drop across the capacitor. This leads to
1098: a bath spectrum $\left\langle \hat{F}\hat{F}\right\rangle _{\omega }^{T=0}=\pi (R/R_{Q})\hbar ^{2}\omega /(1+(RC\omega )^{2})$,
1099: with $R_{Q}=h/e^{2}$ the quantum of resistance. Therefore, the dimensionless
1100: coupling constant $\alpha $ introduced above would be equal to $\alpha =(\pi /2)R/R_{Q}$,
1101: which can have values both larger and smaller than $1$. 
1102: 
1103: Finally, for optical phonons, we use the Fröhlich interaction Hamiltonian
1104: (Ref. \onlinecite{mahan}) with a dimensionless Fröhlich coupling
1105: constant of $\alpha =0.07$ (GaAs) to obtain the rough estimate $\left\langle \hat{F}\hat{F}\right\rangle _{\omega }^{T=0}=\delta (\omega -\omega _{LO})\cdot (1meV)^{2}\cdot (100nm/d_{0})$,
1106: with $\omega _{LO}\approx 5\cdot 10^{13}Hz$. This yields a $z$ deviating
1107: from unity by about $10^{-3}$.
1108: 
1109: However, in the plots we have chosen $z=1/e$ for illustrative purposes.
1110: 
1111: The resulting behaviour of $\Gamma (\varphi ,V)$ at $T=0$, calculated
1112: from Eq. (\ref{DDGammabasic}), is shown in Figs. \ref{DDGamma0}
1113: and \ref{DDGammaPi}. In the case of constructive interference ($\varphi =0$,
1114: Fig. \ref{DDGamma0}), the decay rate for the {}``weak baths'' (a,b,c)
1115: starts out from $\Gamma /2\Gamma _{0}=z$ at $V=0$ and goes to $\Gamma /2\Gamma _{0}=(1+z^{2})/2$
1116: at $eV/\omega _{c}\gg 1$. The initial deviation from the constant
1117: value of $z$ at low voltages is given by the power-law $V^{2(s-1)}$
1118: contained in Eq. (\ref{DDweakgamma}). In contrast, the decay rate
1119: for the Ohmic bath (d,e) starts at $\Gamma =0$, rising with a power-law
1120: and saturating at a value of $\Gamma /2\Gamma _{0}=1/2$, corresponding
1121: to an equal admixture of odd and even states in the reduced density
1122: matrix of the electron coupled to the bath. For destructive interference
1123: ($\varphi =\pi $, Fig. \ref{DDGammaPi}), the behaviour of (a) and
1124: (b) at low voltages is given by $V^{s-1}$ (see Eq. (\ref{DDweakgamma})),
1125: while the decay rate of the Ohmic bath (d,e) remains the same as that
1126: for $\varphi =0$. In the special case (c) of the gapped bath, we
1127: observe perfect destructive interference up to the excitation threshold
1128: of the bath at $eV=\omega _{c}$, where $\Gamma (\varphi =\pi ,V)$
1129: increases in a stepwise manner for the first time, with the next increase
1130: at $eV=3\omega _{c}$. Note that, on the other hand, $\Gamma (\varphi =0,\, V)$
1131: increases at even multiples of the excitation gap. The difference
1132: comes about because it is only the emission of an odd number of phonons
1133: into the bath that reveals the location of the electron, as discussed
1134: above. This feature would be absent if the two dots were coupled to
1135: two independent baths, whereas the other qualitative properties would
1136: remain the same.
1137: 
1138: From the decay rates at $\varphi =0$ and $\varphi =\pi $, we may
1139: calculate the visibility $\upsilon $ of the {}``interference pattern''
1140: that is defined by the dependence of $\Gamma $ on $\varphi $. The
1141: result is shown in Fig. \ref{DDVis}. As we have noted before, the
1142: visibility is always zero for the Ohmic bath. On the other hand, for
1143: the {}``weak baths'', it is perfect (equal to $1$) at $V\rightarrow 0$,
1144: due to the perfect destructive interference, regardless of the suppression
1145: factor $z$ appearing in $\Gamma (\varphi =0)$. In general, the visibility
1146: decreases towards higher bias voltages before saturating at the limiting
1147: value of $z^{2}$. However, in contrast to intuitive expectation,
1148: the decrease may be nonmonotonous, i.e. the visibility of the interference
1149: effect may actually be enhanced by increasing the supply of energy
1150: available to the electron, although the decay rate $\Gamma $ always
1151: increases monotonously at any $V$. This is particularly striking
1152: in case (c), where the visibility drops down to zero in a certain
1153: range before rising again. The decrease down to the exact value of
1154: $0$ is related to the special choice of $\left\langle \hat{\phi }^{2}\right\rangle =1$
1155: ($z=1/e$), which gives equal strengths of the peak at $E=0$ and
1156: the first peak around $E=\omega _{c}$, which then are able to cancel
1157: in the integral $\gamma _{-}$ over $P_{-}(E)$ that is proportional
1158: to the visibility. However, the physical reason for a dip in visibility
1159: is rather generic: In that energy range, the decay rate $\Gamma $
1160: for $\varphi =\pi $ has already increased due to dephasing, while
1161: the blockade-type suppression of the value of $\Gamma $ for $\varphi =0$
1162: has not yet been lifted. This is a consequence of the even-odd effect
1163: discussed above. 
1164: 
1165: 
1166: \section{Sequential tunneling through the double-dot}
1167: 
1168: \label{DDseqTunnelingSection}Up to now, we have discussed in detail
1169: the influence of the bath on the tunneling decay rate of an electron
1170: which has been placed onto the two dots in the symmetric superposition.
1171: In order to complete the picture, we have to calculate the sequential
1172: tunneling current through such a double-dot interference setup. This
1173: will be done by deriving and solving a master equation for the reduced
1174: density matrix of the double-dot system, taking into account the system-bath
1175: coupling exactly, while the tunnel-coupling is treated in leading
1176: order. We are interested specifically in the nonlinear response, i.e.
1177: in how an increasing bias voltage helps to destroy the phase coherence.
1178: The tunneling rates calculated previously will serve as input to the
1179: master equation.
1180: 
1181: However, in order to facilitate the understanding of the results,
1182: we first turn to a qualitative description of the situation without
1183: the bath.
1184: 
1185: At $\varphi =\pi $, tunneling is completely blocked, since the left
1186: reservoir only couples to the even state $\kt{e}$, while the right
1187: reservoir couples to the antisymmetric (odd) superposition, $\kt{o}$.
1188: At $\varphi =0$, both reservoirs couple to $\kt{e}$, whereas $\kt{o}$
1189: is completely decoupled from the leads (compare the discussion in
1190: Ref.~\onlinecite{BoeseHofstetterSchoeller}). This means that a current
1191: may flow if $\kt{o}$ is empty. However, if $\kt{o}$ is filled, the
1192: current vanishes, because double-occupancy is forbidden in our model.
1193: Since there is no way to change the occupation of $\kt{o}$, the stationary
1194: density-matrix of the double-dot at $\varphi =0$ will be any convex
1195: combination of these two possibilities (at $T=0$, in the absence
1196: of other relaxation paths). At any value of $\varphi $ in between
1197: these extremes, there is always the state $\kt{\Psi }=(\kt{+}-e^{-i\varphi }\kt{-})/\sqrt{2}$,
1198: whose decay into the right lead is blocked by destructive interference.
1199: As there is a nonvanishing overlap between $\kt{\Psi }$ and the state
1200: $\kt{e}$ which is reached by tunneling from the left lead, one will
1201: observe an accumulation of population in $\kt{\Psi }$, until the
1202: current is blocked again. This argument holds at $T=0$, while at
1203: finite temperatures the electron can decay towards the left lead and
1204: make a new attempt. Therefore, in this simple picture, the stationary
1205: current at $T=0$ would be zero at any $\varphi $ except for $\varphi =0$,
1206: where it is undefined. 
1207: 
1208: However, one has to take into account that the coupling to the reservoirs
1209: does not only lead to decay but also to an effective tunnel coupling
1210: between $\kt{+}$ and $\kt{-}$. Although this cannot change the blockade
1211: of the current at $\varphi =\pi $ (leading only to an energy shift
1212: of $\kt{e}$ vs. $\kt{o}$), it does lift the blockade at other values
1213: of $\varphi $. This is because the blocked state $\kt{\Psi }$ is
1214: no longer stationary, such that an electron will not remain there
1215: forever. The degeneracy at $\varphi =0$ still remains. Therefore,
1216: in the ideal case without coupling to a bath, we expect the current
1217: to vanish at $\varphi =\pi $ and to rise towards a maximal amplitude
1218: near $\varphi =0$. According to the previous argument, at $T=0$
1219: this maximal amplitude will be determined by the effective tunnel-coupling
1220: between the dot states. 
1221: 
1222: Introducing the bath will then lead to renormalization effects and
1223: spoil the perfect destructive interference at higher values of the
1224: bias voltage (or temperature), qualitatively in the same way as it
1225: has been explained above. We will show that the actual visibility
1226: $\upsilon _{I}$ of the current interference pattern $I(\varphi )$
1227: is given by a monotonous function of the visibility $\upsilon $ introduced
1228: above for the tunneling rate (at symmetric bias). 
1229: 
1230: We start with the Hamiltonian that is obtained after applying the
1231: unitary transformation of the independent boson model (\ref{IBHprime})
1232: onto our Hamiltonian (\ref{DDhamiltonian}):
1233: 
1234: \begin{equation}
1235: \hat{H}'=\epsilon '(\hat{n}_{+}+\hat{n}_{-})+U'\hat{n}_{+}\hat{n}_{-}+\hat{H}_{B}+\hat{H}_{L}+\hat{H}_{R}+\hat{V}'\, \end{equation}
1236: 
1237: 
1238: Here $\epsilon '$ is the (renormalized) energy of the two states,
1239: which we will take to be $\epsilon '=0$ from now on. $U'$ is the
1240: interaction constant that involves both the Coulomb repulsion as well
1241: as the effective attractive interaction induced by the bath. We assume
1242: $U,U'\gg T,eV$, such that double-occupancy is forbidden. 
1243: 
1244: The term which we will treat as a perturbation is given by $\hat{V}'$,
1245: describing the tunneling to the left and the right leads in the presence
1246: of the bath. It is the transformed version of $\hat{V}$ (compare
1247: Eqs. (\ref{DDtunnelHam}) and (\ref{eq:DDtunnelL}) and Appendix \ref{indepBosonApp}),
1248: where the additional fluctuating phase factors $\exp (\pm i\ph )$
1249: have been introduced:
1250: 
1251: \begin{equation}
1252: \hat{V}'=\sum _{j=l,r}\sum _{\alpha =+,-}\hat{j}_{\alpha }\hat{d}_{\alpha }+h.c.\, ,\label{DDseqtunnVprime}\end{equation}
1253: 
1254: 
1255: where
1256: 
1257: \begin{eqnarray}
1258: \hat{l}_{\pm } & = & e^{\pm i\ph }\hat{l}\\
1259: \hat{l} & = & \sum _{k}t_{k}^{L}\hat{a}_{Lk}^{\dagger }\\
1260: \hat{r}_{+} & = & e^{+i\ph }\hat{r}\\
1261: \hat{r}_{-} & = & e^{-i\ph }e^{i\varphi }\hat{r}\\
1262: \hat{r} & = & \sum _{k}t_{k}^{R}\hat{a}_{Rk}^{\dagger }\, .\label{DDseqrDef}
1263: \end{eqnarray}
1264: 
1265: 
1266: As usual, the current through the device does not only depend on the
1267: rates for electrons to tunnel into and out of the dots, but also on
1268: the stationary state which the system assumes in the nonequilibrium
1269: situation, i.e. under an applied bias voltage.
1270: 
1271: We will now derive a master equation for the reduced density matrix
1272: $\hat{\rho }$ of the double-dot system, which contains the populations
1273: $\rho _{++},\rho _{--},\rho _{00}$ ({}``$0$'' denoting {}``no
1274: electron'') and the coherences $\rho _{+-}$ and $\rho _{-+}$ (with
1275: $\rho _{00}=1-\rho _{++}-\rho _{--}$, $\rho _{\alpha 0}=\rho _{0\alpha }=0$
1276: for $\alpha \neq 0$, and $\rho _{-+}=\rho _{+-}^{*}$). We cannot
1277: simply use the standard kind of master equation, since we have to
1278: deal with two degenerate levels $\kt{+}$ and $\kt{-}$, and it is
1279: important that a tunneling event may create a coherent superposition
1280: of $\kt{+}$ and $\kt{-}$ (for example the even state $\kt{e}$).
1281: Such a master equation - for degenerate levels - has also been employed
1282: in Ref.~\onlinecite{koeniggefen} (without coupling to the bath, and evaluated
1283: in the linear-response regime). The equation is different from that
1284: employed in the {}``orthodox'' theory of sequential tunneling, where
1285: no coherent superpositions are involved. Note that for a finite tunnel-coupling
1286: the levels could be treated as degenerate as long as their energetic
1287: distance is much smaller than the level-broadening due to tunneling.
1288: However, as we consider the limit $\Gamma _{0}\rightarrow 0$, we
1289: need to have exactly equal energies. Otherwise, the energy of the
1290: hole that is created in the left electrode would betray the dot state
1291: which the electron has entered, thus preventing any coherent superposition
1292: to form.
1293: 
1294: Given the initial reduced density matrix $\hat{\rho }(0)$, and assuming
1295: the state of the environment (bath and reservoirs) to be independent
1296: of the electronic state on the dot at $t=0$, we obtain the time-evolution
1297: $\hat{\rho }(t)$ by tracing over the environmental degrees of freedom
1298: ({}``$E$''):
1299: 
1300: \begin{eqnarray}
1301: \hat{\rho }(t) & = & tr_{E}[\hat{T}e^{-i\int _{0}^{t}ds\, \hat{V}'(s)}\hat{\rho }(0)\otimes \hat{\rho }_{E}\tilde{\hat{T}}e^{i\int _{0}^{t}ds\, \hat{V}'(s)}]\nonumber \\
1302:  & = & \hat{\rho }(0)-\nonumber \\
1303:  &  & \int _{0}^{t}dt_{1}\int _{0}^{t_{1}}dt_{2}\, tr_{E}[\hat{V}'(t_{1})\hat{V}'(t_{2})\hat{\rho }(0)\otimes \hat{\rho }_{E}+h.c.]\nonumber \\
1304:  &  & +\int _{0}^{t}dt_{1}\int _{0}^{t}dt_{2}\, tr_{E}[\hat{V}'(t_{1})\hat{\rho }(0)\otimes \hat{\rho }_{E}\hat{V}'(t_{2})]\, \, +\ldots 
1305: \end{eqnarray}
1306: 
1307: 
1308: Physically, by using factorized initial conditions, we neglect correlations
1309: between subsequent tunneling events which could be due to excitations
1310: in the electrodes or in the bath: Since the tunneling rate is very
1311: small, these excitations will have traveled away from the double-dot
1312: until the next event takes place. The entanglement between electron
1313: and bath (discussed in the previous sections) would preclude factorized
1314: initial conditions, if it were not treated indirectly in this approach
1315: (via the unitary transformation). Note that we do not have to make
1316: any secular approximation at this point, unlike the usual derivation
1317: of a master equation \cite{Blum}. It turns out that all contributions
1318: only depend on the time-difference $t_{1}-t_{2}$ anyway, because
1319: the dot levels are degenerate. Therefore, in the long-time limit $t\rightarrow \infty $,
1320: the integration over $(t_{1}+t_{2})/2$ results in a factor $t$,
1321: and the endpoints of the integrals over $t_{1}-t_{2}$ may be extended
1322: to $\infty $. This yields the desired master equation that will determine
1323: the stationary $\hat{\rho }$, as well as the current, in the limit
1324: of weak tunnel coupling. 
1325: 
1326: In the expectation values of products $\hat{V}'\hat{V}'$ only those
1327: contributions remain which combine $\hat{d}_{\alpha }\hat{j}_{\alpha }$
1328: (tunneling out of the dots) with $\hat{j}_{\beta }^{\dagger }\hat{d}_{\beta }^{\dagger }$
1329: (tunneling onto the dots):
1330: 
1331: \begin{eqnarray}
1332: \frac{d\hat{\rho }}{dt}=-\sum _{\alpha ,\beta ,j}\int _{0}^{\infty }ds\, \left\{ \hat{d}_{\alpha }(s)\hat{d}_{\beta }^{\dagger }\hat{\rho }\left\langle \hat{j}_{\alpha }(s)\hat{j}_{\beta }^{\dagger }\right\rangle +h.c.\right. &  & \nonumber \\
1333: \left.+\hat{d}_{\alpha }^{\dagger }(s)\hat{d}_{\beta }\hat{\rho }\left\langle \hat{j}_{\alpha }^{\dagger }(s)\hat{j}_{\beta }\right\rangle +h.c.\right\}  &  & \nonumber \\
1334: +\sum _{\alpha ,\beta ,j}\int _{-\infty }^{+\infty }ds\, \left\{ \hat{d}_{\alpha }(s)\hat{\rho }\hat{d}_{\beta }^{\dagger }\left\langle \hat{j}_{\beta }^{\dagger }\hat{j}_{\alpha }(s)\right\rangle +\right. &  & \nonumber \\
1335: \left.\hat{d}_{\alpha }^{\dagger }(s)\hat{\rho }\hat{d}_{\beta }\left\langle \hat{j}_{\beta }\hat{j}_{\alpha }^{\dagger }(s)\right\rangle \right\} \, . &  & 
1336: \end{eqnarray}
1337:  (Note that there is no minus sign from fermion operator re-ordering
1338: in this factorization of dot and reservoir part, as the reservoir
1339: fermion operators are dragged past an even number of dot operators;
1340: compare e.g. \cite{schoellerRTRG}; alternatively, it is also possible
1341: to define them as commuting operators, since there is no interaction
1342: between them). We get for the individual matrix elements (for brevity,
1343: the summation over $j=l,r$ is implied):
1344: 
1345: \begin{eqnarray}
1346: \dot{\rho }_{++} & = & -\rho _{++}\int _{-\infty }^{+\infty }ds\, \left\langle \hat{j}_{+}^{\dagger }(s)\hat{j}_{+}\right\rangle \nonumber \\
1347:  &  & +\rho _{00}\int _{-\infty }^{+\infty }ds\, \left\langle \hat{j}_{+}\hat{j}_{+}^{\dagger }(s)\right\rangle \nonumber \\
1348:  &  & -\rho _{-+}\int _{0}^{\infty }ds\, \left\langle \hat{j}_{+}^{\dagger }(s)\hat{j}_{-}\right\rangle -h.c.\, ,\label{DDrhoppeqGeneral}
1349: \end{eqnarray}
1350: 
1351: 
1352: \begin{eqnarray}
1353: \dot{\rho }_{+-} & = & -\rho _{+-}\int _{0}^{\infty }ds\, \left\langle \hat{j}_{+}^{\dagger }(s)\hat{j}_{+}\right\rangle \nonumber \\
1354:  &  & -\rho _{+-}\int _{0}^{\infty }ds\, \left\langle \hat{j}_{-}^{\dagger }\hat{j}_{-}(s)\right\rangle \nonumber \\
1355:  &  & +\rho _{00}\int _{-\infty }^{+\infty }ds\, \left\langle \hat{j}_{-}\hat{j}_{+}^{\dagger }(s)\right\rangle \nonumber \\
1356:  &  & -\rho _{++}\int _{0}^{\infty }ds\, \left\langle \hat{j}_{+}^{\dagger }\hat{j}_{-}(s)\right\rangle \nonumber \\
1357:  &  & -\rho _{--}\int _{0}^{\infty }ds\, \left\langle \hat{j}_{+}^{\dagger }(s)\hat{j}_{-}\right\rangle \, .\label{DDrhopmEqGeneral}
1358: \end{eqnarray}
1359: 
1360: 
1361: The equation for $\rho _{--}$ follows from that for $\rho _{++}$
1362: by interchanging indices $+$ and $-$. 
1363: 
1364: Now we have to evaluate environment correlators, such as the prefactor
1365: of $\rho _{++}$ in the second equation (e.g. for $j=r$):
1366: 
1367: \begin{equation}
1368: \left\langle \hat{r}_{+}^{\dagger }\hat{r}_{-}(s)\right\rangle =e^{i\varphi }\left\langle e^{-i\ph }e^{-i\ph (s)}\right\rangle \left\langle \hat{r}^{\dagger }\hat{r}(s)\right\rangle \, .\label{eq:rrphiphi}\end{equation}
1369: 
1370: 
1371: By introducing the bare tunneling rates $\Gamma _{R(L)0}=2\pi \, D_{R(L)}\left\langle \left|t_{k}^{R(L)}\right|^{2}\right\rangle $
1372: (compare Eq. (\ref{DDbaretunnelrate})), we get, using Eq. (\ref{DDseqrDef})
1373: (remember $\hat{r}$ \emph{creates} a reservoir electron):
1374: 
1375: \begin{equation}
1376: \left\langle \hat{r}^{\dagger }\hat{r}(s)\right\rangle =\frac{\Gamma _{R0}}{2\pi }\int d\epsilon \, (1-f_{R}(\epsilon ))\, e^{+i\epsilon s}\, .\end{equation}
1377: 
1378: 
1379: Here we have neglected any energy-dependence of the tunnel-coupling
1380: and electrode DOS, assuming the relevant voltages and temperatures
1381: to be sufficiently small (but see below). The bath correlator in (\ref{eq:rrphiphi})
1382: evaluates to $\exp (-\left\langle \ph \ph (s)\right\rangle -\left\langle \ph ^{2}\right\rangle )$,
1383: which can be expressed by using the definition (\ref{DDPEDef}) for
1384: $P_{-}(\omega )$. There, we have to set $s\mapsto -s$ because of
1385: the reversed order in the $\ph $-correlator:
1386: 
1387: \begin{equation}
1388: e^{-\left\langle \ph \ph (s)\right\rangle -\left\langle \ph ^{2}\right\rangle }=\int d\omega \, P_{-}(\omega )e^{i\omega s}\, .\end{equation}
1389: 
1390: 
1391: Therefore, we obtain:
1392: 
1393: \begin{eqnarray}
1394: \int _{0}^{\infty }ds\, \left\langle \hat{r}_{+}^{\dagger }\hat{r}_{-}(s)\right\rangle = &  & \nonumber \\
1395: e^{i\varphi }\frac{\Gamma _{R0}}{2}\int d\epsilon \, (1-f_{R}(\epsilon ))\, \tilde{P}_{-}^{*}(-\epsilon )\, , &  & \label{DDexamplerr}
1396: \end{eqnarray}
1397: 
1398: 
1399: with
1400: 
1401: \begin{eqnarray}
1402: \tilde{P}_{-}(\epsilon )=\frac{1}{\pi }\int d\omega \, P_{-}(\omega )\int _{0}^{\infty }ds\, e^{i(\epsilon -\omega )s}= &  & \nonumber \\
1403: P_{-}(\epsilon )+\frac{i}{\pi }\int d\omega \, \frac{P_{-}(\omega )}{\epsilon -\omega }\, . &  & 
1404: \end{eqnarray}
1405: 
1406: 
1407: The integral in the second line is understood as a principal-value
1408: integral. In order to abbreviate expressions like this, we introduce
1409: the following definitions for the effective in- and out-tunneling
1410: rates as well as the effective tunnel couplings generated by the electrodes:
1411: 
1412: \begin{eqnarray}
1413: \gamma _{L(-)} & \equiv  & \Gamma _{L0}\int d\epsilon \, (1-f_{L}(\epsilon ))\, P_{(-)}(-\epsilon )\label{DDgammaseconddef}\\
1414: \gamma _{L(-)}^{in} & \equiv  & \Gamma _{L0}\int d\epsilon \, f_{L}(\epsilon )\, P_{(-)}(\epsilon )\\
1415: \Delta _{L} & \equiv  & -\frac{\Gamma _{L0}}{\pi }\int _{-\infty }^{\Lambda }d\epsilon \, (1-f_{L}(\epsilon ))\, \int d\omega \, \frac{P_{-}(\omega )}{\epsilon +\omega }\\
1416: \tilde{\gamma }_{L-} & \equiv  & \gamma _{L-}[P\mapsto \tilde{P}]=\gamma _{L-}+i\Delta _{L}\, .\label{DDgammatildedef}
1417: \end{eqnarray}
1418: 
1419: 
1420: Analogous definitions hold for $L\mapsto R$. Eq. (\ref{DDgammaseconddef})
1421: is equivalent to the definition (\ref{DDgammaDefinition}) used for
1422: $\gamma _{(-)}$ in previous sections. Note that the effective tunnel
1423: coupling $\Delta _{L(R)}$ depends on $P_{-}$, because it arises
1424: from transitions between the states $\kt{+}$ and $\kt{-}$, via an
1425: intermediate lead state. In the expression for $\Delta _{L(R)}$,
1426: the energy-dependence of the density of states and the tunnel coupling
1427: to the reservoir electrode should be kept in order to have a convergent
1428: integral. We will take this into account by introducing an effective
1429: upper energy cutoff $\Lambda $ in the integral. Using these definitions,
1430: (\ref{DDexamplerr}) is equal to $\exp (i\varphi )\tilde{\gamma }_{R-}^{*}/2$.
1431: 
1432: One might wonder why the effective tunnel couplings $\Delta _{L(R)}$
1433: do depend on the occupation of electron states in the reservoirs.
1434: After all, in the non-interacting case, it is possible to calculate
1435: such a change of the effective single-particle Hamiltonian prior to
1436: filling in the electron states. Alternatively, in a calculation that
1437: already takes into account occupation factors, there would be two
1438: contributions which add up to an integral that does not depend on
1439: the Fermi function. However, we consider the interacting case $U=\infty $,
1440: such that (even without the bath) one of these contributions is missing
1441: (since it would involve intermediate states with double occupancy).
1442: The resulting logarithm is analogous to that which appears in the
1443: Kondo problem. This effective tunnel coupling has also been discussed
1444: in Ref.~\onlinecite{BoeseDD}, for the case without a bath. There, the upper
1445: cutoff $\Lambda $ was provided by the Coulomb coupling $U$, since
1446: for higher energies double-occupancy is no longer forbidden and the
1447: non-interacting case takes over (where two contributions arise that
1448: cancel each other). If we take the limit $U\rightarrow \infty $,
1449: then $\Lambda $ will be set by a cutoff in the tunnel matrix elements
1450: (or the electron reservoir's density of states).
1451: 
1452: The general master equation for the reduced density matrix of the
1453: double-dot, derived in the limit of weak tunnel coupling but arbitrary
1454: electron-bath coupling, follows by inserting the definitions (\ref{DDgammaseconddef})-(\ref{DDgammatildedef})
1455: into Eqs. (\ref{DDrhoppeqGeneral}) and (\ref{DDrhopmEqGeneral}):
1456: 
1457: \begin{eqnarray}
1458: \dot{\rho }_{++} & = & -\rho _{++}(\gamma _{L}+\gamma _{R})\nonumber \\
1459:  &  & +\rho _{00}(\gamma _{L}^{in}+\gamma _{R}^{in})\nonumber \\
1460:  &  & -\frac{\rho _{-+}}{2}(e^{i\varphi }\tilde{\gamma }_{R-}+\tilde{\gamma }_{L-})-h.c.\, ,\label{DDrhoppeq}
1461: \end{eqnarray}
1462: 
1463: 
1464: \begin{eqnarray}
1465: \dot{\rho }_{--} & = & -\rho _{--}(\gamma _{L}+\gamma _{R})\nonumber \\
1466:  &  & +\rho _{00}(\gamma _{L}^{in}+\gamma _{R}^{in})\nonumber \\
1467:  &  & -\frac{\rho _{+-}}{2}(e^{-i\varphi }\tilde{\gamma }_{R-}+\tilde{\gamma }_{L-})-h.c.\, ,\label{DDrhommeq}
1468: \end{eqnarray}
1469: 
1470: 
1471: \begin{eqnarray}
1472: \dot{\rho }_{+-} & = & -\rho _{+-}(\gamma _{L}+\gamma _{R})\nonumber \\
1473:  &  & +\rho _{00}(e^{i\varphi }\gamma _{R-}^{in}+\gamma _{L-}^{in})\nonumber \\
1474:  &  & -\frac{\rho _{++}}{2}(e^{i\varphi }\tilde{\gamma }_{R-}^{*}+\tilde{\gamma }_{L-}^{*})\nonumber \\
1475:  &  & -\frac{\rho _{--}}{2}(e^{i\varphi }\tilde{\gamma }_{R-}+\tilde{\gamma }_{L-})\, .\label{DDrhopmEq}
1476: \end{eqnarray}
1477: 
1478: 
1479: The ingredients of the master equation obtained here may be interpreted
1480: as follows: 
1481: 
1482: One part of the right hand side corresponds to the unitary time-evolution
1483: generated by the effective tunneling Hamiltonian
1484: 
1485: \begin{equation}
1486: \hat{H}_{eff}^{T}=\frac{1}{2}(e^{i\varphi }\Delta _{R}+\Delta _{L})\kt{+}\left\langle -\right|+h.c.\, .\end{equation}
1487: 
1488: 
1489: Furthermore, the in-tunneling contributions in the equations for $\rho _{++}$
1490: and $\rho _{--}$ depend on $P(E)$, while that for $\rho _{+-}$
1491: is determined by $P_{-}(E)$, since it describes the creation of a
1492: coherent superposition of $\kt{+}$ and $\kt{-}$ (which is hindered
1493: by the bath). This term would be absent in the usual master equation.
1494: In particular, if $\gamma _{L-}^{in}\rightarrow \gamma _{L}^{in}$,
1495: which will be the case at $T=0$ for vanishing bias between the dots
1496: and the left electrode, an electron tunneling from the left lead will
1497: end up in the coherent superposition where $\rho _{+-}=\rho _{++}=\rho _{--}$.
1498: Taking into account that we are working in a transformed basis, this
1499: describes just the entangled state (\ref{DDentangled}), confirming
1500: the starting point of our earlier discussion. Note that the out-tunneling
1501: contribution for $\rho _{++}$ also depends on $\rho _{+-}$, for
1502: example. This reflects the fact that a superposition between the two
1503: states may be blocked from decaying into the lead, while each state
1504: separately can decay.
1505: 
1506: The stationary density matrix is obtained by demanding $d\hat{\rho }/dt=0$
1507: (and using the relations $\rho _{00}=1-\rho _{++}-\rho _{--}$ and
1508: $\rho _{-+}=\rho _{+-}^{*}$). This will give us the density matrix
1509: in zeroth order $\Gamma _{0}^{0}$ in the bare tunnel coupling, which
1510: we need to calculate the current in leading order $\Gamma _{0}^{1}$.
1511: 
1512: We can obtain the current from the contribution of the left electrode
1513: to the change $\dot{\rho }_{++}+\dot{\rho }_{--}$ in the double-dot
1514: occupation (i.e. keeping only terms that stem from the left electrode
1515: in the master equation). This is equal to the right-going current
1516: in the stationary limit:
1517: 
1518: \begin{eqnarray}
1519: \frac{I}{e}=(\dot{\rho }_{++}+\dot{\rho }_{--})_{L}= &  & \nonumber \\
1520: 2\rho _{00}\gamma _{L}^{in}-\gamma _{L}(\rho _{++}+\rho _{--})-2\gamma _{L-}Re[\rho _{+-}]\, . &  & \label{DDcurrentexpression}
1521: \end{eqnarray}
1522: 
1523: 
1524: An alternative way of deriving the current would be to start from
1525: the general Meir-Wingreen formula\cite{meirwingreen} which expresses
1526: the current in terms of the exact Green's functions of the double
1527: dot, to be calculated in presence of the tunnel-coupling and the bath.
1528: This has been the approach of Ref.~\onlinecite{koeniggefen} for the case
1529: without the bath, and we have checked (\ref{DDcurrentexpression})
1530: to give the same result in that case.
1531: 
1532: 
1533: \section{Evaluation of the sequential tunneling current and the visibility}
1534: 
1535: \label{sec:Evaluation-of-the}In order to evaluate the current as
1536: a function of temperature $T$, bias voltage $V$ and phase difference
1537: $\varphi $, we will now specialize to the case of symmetric bias
1538: and left-right symmetric tunnel couplings ($\Gamma _{R0}=\Gamma _{L0}=\Gamma _{0}$).
1539: All essential features (in particular the perfect destructive interference
1540: in absence of the bath) are independent of this assumption. We will
1541: find that the current is symmetric under $\varphi \mapsto -\varphi $
1542: even for the nonlinear response considered here, due to the symmetry
1543: of the model (compare Ref.~\onlinecite{koeniggefen} for a systematic analysis
1544: of phase-locking in a variety of interference geometries).
1545: 
1546: %
1547: \begin{figure}
1548: \begin{center}\includegraphics[  height=7cm]{Ishape.eps}\end{center}
1549: 
1550: 
1551: \caption{\label{DDIshapeFigure}The current $I$ for different values of the
1552: visibility $\upsilon =\gamma _{-}/\gamma =0.8,\, 0.9,\, 0.99,\, 0.999,\, 0.9999$
1553: (from top to bottom). The limits $\varphi \rightarrow 0$ and $\upsilon \rightarrow 1$
1554: do not commute. Other parameters held fixed: $\lambda =e^{-\beta \mu }=0.2$
1555: and $\delta _{L}=\delta _{R}=-1$. }
1556: \end{figure}
1557: 
1558: 
1559: We find from Eqs. (\ref{DDgammaseconddef})-(\ref{DDgammatildedef}),
1560: using $f(\epsilon )=1-f(-\epsilon )$:
1561: 
1562: \begin{equation}
1563: \gamma _{R(-)}=\gamma _{L(-)}^{in}=\gamma _{(-)}\equiv \Gamma _{0}\int d\epsilon \, f(\epsilon -\mu )\, P_{(-)}(\epsilon )\, ,\end{equation}
1564: 
1565: 
1566: where $\mu =eV/2$ is the chemical potential of the left reservoir.
1567: This is definition (\ref{DDgammaDefinition}), with $eV$ replaced
1568: by $\mu =eV/2$ (since we deal with the symmetric bias case). Furthermore,
1569: we use the condition of detailed balance, $P_{(-)}(-E)=\exp (-\beta E)\, P_{(-)}(E)$
1570: (see, for example, Ref.~\onlinecite{ingoldhabil}), which leads to
1571: 
1572: \begin{equation}
1573: \gamma _{L(-)}=\gamma _{R(-)}^{in}=e^{-\beta \mu }\gamma _{(-)}\, .\end{equation}
1574: 
1575: 
1576: The effective tunnel couplings are still different (because of the
1577: different Fermi distributions):
1578: 
1579: \begin{equation}
1580: \Delta _{L(R)}=-\frac{\Gamma _{0}}{\pi }\int _{-\infty }^{\Lambda }d\epsilon \, f(-(\epsilon \mp \mu ))\int d\omega \, \frac{P_{-}(\omega )}{\epsilon +\omega }\, .\end{equation}
1581: 
1582: 
1583: The lower sign belongs to the right electrode. 
1584: 
1585: For the special case of $T=0$, electrons always enter from the left
1586: and go to the right, such that we have $\gamma _{L}=\gamma _{L-}=\gamma _{R}^{in}=\gamma _{R-}^{in}=0$
1587: and $\gamma _{R(-)}=\gamma _{L(-)}^{in}=\gamma _{(-)}$, with
1588: 
1589: \begin{equation}
1590: \gamma _{(-)}=\Gamma _{0}\int _{0}^{\mu }d\epsilon \, P_{(-)}(\epsilon )\, .\end{equation}
1591: 
1592: 
1593: The effective tunnel couplings are, at $T=0$:
1594: 
1595: \begin{equation}
1596: \Delta _{L(R)}=-\frac{\Gamma _{0}}{\pi }\int d\omega \, P_{-}(\omega )\, \ln \left[\frac{\Lambda +\omega }{|\mu \pm \omega |}\right]\, .\end{equation}
1597: 
1598: 
1599: Note that, without any bath present, $\Delta _{L(R)}$ will have a
1600: logarithmic singularity at $\mu \rightarrow 0$, for $T=0$. The upper
1601: cutoff $\Lambda $ will be given by the minimum of the Coulomb repulsion
1602: energy $U$ and the bandwidth of the reservoir's electronic energy
1603: band (or by some cutoff in the tunnel matrix elements). For the purposes
1604: of our discussion, we assume $\Lambda \gg \mu ,\omega $.
1605: 
1606: In the the limit of high bias voltages ($\omega \ll \Lambda ,\mu $),
1607: we obtain effective tunnel couplings whose magnitude goes as $z^{2}$
1608: and decreases logarithmically with increasing $\mu $:
1609: 
1610: \begin{equation}
1611: \Delta _{L}\approx \Delta _{R}\approx -\frac{\Gamma _{0}}{\pi }\ln \left[\frac{\Lambda }{\mu }\right]\int d\omega \, P_{-}(\omega )\, =-z^{2}\frac{\Gamma _{0}}{\pi }\ln \left[\frac{\Lambda }{\mu }\right]\, .\end{equation}
1612: 
1613: 
1614: By solving the master equation for the stationary density matrix and
1615: inserting the result into Eq. (\ref{DDcurrentexpression}), we obtain
1616: the expression for the current through the double dot in terms of
1617: all of the quantities mentioned previously. In general (at arbitrary
1618: $T$), it is found that the current may be written as the product
1619: of $\gamma $ with a dimensionless function of the phase difference
1620: $\varphi $ and the ratios $\upsilon =\gamma _{-}/\gamma $, $\delta _{L(R)}=\Delta _{L(R)}/\gamma $
1621: and $\beta \mu $:
1622: 
1623: \begin{equation}
1624: I=e\gamma \, I_{0}[\varphi ,\beta \mu ,\upsilon ,\delta _{L},\delta _{R}]\, .\end{equation}
1625: 
1626: 
1627: The complete expression for $I_{0}$ is very cumbersome, although
1628: it may be found analytically by straightforward solution of the master
1629: equation (it is listed for $T=0$ in Appendix \ref{DDseqappendix}).
1630: Therefore, let us first discuss the situation without coupling to
1631: a bath. In that case, we obtain 
1632: 
1633: \begin{equation}
1634: \delta _{L}=\delta _{R}\equiv \delta =-\frac{\Gamma _{0}}{\pi }\int _{-\infty }^{\Lambda }\frac{d\epsilon }{\epsilon }\, f(\mu -\epsilon )\, \end{equation}
1635:  and $\gamma =\gamma _{-}=\Gamma _{0}f(-\mu )$. The current turns
1636: out to be (with $\lambda \equiv e^{-\beta \mu }$):
1637: 
1638: \begin{equation}
1639: \frac{I}{e\gamma }=\frac{4(1-\lambda )(\delta ^{2}+\lambda )\cos ^{2}(\frac{\varphi }{2})}{3\delta ^{2}+2(1+\lambda +\lambda ^{2})+3\delta ^{2}\cos (\varphi )}\, .\label{DDidealcurrent}\end{equation}
1640: 
1641: 
1642: Several points should be noticed about this expression: Firstly, the
1643: destructive interference at $\varphi =\pi $ remains perfect regardless
1644: of temperature, because there are no current-carrying states at all.
1645: At zero temperature ($\lambda =0$), the maximal amplitude of the
1646: current is $I_{max}/e\gamma =2\delta ^{2}/(3\delta ^{2}+1)$, which
1647: vanishes when the effective tunnel coupling $\delta $ goes to zero.
1648: This has been explained above as a consequence of the possible transition
1649: into a current-blocking state, which can only be undone by the effective
1650: tunnel coupling. At finite temperatures ($\lambda >0$), the maximal
1651: current is nonzero even for $\delta \rightarrow 0$, where it approaches
1652: the value of $I_{max}/e\gamma =2\lambda (1-\lambda )/(1+\lambda +\lambda ^{2})$.
1653: This has a maximum at around $T\sim \mu $. It vanishes for larger
1654: temperatures as $\mu /T$, which is to be expected for tunneling through
1655: a localized level (decreasing derivative of the Fermi function). In
1656: addition, the shape of $I(\varphi )$ depends on $\delta $ and $\lambda $,
1657: with a sharper minimum at $\varphi =\pi $ in the case of larger $|\delta |$.
1658: In the limit of $\delta \rightarrow 0$, the current becomes a pure
1659: cosine. At finite temperatures (as well as for $\upsilon \neq 1$)
1660: the behaviour is similar, except for the finite amplitude of the current
1661: at $\delta \rightarrow 0$.
1662: 
1663: Now we turn to the situation including the bath. The general expression
1664: for the current is very lengthy, and we will omit it here. However,
1665: it turns out that the maximal and minimal current are functions merely
1666: of $\upsilon $ and $\lambda =e^{-\beta \mu }$, while they are independent
1667: of $\delta _{L,R}$.
1668: 
1669: The amplitude of the minimal current (at $\varphi =\pi $) is given
1670: by
1671: 
1672: \begin{equation}
1673: \frac{I(\varphi =\pi )}{e\gamma }=\frac{2(1+\lambda )(1-\lambda ^{2})(1-\upsilon ^{2})}{3(1+\lambda )^{2}+(1-\lambda )^{2}\upsilon ^{2}}\, ,\end{equation}
1674: while the maximal current (at $\varphi =0$) is
1675: 
1676: \begin{equation}
1677: \frac{I(\varphi =0)}{e\gamma }=\frac{2}{3}(1-\lambda )\, .\label{DDmaxcurrentgeneral}\end{equation}
1678: It should be noted that the expression (\ref{DDidealcurrent}) for
1679: the current in the ideal case seems to contradict this simple formula.
1680: However, that is because the limits $\varphi \rightarrow 0$ and $\upsilon \rightarrow 1$
1681: do not commute. This is shown in Fig. \ref{DDIshapeFigure}. It means
1682: that for $T=0$ and $\delta _{L,R}\rightarrow 0$ the maximal current
1683: calculated according to (\ref{DDmaxcurrentgeneral}), which is independent
1684: of $\delta _{L,R}$, and the {}``typical'' amplitude of the current
1685: ($\propto \delta _{L}^{2}$) may deviate strongly. The peculiar behaviour
1686: near $\varphi =0$ seems to be connected to the physical degeneracy
1687: of the case $\varphi =0,\, \upsilon =1$ which has been discussed
1688: above.
1689: 
1690: From these formulas, we obtain the visibility, defined in terms of
1691: the current:
1692: 
1693: \begin{equation}
1694: \upsilon _{I}\equiv \frac{I(\varphi =0)-I(\varphi =\pi )}{I(\varphi =0)+I(\varphi =\pi )}\, .\end{equation}
1695: 
1696: 
1697: It can be expressed entirely by the visibility $\upsilon $ defined
1698: previously in terms of the tunneling rates (Eqs. (\ref{eq:visDef}),
1699: (\ref{eq:visGamm})), as well as the temperature-dependent factor
1700: $\lambda =e^{-\beta \mu }$ ($\mu =eV/2$):
1701: 
1702: \begin{equation}
1703: \upsilon _{I}=\frac{2(1+\lambda +\lambda ^{2})\upsilon ^{2}}{3(1+\lambda )^{2}-(1+4\lambda +\lambda ^{2})\upsilon ^{2}}\, .\label{DDvisibilityI}\end{equation}
1704: 
1705: 
1706: This is a monotonous mapping of $\upsilon $ to the interval $[0,1]$,
1707: with only a weak dependence on $\lambda $. The other parameters $\delta _{L},\delta _{R}$
1708: only modify the amplitude and shape of the current pattern $I(\varphi )$.
1709: Therefore, all the statements about the visibility made in the previous
1710: discussion of the tunneling decay out of the symmetric superposition
1711: continue to hold up to this monotonous transformation (and with $eV$
1712: replaced by $\mu =eV/2$). In particular, at $T=0$, we have
1713: 
1714: \begin{equation}
1715: \upsilon _{I}=\frac{2\upsilon ^{2}}{3-\upsilon ^{2}}\, .\end{equation}
1716: 
1717: 
1718: %
1719: \begin{figure}
1720: \begin{center}\includegraphics[  height=7cm]{visAll.eps}\end{center}
1721: 
1722: 
1723: \caption{\label{DDvisIvsMu}The visibility $\upsilon _{I}$ of the pattern
1724: $I(\varphi )$, for piezoelectric coupling to acoustic phonons (b)
1725: (solid line) and for the optical phonon bath (c) (dashed line), plotted
1726: vs. $\mu =eV/2$, at different temperatures $T/\omega _{c}=0.01,\, 0.05,\, 0.1,\, 0.2,\, 0.4,\, 0.5$
1727: (top to bottom). Inset depicts energy diagram for tunneling in this
1728: situation.}
1729: \end{figure}
1730: The dependence of the visibility $\upsilon _{I}$ on the bias voltage
1731: $eV=2\mu $, the temperature $T$ and the bath spectrum is displayed
1732: in Fig. \ref{DDvisIvsMu}, for bath spectra of type (b) and (c). The
1733: decrease of $\upsilon _{I}$ at $\mu =0$ with increasing temperature
1734: $T$ in case (b) is well approximated by Eq. (\ref{DDvisfiniteT})
1735: for $\upsilon (T,V\rightarrow 0)$ (employing the relation $\upsilon _{I}=\upsilon ^{2}/(2-\upsilon ^{2})$
1736: for $\mu =0$). (The functions $P_{(-)}(E)$ for finite temperatures
1737: have been calculated numerically using the fast Fourier transform,
1738: from the defining equation (\ref{DDPEDef})).
1739: 
1740: Note that for bath spectra with $z=0$ (i.e. exponent $s\leq 1$ at
1741: $T=0$ and $s\leq 2$ at $T>0$) the visibility vanishes entirely
1742: (at any $V$), as has been explained in the previous sections. We
1743: have already pointed out that this picture is expected to change if
1744: one treats the tunnel-coupling to higher order. However, we have to
1745: leave this analysis for the future. One possible approach to a nonperturbative
1746: (but still approximate) treatment of both the tunnel-coupling and
1747: the system-bath coupling at the same time seems to be the numerical
1748: {}``real-time renormalization group'' scheme \cite{schoellerRTRG}.
1749: 
1750: 
1751: \section{Conclusions}
1752: 
1753: We have analyzed dephasing in tunneling through two parallel single-level
1754: quantum dots with a fluctuating energy difference between the dots.
1755: The disappearance of perfect destructive interference in a symmetric
1756: setup has been taken as a criterion for {}``genuine'' dephasing,
1757: as opposed to mere renormalization. The coupling to the bath has been
1758: taken into account exactly, via the {}``independent boson model''
1759: and the concepts of the {}``$P(E)$ theory'' of tunneling in a dissipative
1760: environment, while the tunnel coupling has been treated in leading
1761: order. 
1762: 
1763: We have discussed in detail the behaviour of the density matrix of
1764: a single electron that has been placed in a superposition of the two
1765: dot levels. The bath measures (to some extent) the position of the
1766: electron, such that the electron's density matrix becomes mixed. However,
1767: this allows direct conclusions about the {}``incoherent current''
1768: only in the limit of high bias voltages, corresponding to a fast {}``projection''
1769: measurement of the electron's state. For lower voltages, only the
1770: low-frequency part of the bath spectrum contributes to the lifting
1771: of destructive interference. Thus, for any {}``weak bath'', whose
1772: spectrum falls off fast towards low frequencies, the visibility of
1773: the interference effect becomes perfect in the limit of low bias voltages
1774: $V$ and temperatures $T$, when the energy supplied to the electron
1775: is vanishingly small. This is the case for a fluctuation spectrum
1776: $\propto \omega ^{s}$ with $s>1$ ($s>2$) for $T=0$ ($T>0$). The
1777: visibility may show a nonmonotonous behaviour as a function of bias
1778: voltage. For {}``stronger'' spectra (smaller exponent $s$), including
1779: the Ohmic bath ($s=1$), there is the well-known zero-bias anomaly
1780: (suppression of the tunneling current at low voltages), which affects
1781: equally both the cases of constructive and destructive interference.
1782: Therefore, the visibility vanishes exactly at any bias voltage in
1783: our approach, where the tunnel coupling has been treated only in leading
1784: order. Although there is always a suppression of the magnitude of
1785: the tunnel current for the case of constructive interference, this
1786: may be interpreted as a mere renormalization of the effective tunnel-coupling,
1787: since the perfect destructive interference is not affected and since
1788: it occurs even for a bath with an excitation gap. The full dependence
1789: of the sequential tunneling current $I(\varphi )$ on voltage, temperature,
1790: bath spectrum and phase difference $\varphi $ between the interfering
1791: paths has been derived by setting up a master equation for the state
1792: of the double-dot (which is special due to the degeneracy of dot levels). 
1793: 
1794: The major questions that have remained open in our analysis are related
1795: to the behaviour at stronger tunnel coupling. In particular, the perfect
1796: destructive interference may also be overcome by correlated tunneling
1797: of several particles (with an intermediate {}``virtual'' excitation
1798: of the bath), and this process will therefore contribute to dephasing,
1799: although it is expected to be suppressed strongly at low voltages
1800: and temperatures. Likewise, the visibility for the Ohmic bath (or
1801: other strong baths), which turns out to be zero in the present approximation,
1802: may be changed at low bias voltages and temperatures comparable to
1803: the tunneling rate. This will require other methods to analyze the
1804: competition between strong tunnel coupling and system-bath coupling.
1805: 
1806: \begin{acknowledgments}
1807: We thank J. König, H. Schoeller, D. Loss and H. Grabert for useful
1808: discussions. This work has been supported through the Swiss NSF and
1809: the Swiss NCCR for Nanoscience.
1810: \end{acknowledgments}
1811: \appendix
1812: 
1813: 
1814: \section{Independent boson model}
1815: 
1816: \label{indepBosonApp}For reference purposes, we describe here the
1817: canonical transformation employed in the independent boson model.
1818: See Ref.\onlinecite{mahan} for more details (concerning the case
1819: of at most a single particle). Consider a set of electronic levels
1820: $j$ that couple to bath operators $\hat{F}_{j}$ which are assumed
1821: to be linear in the coordinates (and momenta) of a bath of harmonic
1822: oscillators, $\hat{H}_{B}$:
1823: 
1824: \begin{equation}
1825: \hat{H}=\sum _{j}(\varepsilon _{j}+\hat{F}_{j})\hat{n}_{j}+\hat{H}_{B}\, .\label{indepbosonham}\end{equation}
1826: 
1827: 
1828: Here $\varepsilon _{j}$ is the unperturbed level energy and $\hat{n}_{j}=\hat{d}_{j}^{\dagger }\hat{d}_{j}$
1829: is the number of particles on level $j$. The fluctuating fields are
1830: characterized completely by their power spectra at $T=0$, 
1831: 
1832: \begin{equation}
1833: \left\langle \hat{F}_{l}\hat{F}_{j}\right\rangle _{\omega }^{T=0}\equiv \frac{1}{2\pi }\int _{-\infty }^{+\infty }dt\, e^{i\omega t}\, \left\langle \hat{F}_{l}(t)\hat{F}_{j}\right\rangle ^{T=0}\, .\end{equation}
1834: 
1835: 
1836: Here we will restrict ourselves to the case where the different variables
1837: commute, $[\hat{F}_{l},\hat{F}_{j}]=0$. As a consequence, the spectrum
1838: $\left\langle \hat{F}_{l}\hat{F}_{j}\right\rangle _{\omega }^{T=0}$
1839: is real-valued, but there may still be correlations. 
1840: 
1841: The most straightforward solution proceeds via a unitary transformation\cite{mahan}
1842: (essentially a gauge transformation). One introduces the fluctuating
1843: phases $\ph _{j}$, whose time-derivatives are given by the $\hat{F}_{j}$:
1844: 
1845: \begin{equation}
1846: \dot{\ph }_{j}\equiv i[\hat{H}_{B},\ph _{j}]=-\hat{F}_{j}\, .\label{IBphidef}\end{equation}
1847: 
1848: 
1849: The exponent generating the unitary transformation is defined as:
1850: 
1851: \begin{equation}
1852: \hat{\chi }=\sum _{j}\ph _{j}\hat{n}_{j}\, .\end{equation}
1853: 
1854: 
1855: Applying the transformation to the Hamiltonian in Eq. (\ref{indepbosonham})
1856: yields:
1857: 
1858: \begin{equation}
1859: \hat{H}'=e^{-i\hat{\chi }}\hat{H}e^{+i\hat{\chi }}=\sum _{j}\varepsilon _{j}\hat{n}_{j}-\sum _{lj}J_{lj}\hat{n}_{l}\hat{n}_{j}+\hat{H}_{B}\, .\label{IBHprime}\end{equation}
1860: 
1861: 
1862: The coupling between system and bath has been eliminated, resulting
1863: in an effective interaction between particles on the different levels,
1864: with:
1865: 
1866: \begin{equation}
1867: J_{lj}=\int _{0}^{\infty }d\omega \, \frac{\left\langle \hat{F}_{l}\hat{F}_{j}\right\rangle _{\omega }^{T=0}}{\omega }\, .\label{IBcoupling}\end{equation}
1868: 
1869: 
1870: The $J_{lj}$ are real-valued and independent of temperature. For
1871: $l=j$ they describe energy shifts of single-particle levels. The
1872: canonical transformation also changes the particle annihilation and
1873: creation operators,
1874: 
1875: \begin{equation}
1876: \hat{d}_{j}'=e^{-i\hat{\chi }}\hat{d}_{j}e^{+i\hat{\chi }}=e^{i\ph _{j}}\hat{d}_{j}\, ,\label{IBdtransform}\end{equation}
1877: 
1878: 
1879: and $\hat{d}_{j}'^{\dagger }=\hat{d}_{j}^{\dagger }e^{-i\ph _{j}}$.
1880: This will affect all Green's functions and, therefore, also the time-evolution
1881: of the single-particle density matrix. In addition, it becomes important
1882: if a tunneling part is added to the Hamiltonian, where the operators
1883: $\hat{d}_{j}^{(\dagger )}$ appear, such that they have to be transformed
1884: according to (\ref{IBdtransform}). However, since the phases $\hat{\phi }_{j}$
1885: and the particle operators $\hat{d}_{j}^{(\dagger )}$ commute (even
1886: at different times, when evolved according to $\hat{H}'$), the evaluation
1887: of Green's functions always splits into a part referring to the particles
1888: and a separate average over the bath operators. This is the major
1889: simplification brought about by the {}``diagonal coupling'' between
1890: system and bath.
1891: 
1892: 
1893: \section{Current expression for sequential tunneling through the double-dot}
1894: 
1895: \label{DDseqappendix}At $T=0$, for the symmetric situation, the
1896: current $I$ is given by $I=e\gamma \, I_{0}[\upsilon ,\delta _{L},\delta _{R}]$,
1897: with:
1898: 
1899: \begin{eqnarray}
1900: I_{0}[\upsilon ,\delta _{L},\delta _{R}] & = & 2\cdot [-\delta _{L}^{2}+(\upsilon ^{2}-1)(1+\delta _{R}^{2})+\nonumber \\
1901:  &  & 2\delta _{L}\delta _{R}(\upsilon ^{2}-1)\cos \varphi +\delta _{L}^{2}\upsilon ^{2}\cos ^{2}\varphi ]\cdot \nonumber \\
1902:  &  & [-3\delta _{L}^{2}+2\delta _{L}\delta _{R}\upsilon ^{2}+(1+\delta _{R}^{2})(\upsilon ^{2}-3)+\nonumber \\
1903:  &  & 2(\upsilon ^{2}(1+\delta _{L}^{2}+\delta _{R}^{2})+\delta _{L}\delta _{R}(\upsilon ^{2}-3))\cos \varphi +\nonumber \\
1904:  &  & \delta _{L}(\delta _{L}+2\delta _{R})\upsilon ^{2}\cos ^{2}\varphi ]^{-1}
1905: \end{eqnarray}
1906: 
1907: 
1908: \begin{thebibliography}{10}
1909: \bibitem{mohantywebb}P. Mohanty, E.~M.~Q. Jariwala, and R.~A. Webb, Phys. Rev. Lett.
1910: \textbf{78}, 3366 (1997).
1911: \bibitem{cedraschi}P. Cedraschi, V. V. Ponomarenko, and M. Büttiker, Phys. Rev. Lett.
1912: \textbf{84}, 346 (2000).
1913: \bibitem{buettrevb}P. Cedraschi and M. Büttiker, Phys. Rev. B \textbf{63}, 165312 (2001);
1914: Annals of Physics, \textbf{289}, 1 (2001).
1915: \bibitem{GZ_PB}A. D. Zaikin and D. S. Golubev, Physica B \textbf{280}, 453 (2000).
1916: \bibitem{OurABring}F. Marquardt and C. Bruder, Phys. Rev. B \textbf{65}, 125315 (2002).
1917: \bibitem{NagaevBuettikerHO}K. E. Nagaev and M. Büttiker, Europhys. Lett. \textbf{58}, 475 (2002).
1918: \bibitem{Guinea}F. Guinea, Phys. Rev. B \textbf{65}, 205317 (2002).
1919: \bibitem{GSZ_modelsLowTDeph}D.~S.~Golubev, G.~Schön, and A.~D.~Zaikin, cond-mat/0208548 (2002).
1920: \bibitem{ImryInexistenceZPDP}Y. Imry, cond-mat/0202044 (2002).
1921: \bibitem{sai}A. Stern, Y. Aharonov, and Y. Imry, Phys. Rev. A \textbf{41}, 3436
1922: (1990).
1923: \bibitem{koeniggefen}J. König and Y. Gefen, Phys. Rev. Lett. \textbf{86}, 3855 (2001);
1924: Phys. Rev. B \textbf{65}, 045316 (2002). 
1925: \bibitem{KouwenhouwenDD}T. Fujisawa et al., Science \textbf{282}, 932 (1998).
1926: \bibitem{StoofNazarov}T. H. Stoof and Y. V. Nazarov, Phys. Rev. B \textbf{53}, 1050 (1996).
1927: \bibitem{brune}Ph. Brune, C. Bruder, and H. Schoeller, Phys. Rev. B \textbf{56},
1928: 4730 (1997); Physica E \textbf{1}, 216 (1997).
1929: \bibitem{ziegler}R. Ziegler, C. Bruder, and H. Schoeller, Phys. Rev. B \textbf{62},
1930: 1961 (2000).
1931: \bibitem{BrandesKramer}T. Brandes and B. Kramer, Phys. Rev. Lett. \textbf{83}, 3021 (1999).
1932: \bibitem{aguado}R. Aguado and L. P. Kouwenhoven, Phys. Rev. Lett. \textbf{84}, 1986
1933: (2000).
1934: \bibitem{Debald}S. Debald, T. Brandes, and B. Kramer, Phys. Rev. B \textbf{66}, 041301
1935: (2002).
1936: \bibitem{KeilSchoellerRTRG}M. Keil and H. Schoeller, cond-mat/0205308 (2002).
1937: \bibitem{akera}H. Akera, Phys. Rev. B \textbf{47}, 6835 (1993).
1938: \bibitem{shahbazyan}T. V. Shahbazyan and M. E. Raikh, Phys. Rev. B \textbf{49}, 17123
1939: (1994).
1940: \bibitem{KotthausDD}A. W. Holleitner et al., Science \textbf{297}, 70 (2002).
1941: \bibitem{BoeseDD}D. Boese, W. Hofstetter, and H. Schoeller, Phys. Rev. B \textbf{64},
1942: 125309 (2001).
1943: \bibitem{LossShenjaDD}D. Loss and E. V. Sukhorukov, Phys. Rev. Lett. \textbf{84}, 1035 (2000).
1944: \bibitem{mahnsoo}M.-S. Choi, C. Bruder, and D. Loss, Phys. Rev. B \textbf{62}, 13569
1945: (2000).
1946: \bibitem{BoeseHofstetterSchoeller}D. Boese, W. Hofstetter, and H. Schoeller, Phys. Rev. B 66, 125315
1947: (2002).
1948: \bibitem{HauleBonca}K. Haule and J. Bonca, Phys. Rev. B \textbf{59}, 13087 (1999).
1949: \bibitem{key-2}L. G. Mourokh, N. J. M. Horing, and A. Yu. Smirnov, Phys. Rev. B \textbf{66},
1950: 085332 (2002).
1951: \bibitem{key-1}E. Buks et al., Nature \textbf{391}, 871 (1998).
1952: \bibitem{aleinerwhichpath}I. L. Aleiner, N. S. Wingreen, and Y. Meir, Phys. Rev. Lett. \textbf{79},
1953: 3740 (1997).
1954: \bibitem{ingoldhabil}G.-L. Ingold and Y. V. Nazarov, in: \emph{Single Charge Tunneling},
1955: ed. by H. Grabert and M. Devoret, NATO ASI Series B, vol. 294, Plenum,
1956: New York (1992).
1957: \bibitem{SchoenPE}G. Schön, in: T. Dittrich et al., \emph{Quantum transport and dissipation},
1958: Wiley-VCH, Weinheim (1998).
1959: \bibitem{mahan}G. D. Mahan: \emph{Many-Particle Physics}, Plenum, New York (1981),
1960: pp. 269-310.
1961: \bibitem{devoret}M. H. Devoret et al., Phys. Rev. Lett. \textbf{64}, 1824 (1990).
1962: \bibitem{minnhagen}P. Minnhagen, Phys. Lett. A \textbf{56}, 327 (1976).
1963: \bibitem{gant}V.~F. Gantmakher and Y.~B. Levinson, \emph{Carrier scattering in
1964: metals and semiconductors}, North-Holland, Amsterdam (1987), § 3.5. 
1965: \bibitem{Blum}K. Blum, \emph{Density Matrix Theory and Applications}, Plenum, New
1966: York (1996).
1967: \bibitem{schoellerRTRG}H. Schoeller, Lect. Notes Phys. \textbf{544}, 137 (2000). (cond-mat/9909400)\bibitem{meirwingreen}Y. Meir and N. Wingreen, Phys. Rev. Lett. \textbf{68}, 2512 (1992).
1968: \end{thebibliography}
1969: 
1970: \end{document}
1971: