cond-mat0610572/text.tex
1: % Use only LaTeX2e, calling the article.cls class and 12-point type.
2: 
3: %\documentclass[12pt]{article}
4: \documentclass[twocolumn,english,aps,prl,superscriptaddress]{revtex4}
5: % Users of the {thebibliography} environment or BibTeX should use the
6: % scicite.sty package, downloadable from *Science* at
7: % www.sciencmag.org/misc/con-info.shtml .  This package should properly
8: % format in-text reference calls and reference-list numbers.
9: 
10: %\usepackage{scicite}
11: 
12: % Use times if you have the font installed; otherwise, comment out the
13: % following line.
14: 
15: \usepackage{times}
16: 
17: \usepackage{graphicx}
18: \usepackage{babel}
19: \usepackage{amsmath}
20: \usepackage{amsfonts}
21: \usepackage{amssymb}
22: \usepackage{epsfig}
23: 
24: 
25:  \oddsidemargin 0.2cm \textwidth 16cm \textheight
26: 25cm \footskip 1.0cm
27: 
28: \begin{document}
29: \normalsize{\textbf{Science, Vol.313, p.499, 2006}}\\\\
30: \title{Violation of Kirchhoff's
31: Laws for a Coherent RC Circuit}
32: 
33: \author{J. Gabelli}\affiliation{Laboratoire Pierre Aigrain, D{\'e}partement
34: de Physique de l'Ecole Normale Sup\'erieure, 24 rue Lhomond, 75231
35: Paris Cedex 05, France} \author{G. F\`{e}ve}\affiliation{Laboratoire
36: Pierre Aigrain, D{\'e}partement de Physique de l'Ecole Normale
37: Sup\'erieure, 24 rue Lhomond, 75231 Paris Cedex 05,
38: France}\author{J.-M. Berroir}\affiliation{Laboratoire Pierre
39: Aigrain, D{\'e}partement de Physique de l'Ecole Normale
40: Sup\'erieure, 24 rue Lhomond, 75231 Paris Cedex 05,
41: France}\author{B. Pla\c{c}ais}\affiliation{Laboratoire Pierre Aigrain,
42: D{\'e}partement de Physique de l'Ecole Normale Sup\'erieure, 24 rue
43: Lhomond, 75231 Paris Cedex 05, France}\author{A.
44: Cavanna}\affiliation{Laboratoire de Photonique et Nanostructures,
45: UPR20 CNRS, Route de Nozay, 91460 Marcoussis Cedex,
46: France}\author{B. Etienne}\affiliation{Laboratoire de Photonique et
47: Nanostructures, UPR20 CNRS, Route de Nozay, 91460 Marcoussis Cedex,
48: France}\author{Y.Jin}\affiliation{Laboratoire de Photonique et
49: Nanostructures, UPR20 CNRS, Route de Nozay, 91460 Marcoussis Cedex,
50: France}\author{D.C. Glattli} \email{glattli@lpa.ens.fr}
51: \affiliation{Laboratoire Pierre Aigrain, D{\'e}partement de Physique
52: de l'Ecole Normale Sup\'erieure, 24 rue Lhomond, 75231 Paris Cedex
53: 05, France} \affiliation{Service de Physique de l'Etat Condens{\'e},
54: CEA Saclay, F-91191 Gif-sur-Yvette, France}
55: 
56: 
57: 
58: 
59: \begin{abstract}
60: What is the complex impedance of a fully coherent quantum
61: resistance-capacitance (RC)circuit at GHz frequencies in which a
62: resistor and a capacitor are connected in series? While Kirchhoff's
63: laws predict addition of capacitor and resistor impedances, we
64: report on observation of a different behavior. The resistance,here
65: associated with charge relaxation, differs from the usual transport
66: resistance given by the Landauer formula. In particular, for a
67: single mode conductor, the charge relaxation resistance is half the
68: resistance quantum, regardless of the transmission of the mode. The
69: new mesoscopic effect reported here is relevant for the dynamical
70: regime of all quantum devices.
71: \end{abstract}
72: 
73: \maketitle
74: For a classical circuit, Kirchhoff's laws prescribe the
75: addition
76:  of resistances in series. Its failure has been a central issue in developing our understanding
77:  of electronic transport in mesoscopic conductors. Indeed, coherent multiple electronic
78: reflections between scatterers in the conductor were
79:  found to make the conductance non-local \cite{Landauer57,Landauer70}. A new composition law of
80:  individual scatterer contribution to resistance was found that led to
81:  the solution of the problem of electron localization \cite{Anderson81} and, later,
82:  to formulation of the electronic conduction in terms of
83:  scattering of electronic waves \cite{ButtImLandauer83}. Nonadditivity of series
84:  resistances, or of parallel conductances, nonlocal effects and
85:  negative four-point resistances \cite{GaoBO} have been
86:  observed in a series of transport experiments at low temperature, where phase
87:  coherence extends over the mesoscopic scale \cite{Review1,Review2}.
88:  It is generally accepted that the conductance of a phase-coherent quantum
89:  conductor is given by the Landauer formula and its generalization to multi-lead conductors \cite{ButtLand},
90: which relate the conductance to the transmission of electronic waves
91: by the conductance quantum $e^2/h$. But, how far is this description
92: robust at finite frequency where conductance combines with
93: nondissipative circuit elements such as capacitors or inductors ?
94: Are there significant departures from the dc result ? The question
95: is important, as recent advances in quantum information highlight
96: the need for fast manipulation of quantum systems, in particular
97: quantum conductors. High frequency quantum transport has been
98: theoretically addressed, showing that a quantum RC circuit displays
99: discrepancies with its classical counterpart
100: \cite{Buttiker93,Pretre96}.
101: \begin{figure}
102: \centerline{\includegraphics[width=6 cm,
103: keepaspectratio]{fig1.eps}}\caption{{\small The quantum capacitor
104: realized using a 2DEG (A) and its equivalent circuit (B). The
105: capacitor consists of a metallic electrode (in gold) on top of a
106: submicrometer 2DEG quantum dot (in blue) defining the second
107: electrode. The resistor is a QPC linking the dot to a wide 2DEG
108: reservoir (in blue), itself connected to a metallic contact (dark
109: gold). The QPC voltage $V_G$ controls the number of electronic modes
110: and their transmission. The radio frequency voltage $V_{ac}$, and
111: eventually a dc voltage $V_{dc}$, are applied to the
112: counter-electrode whereas the ac current, from which the complex
113: conductance is deduced, is collected at the ohmic contact. As
114: predicted by theory, the relaxation resistance $R_q$, which enters
115: the equivalent circuit for the coherent conductance, is
116: transmission-independent and equal to half the resistance quantum.
117: The capacitance is the serial combination $C_\mu$ of the quantum and
118: the geometrical capacitances ($C_q$ and $C$ respectively). $C_q$ is
119: transmission-dependent and strongly modulated by $V_{dc}$ and/or
120: $V_G$. The combination of $R_q$ and $C_q$ forms the impedance
121: $1/g_q$ of the coherent quantum conductor.}}\label{HBT1}\end{figure}
122: It was shown that a counter-intuitive modification of the series
123: resistance lead to the situation in which the resistance is no
124: longer described by the Landauer formula and does not depend on
125: transmission in a direct way \cite{Buttiker93,Pretre96}. Instead it
126: is directly related to the dwell time of electrons in the capacitor.
127: Moreover, when the resistor transmits in a single electronic mode, a
128: constant resistance was found, equal to the half-resistance quantum
129: $h/2e^2$, i.e., it was not transmission-dependent. This resistance,
130: modified by the presence of the coherent capacitor, was termed a
131: "charge-relaxation resistance" to distinguish it from the usual dc
132: resistance, which is sandwiched between macroscopic reservoirs and
133: described by the Landauer formula. The quantum charge-relaxation
134: resistance, as well as its generalization in nonequilibrium systems,
135: is an important concept that can be applied to quantum information.
136: For example, it enters into the problem of quantum-limited detection
137: of charge qubits \cite{Pilgram} \cite{Clerk}, in the study of
138: high-frequency-charge quantum noise
139: \cite{ChargeNoise1,ChargeNoise2,ChargeNoise3}, or in the study of
140: dephasing of an electronic quantum interferometer \cite{Seelig}. In
141: molecular electronics, the charge relaxation resistance is also
142: relevant to the THz frequency response of systems such as carbon
143: nanotubes\cite{Burke}.
144: 
145: We report on the observation and quantitative measurement of the
146: quantum charge-relaxation resistance in a coherent RC circuit
147: realized in a two-dimensional electron gas (2DEG) (see Fig.1A). The
148: capacitor is made of a macroscopic metallic electrode on top of a
149: 2DEG submicrometer dot defining the second electrode. The resistor
150: is a quantum point contact (QPC) connecting the dot to a wide 2DEG
151: macroscopic reservoir.  We address the coherent regime in which
152: electrons emitted from the reservoir to the dot are backscattered
153: without loss of coherence. In this regime, we have checked the
154: prediction made in refs.\cite{Buttiker93,Pretre96} that the
155: charge-relaxation resistance is not given by the Landauer formula
156: resistance but instead is constant and equals $h/2e^2$, as the QPC
157: transmission is varied. Note that we consider here a spin-polarized
158: regime and that the factor $1/2$ is not the effect of spin, but a
159: hallmark of a charge-relaxation resistance. When coherence is washed
160: out by thermal broadening, the more conventional regime pertaining
161: to dc transport is recovered. The present work differs from previous
162: capacitance measurements where, for spectroscopic purpose, the dot
163: reservoir coupling was weak and the ac transport regime was
164: incoherent \cite{Ashoori92a,Ashoori92b}. As a consequence, although
165: quantum effects in the capacitance were observable, the quantum
166: charge-relaxation resistance was not accessible in these earlier
167: experiments.
168: 
169: At zero temperature in the coherent regime and when a single mode is
170: transmitted by the QPC, the mesoscopic RC circuit is represented by
171: the equivalent circuit of Fig.1B \cite{Buttiker93,Pretre96}. The
172: geometrical capacitance $C$ is in series with the quantum admittance
173: $g_{q}(\omega)$ connecting the ac current flowing in the QPC to the
174: ac internal potential of the dot:
175: \begin{equation}\label{cond0}
176: g_{q}(\omega)=\frac{1}{\frac{h}{2e^{2}}+\frac{1}{-i\omega C_{q}}}
177: \qquad (T = 0)
178: \end{equation}
179: The nonlocal quantum impedance behaves as if it were the series
180: addition of a quantum capacitance $C_{q}$ with a constant contact
181: resistance $h/2e^2$. $C_{q}=e^{2}\frac{d\mathcal{N}}{d\varepsilon}$
182: is associated with the local density of state
183: $\frac{d\mathcal{N}}{d\varepsilon}$ of the mode propagating in the
184: dot, taken at the Fermi energy. The striking effect of phase
185: coherence is that the QPC transmission probability $D$ affects the
186: quantum capacitance (see Eq.\ref{Cq}) but not the resistance. The
187: total circuit admittance $G$ is simply :
188: \begin{equation}\label{grandcon}
189:  G=\frac{-i\omega C g_{q}(\omega)}{-i\omega C+g_{q}(\omega)}= \frac{-i\omega C_{\mu}\frac{2e^{2}}{h}}{-i\omega
190:  C_{\mu}+\frac{2e^{2}}{h}}\quad , \qquad (T = 0)
191: \end{equation}
192: where $C_{\mu}=\frac{CC_{q}}{C+C_{q}}$ is the electrochemical
193: capacitance. In the incoherent regime, both resistance and quantum
194: capacitance vary with transmission. The dot forms a second reservoir
195: and the electrochemical capacitance $C_{\mu}$ is in series with the
196: QPC resistance $R$. In particular, when the temperature is high
197: enough to smooth the capacitor density of states, the Landauer
198: formula $R=\frac{h}{e^{2}}\times\frac{1}{D}$ is recovered.
199: 
200: Several samples have been measured at low temperatures, down to 30
201: mK, which show analogous features. We present results on two samples
202: made with 2DEG defined in the same high-mobility GaAsAl/GaAs
203: heterojunction, with nominal density $n_{s}=1.7 \times 10^{15}$
204: m$^{-2}$ and mobility $\mu =260$ V$^{-1}$m$^{2}$s$^{-1}$. A finite
205: magnetic field ($B=1.3$ T) is applied, so as to work in the
206: ballistic quantum Hall regime with no spin degeneracy \cite{SOM}.
207: 
208: \begin{figure}[hhh]
209: \centerline{\includegraphics[width=9 cm,
210: keepaspectratio]{fig2.eps}}\caption{{\small Complex conductance of
211: sample E3 as function of the gate voltage $V_G$ for $T=100$ mK and
212: $\omega /2\pi=1.2$ GHz, at the opening of the first conduction
213: channel (C) and its Nyquist representation in (D). The theoretical
214: circle characteristic of the coherent regime is shown as a solid
215: line. (A and B) show the corresponding curves for the simulation of
216: sample E3 using the 1D model with $C=4$ $ fF$, $C_\mu=1$ $
217: fF$.}}\label{HBT2}\end{figure}
218: 
219: The real and imaginary parts of the admittance $Im(G)$ and $Re(G)$
220: as a function of QPC gate voltage $V_{G}$ at the opening of the
221: first conduction channel are shown in Fig.2C. On increasing $V_{G}$,
222: we can distinguish three regimes. At very negative $V_{G}\leq -0.86$
223: V, the admittance is zero. Starting from this pinched state, peaks
224: are observed in both $Im(G)$ and $Re(G)$. Following a maximum in the
225: oscillations, a third regime occurs where $Im(G)$ oscillates nearly
226: symmetrically about a plateau while the oscillation amplitude
227: decreases smoothly. Simultaneously, peaks in $Re(G)$ quickly
228: disappear to vanish in the noise.
229: 
230: Comparing these observations with the results of
231: refs.\cite{Buttiker93,Pretre96},using a simplified one-dimensional
232: (1D) model for $C_q$ with one conduction mode and a constant energy
233: level spacing in the dot $\Delta$ \cite{Gabelli06}, the simulation
234: (Fig.2A) shows a striking similarity to the experimental conductance
235: traces in Fig.2C. In this simulation, $V_{G}$ determines the
236: transmission $D$ but also controls linearly the 1D dot potential.
237: The transmission is chosen to vary with $V_G$ according to a
238: Fermi-Dirac-like dependence appropriate to describe QPC transmission
239: \cite{Butt90}. This model can be used to get a better understanding
240: of the different conductance regimes. Denoting $r$ and $t$ the
241: amplitude reflection and transmission coefficients of the QPC
242: ($r^{2}=1-D$, $t=\sqrt{D}$), we first calculate the scattering
243: amplitude of the RC circuit:
244: \begin{equation}\label{S}
245: s(\varepsilon)=r-t^{2}e^{i\varphi}\sum_{n=0}^{\infty}(re^{i\varphi})^{n}=\frac{r-e^{i\varphi}}{1-re^{i\varphi}}
246: \end{equation}
247: where $\varepsilon$ is the Fermi energy relative to the dot
248: potential and $\varphi=2\pi\varepsilon/\Delta$ is the phase
249: accumulated for a single turn in the quantum dot. The
250: zero-temperature quantum capacitance is then given by:
251: \begin{equation}\label{Cq}
252: C_q=e^{2}\frac{d\mathcal{N}}{d\varepsilon}=\frac{1}{2i\pi}s^{+}\frac{\partial
253: s}{\partial
254: \varepsilon}=\frac{e^{2}}{\Delta}\frac{1-r^{2}}{1-2r\cos2\pi\frac{\varepsilon}{\Delta}+r^{2}}
255: \end{equation}
256: Therefore, $C_q$ exhibits oscillations when the dot potential is
257: varied. When $r \rightarrow 0$,  these oscillations vanish and
258: $C_q\rightarrow e^{2}/\Delta$. As reflection increases, oscillations
259: are growing with maxima $\frac{e^{2}}{\Delta} \frac{1+r}{1-r}$ and
260: minima $\frac{e^{2}}{\Delta} \frac{1-r}{1+r}$. For strong
261: reflection, Eq \ref{Cq} gives resonant Lorentzian peaks with an
262: energy width $D\Delta/2$ given by the escape rate of the dot.
263: However, at finite temperature, the conductance in equation
264: (\ref{cond0}) has to be thermally averaged to take into account the
265: finite energy width of the electron source so that :
266: \begin{eqnarray}\label{condT}
267:      g_{q}(\omega) = \int d\varepsilon \left ( - \frac{\partial f}{\partial \varepsilon}\right ) \frac{1}{h/2e^2+1/(-i\omega
268:      C_{q})}\quad \\
269:      (T\neq 0)  \nonumber
270: \end{eqnarray}
271: where $f$ is the Fermi-Dirac distribution. Again, the nonlocal
272: admittance behaves as if it were the serial association of a
273: charge-relaxation resistance $R_q$ and a capacitance that we still
274: denote $C_q$. In the weak transmission regime ($D\rightarrow 0$),
275: when $D\Delta \ll k_{B}T $, equation (\ref{condT}) yields thermally
276: broadened capacitance peaks with
277: \begin{equation}\label{fanC}
278: C_{q} \simeq
279: \frac{e^{2}}{4k_{B}T\cosh^{2}(\delta\varepsilon/2k_{B}T)}\quad,\qquad
280: (D<<1)
281: \end{equation}
282: where $\delta \varepsilon$ denotes the energy distance to a resonant
283: dot level. Note that these capacitance peaks do not depend on the
284: dot parameters and can be used as a primary thermometer. Similar but
285: transmission-dependent peaks are predicted in the inverse resistance
286: \begin{equation}\label{fanR}
287: 1/R_{q} \simeq
288: D\frac{e^{2}}{h}\frac{\Delta}{4k_{B}T\cosh^{2}(\delta\varepsilon/2k_{B}T)},\qquad
289: (D<<1)
290: \end{equation}
291: This result is reminiscent of the thermally broadened resonant
292: tunneling conductance for transport through a quantum dot. A
293: consequence of the finite temperature is the fact that the
294: resistance is no longer constant. This thermally-induced divergence
295: of $R_q$ at low transmission restores a frequency-dependent
296: pinch-off for $R_q\gg 1/C_q\omega$, as can be seen in both model and
297: experiment in Figs.2A.C. As mentioned above, for $k_{B}T \gg
298: D\Delta$, the quantum dot looks like a reservoir and the Landauer
299: formula is recovered.
300: 
301: The coherent and the thermally broadened regimes are best
302: demonstrated in the Nyquist representation $Im(G)$ versus $Re(G)$ of
303: the experimental data in Fig.2D. This representation allows to
304: easily distinguish constant resistance from constant capacitance
305: regimes, as they correspond to circles respectively centered on the
306: real and imaginary axis. Whereas, for low transmission, the Nyquist
307: diagram strongly depends on transmission, the conductance
308: oscillations observed in Fig.2C collapse on a single curve in the
309: coherent regime. Moreover this curve is the constant $R_q=h/2e^2$
310: circle. By contrast, admittance peaks at low transmission correspond
311: to a series of lobes in the Nyquist diagram, with slopes increasing
312: with transmission in qualitative agreement with Eqs.\ref{fanC} and
313: \ref{fanR}. These lobes and the constant $R_q$ regime are well
314: reproduced by the simulations in Fig.2B. Here, the value of $C_\mu$
315: and the electronic temperature are deduced from measurement. In our
316: experimental conditions, the simulated traces are virtually free of
317: adjustable parameters as $C\geq 4C_\mu \gg C_{q}$.
318: 
319: 
320: It is important to note that in a real system, the weak transmission
321: regime is accompanied by Coulomb blockade effects that are not taken
322: into account in the above model. In the weak transmission regime and
323: $T=0$, using an elastic co-tunneling approach
324: \cite{Averin92,Glattli93}, we have checked that there is no
325: qualitative change except for the energy scale that now includes the
326: charging energy so that $\Delta$ is replaced by
327: $\Delta$+$e^2/C$=$e^2/C_{\mu}$. At large transmission, the problem
328: is nonperturbative in tunnel coupling and highly nontrivial.
329: Calculations of the thermodynamic capacitance exist
330: \cite{Matveev,Glazman98,Brouwer05}, but at present, no comprehensive
331: model is available that would include both charge-relaxation
332: resistance and quantum capacitance for finite temperature and/or
333: large transmission.
334: \begin{figure}[hhh]
335: \centerline{\includegraphics[width=3.5 in,
336: keepaspectratio]{fig3.eps}}\caption{{\small Coulomb blockade
337: oscillations in the real part of the ac conductance in the
338: low-transmission regime. The control voltage is applied to the
339: counter-electrode for Sample E3 (A) and to the QPC gate for Sample
340: E1 (B). The temperature dependence is used for absolute calibration
341: of our setup, as described in the text : the peak width, shown in (C
342: and D) as a function of temperature, is deduced from theoretical
343: fits (solid lines) using Eq.(\ref{fanR}) and taking a linear
344: dependence of energy with the control voltage. Lines in (C) and (D)
345: are fits of the experimental results using a $\sqrt{T^2+T_0^2}$ law
346: to take into account a finite residual electronic temperature
347: $T_{0}$.}}\label{HBT3}\end{figure} Calibration of our admittance
348: measurements is a crucial step toward extracting the absolute value
349: of the constant charge-relaxation resistance. As at GHz frequencies,
350: direct calibration of the whole detection chain is hardly better
351: than 3dB, we shall use here an indirect, but absolute, method, often
352: used in Coulomb blockade spectroscopy, that relies on the comparison
353: between the gate voltage width of a thermally broadened Coulomb peak
354: ($\propto k_{B}T$) and the Coulomb peak spacing ($\propto
355: e^2/C_\mu$). From this, an absolute value of $C_{\mu}$ can be
356: obtained. The real part of the admittance of Sample E3 is shown as a
357: function of the dc voltage $V_{dc}$ at the counter-electrode, for a
358: given low transmission (Fig.3A). A series of peaks with periodicity
359: $\Delta V_{dc}=370\, \mu$V are observed, with the peaks accurately
360: fitted by Eq.\ref{fanR}. Their width, proportional to the electronic
361: temperature $T_{el}$, is plotted versus the refrigerator temperature
362: $T$ (see Fig.3C). When corrected for apparent electron heating
363: arising from gaussian environmental charge noise, and if we assume
364: $T_{el}=\sqrt{T^{2}+T_{0}^{2}}$, the energy calibration of the gate
365: voltage yields $C_\mu$ and the amplitude $1/C_{\mu}\omega$ of the
366: conductance plateau in Fig.2. A similar analysis is done in Fig.3, B
367: and D, for sample E1 using $V_G$ to control the dot potential. Here
368: peaks are distorted because of a transmission-dependent background
369: and show a larger periodicity $\Delta V_{G}=2$ mV, which reflects
370: the weaker electrostatic coupling to the 2DEG.
371: 
372: \begin{figure}[hhh]
373: \centerline{\includegraphics[width=9 cm,
374: keepaspectratio]{fig4.eps}}\caption{{\small Complex impedance of
375: Sample E3 (A and B) and Sample E1 (C and D) as a function of QPC
376: voltage for $T= 30$ mK and $B=1.3$ T. The dashed lines in (B and D)
377: correspond to the values of $C_{\mu}$ deduced from calibration. The
378: horizontal solid lines in (A and C) indicate the half-quantum of
379: resistance expected for the coherent regime. Uncertainties on
380: $R_{q}$ are displayed as hatched areas.}}\label{HBT4}\end{figure}
381: 
382: 
383: Finally, after numerical inversion of the conductance data, we can
384: separate the complex impedance into the contributions of the
385: capacitance, $1/C_\mu\omega$, and the relaxation resistance $R_q$.
386: The results in Fig.4 demonstrate deviations from standard
387: Kirchhoff's laws : the charge-relaxation resistance $R_q$ remains
388: constant in the regime where the quantum capacitance exhibits strong
389: transmission-dependent oscillations; this constant value equals,
390: within experimental uncertainty, half the resistance quantum as
391: prescribed by theory \cite{Buttiker93,Pretre96}.  In the weak
392: transmission regime, the Landauer formula is recovered because of
393: thermal broadening, and $R_q$ diverges as it does in the dc regime.
394: Furthermore, additional measurements at $4K$ prove that the
395: classical behavior is indeed recovered in the whole transmission
396: range whenever $k_BT\gg e^2/C_\mu$.
397: 
398: In conclusion, we have experimentally shown that the series
399: association of a quantum capacitor and a model quantum resistor
400: leads to a violation of the dynamical Kirchhoff's law of impedance
401: addition. In the fully coherent regime, the quantum resistor is no
402: longer given by the Landauer formula but by the half-quantized
403: charge-relaxation resistance predicted in
404: Ref.\cite{Buttiker93,Pretre96}.\\
405: The LPA is the CNRS-ENS UMR8551 associated with universities Paris 6
406: and Paris 7. The research has been supported by AC-Nanoscience,
407: SESAME grants and ANR-05-NANO-028.
408: 
409: 
410: \begin{thebibliography}{9}
411: 
412: \bibitem{Landauer57} R. Landauer, \emph{IBM J. Res. Dev.} \textbf{1}, 233
413: (1957).
414: 
415: \bibitem{Landauer70} R. Landauer, \emph{Phil. Mag.} \textbf{21}, 863 (1970).
416: %Electrical resistance of disordered one-dimensional lattice%  .
417: 
418: \bibitem{Anderson81} P.W. Anderson, \emph{Phys. Rev. B} \textbf{23}, 4828 (1981).
419: %New method for scaling theory of localization. II. Multichannel theory of a "wire" and possible extension to higher dimensionality %.
420: 
421: \bibitem{ButtImLandauer83} M. B\"{u}ttiker, Y. Imry, R. Landauer, S. Pinhas, \emph{Phys. Rev. B.} \textbf{31}, 6207 (1985).
422: %Generalized many Channel Conductance Formula with Application to small Rings %
423: 
424: \bibitem{GaoBO} B. Gao, A. Komnik, R. Egger, D.C. Glattli, A. Bachtold,
425: \emph{Phys. Rev. Lett.} \textbf{92}, 216804 (2004).
426: 
427: \bibitem{Review1} see for a review, S. Datta, \emph{Electronic Transport in Mesoscopic
428: Systems}, Cambridge University Press, Cambridge, (1997).
429: 
430: \bibitem{Review2} see for a review, Y Imry, \emph{Introduction
431: to Mesoscopic Physics}, Oxford University Press, Oxford (1997).
432: 
433: \bibitem{ButtLand} M. B\"uttiker, \emph{Phys. Rev. Lett.} \textbf{57}, 17611764 (1986)
434: 
435: \bibitem{Buttiker93} M. B\"{u}ttiker, A. Pr\^{e}tre, H. Thomas, \emph{Phys. Rev. Lett.} \textbf{70}, 4114
436: (1993).
437: 
438: \bibitem{Pretre96} A. Pr\^{e}tre, H. Thomas, M. B\"{u}ttiker, \emph{Phys. Rev. B.}
439: \textbf{54}, 8130 (1996).
440: 
441: \bibitem{Pilgram} S. Pilgram, M. B\"uttiker, \emph{Phys. Rev. Lett.} \textbf{89}, 200401
442: (2002).
443: 
444: \bibitem{Clerk} A.A. Clerk, S.M. Girvin, A.D. Stone, \emph{Phys. Rev. B.}
445: \textbf{67}, 165324  (2003).
446: %Quantum-limited measurement and information in mesoscopic detectors%
447: 
448: \bibitem{ChargeNoise1} M. B\"uttiker, H. Thomas, A. Pr\^{e}tre, \emph{Phys. Lett. A} \textbf{180},
449: 364 (1993).
450: %mesoscopic capacitors%
451: 
452: \bibitem{ChargeNoise2} Ya. M. Blanter, M. B\"uttiker, \emph{Phys. Rep.} \textbf{336}, 1
453: (2000).
454: %Shot Noise in Mesoscopic Conductors%
455: 
456: \bibitem{ChargeNoise3} F.W.J. Hekking, J.P. Pekola, \emph{Phys. Rev. Lett.} \textbf{96}, 056603
457: (2006).
458: 
459: \bibitem{Seelig} G. Seelig, S. Pilgram, A. N. Jordan, M. B\"uttiker, \emph{Phys. Rev. B.} \textbf{68}, 161310
460: (2003).
461: 
462: \bibitem{Burke} P. J. Burke, \emph{IEEE T Nanotechnol} \textbf{2} (1),
463: 55 (2003).
464: %"An RF Circuit Model for Carbon Nanotubes,"%
465: 
466: \bibitem{Ashoori92a} R. C. Ashoori \emph{et al.},  \emph{Phys. Rev. Lett.} \textbf{68},
467: 3088 (1992).
468: %single electron capacitance spectroscopy of discrete quantum levels%
469: 
470: \bibitem{Ashoori92b} R. C. Ashoori \emph{et al.},
471: \emph{Phys. Rev. Lett.} \textbf{71}, 613 (1992).
472: %N-electron ground state energies of a quantum dot in magnetic field%
473: 
474: \bibitem{SOM} Materials and methods are available as supporting material on
475: Science Online.
476: 
477: \bibitem{Gabelli06} J. Gabelli, thesis,
478: Universit\'e Pierre et Marie Curie, Paris, 2006;
479: %{\emph Evidence for quantum coherence in dynamical electronic transport}%.
480: On line access at (http://tel.ccsd.cnrs.fr/tel-00011619).
481: 
482: \bibitem{Butt90} M. B\"{u}ttiker,
483: \emph{Phys. Rev. B} \textbf{41}, 7906 (1990).
484: % Quantized transmission of a saddle-point constriction %.
485: 
486: \bibitem{Averin92} D.V. Averin, Y. Nazarov, \emph{Single charge tunneling}, chap.6, NATO-ASI series B, Vol.\textbf{294}, Plenum, New York, 1992.
487: 
488: \bibitem{Glattli93} D.C. Glattli,
489: \emph{Physica B} \textbf{189}, 88 (1993).
490: % Coulomb blockade and off-resonance tunneling in small electronic systems %.
491: 
492: \bibitem{Matveev} K.A. Matveev, \emph{Phys. Rev. B.}\textbf{51} , 1743
493: (1995).
494: 
495: \bibitem{Glazman98} L.I. Glazman, I.L. Aleiner, \emph{Phys. Rev. B.} \textbf{57},
496: 9608 (1998).
497: 
498: \bibitem{Brouwer05} P.W. Brouwer, A. Lamacraft, K. Flensberg
499: \emph{Phys. Rev. B.} \textbf{72}, 075316 (2005), and references
500: therein.
501: %Nonequilibrium theory of Coulomb blockade in open quantum dots%
502: 
503: %\bibitem{Chaotic}  P. W. Brouwer and M. B\(uttiker, Europhys. Lett., 37 (7), pp. 441-446 (1997)
504: 
505: \end{thebibliography}
506: 
507: \end{document}
508: