cond-mat0510289/pml.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %%%% Gilad Rosenberg and Doron Cohen (Sept 2005, Jan2006)
3: %%%% "pml" Quantum stirring of particles in closed devices
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5: 
6: \documentclass{iopart}
7: 
8: \usepackage{fleqn}
9: 
10: % special 
11: \usepackage{ifthen}
12: \usepackage{ifpdf}
13: 
14: % fonts
15: \usepackage{latexsym}
16: %\usepackage{amsmath}
17: %\usepackage{setstack}
18: \usepackage{txfonts}
19: \usepackage{amssymb}
20: \usepackage{bm}
21: 
22: 
23: % figures
24: \ifpdf
25: \usepackage{graphicx}
26: \usepackage{epstopdf}
27: \else
28: \usepackage{graphicx}
29: \usepackage{epsfig}
30: \fi
31: 
32: % math symbols I
33: \newcommand{\sinc}{\mbox{sinc}}
34: \newcommand{\const}{\mbox{const}}
35: \newcommand{\trc}{\mbox{trace}}
36: \newcommand{\intt}{\int\!\!\!\!\int }
37: \newcommand{\ointt}{\int\!\!\!\!\int\!\!\!\!\!\circ\ }
38: \newcommand{\ar}{\mathsf r}
39: \newcommand{\im}{\mbox{Im}}
40: \newcommand{\re}{\mbox{Re}}
41: 
42: 
43: % math symbols II
44: \newcommand{\eexp}{\mbox{e}^}
45: \newcommand{\bra}{\left\langle}
46: \newcommand{\ket}{\right\rangle}
47: 
48: 
49: % more math commands
50: \newcommand{\tbox}[1]{\mbox{\tiny #1}}
51: \newcommand{\bmsf}[1]{\bm{\mathsf{#1}}} 
52: %\newcommand{\amatrix}[1]{\begin{matrix} #1 \end{matrix}} 
53: \newcommand{\amatrix}[1]{\matrix{#1}} 
54: \newcommand{\pd}[2]{\frac{\partial #1}{\partial #2}}
55: 
56: 
57: % equations
58: \newcommand{\be}[1]{\begin{eqnarray}\ifthenelse{#1=-1}{\nonumber}{\ifthenelse{#1=0}{}{\label{e#1}}}}
59: \newcommand{\ee}{\end{eqnarray}} 
60: 
61: 
62: % graphics
63: \newcommand{\drawline}{\begin{picture}(500,1)\line(1,0){500}\end{picture}}
64: \newcommand{\hide}[1]{ }
65: \newcommand{\Cn}[1]{\begin{center} #1 \end{center}}
66: \newcommand{\mpg}[2][\hsize]{\begin{minipage}[b]{#1}{#2}\end{minipage}}
67: \newcommand{\putgraph}[2][\hsize]   {\includegraphics[#1]{#2} }
68: 
69: \begin{document}
70: 
71: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
72: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
73: 
74: \hide{
75: 
76: LIST OF FIGURES
77: 
78: -- Fig1 (models)
79: pmp_scat
80: pmt_fig_d
81: ClosedGeometry2Deltas
82: ClosedGeometry2DeltasWithPerturbation
83: 
84: -- Fig2 (models/schematic)
85: ClosedGeometry
86: OpenGeometry3
87: 
88: -- Fig.3 (cycles)
89: PumpingCycle2
90: pmp_cyc_b3
91: 
92: -- Fig.4 (cycles/B)
93: pmp_cyc_a4
94: pmp_cyc_a3
95: 
96: -- Fig.5  (Q, classical) 
97: ClassicalChargeAllBW
98: ClassicalChargelogX2BW
99: 
100: -- Fig.6 (sigma)
101: DensityOfDegeneracies_normalX2allBW
102: 
103: -- Fig.7 (Q, quantum)
104: OneDegeneracyRoutesDiagram
105: QuantumChargeDiagram
106: NDegeneraciesRoutesDiagram
107: QuantumChargeNDegeneraciesDiagram
108: 
109: -- Fig.8 (degeneracies, 2delta)
110: Diagram7LevelsDegeneraciesFlat
111: Diagram7LevelsDegeneraciesAndAvoidedCrossings2
112: 
113: 
114: -- Fig.9 (degeneracies, 3delta)
115: Diagram7LevelsDegeneraciesPerturbation
116: 
117: -- Fig.10 (Q, numerical)
118: QuantumChargeRoutesBW
119: QuantumChargeBW
120: 
121: 
122: -- Fig.11 (g1, chaotic)
123: OneDeltaOneComplexBarrierGraphBW
124: 
125: -- Fig.12 (UCF)
126: DensityLowTransmissionLogBW
127: 
128: }
129: 
130: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
131: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
132: 
133: \title[Quantum Stirring]
134: {Quantum stirring of particles in closed devices}
135: 
136: \author{Gilad Rosenberg and Doron Cohen}
137: 
138: \address{
139: Department of Physics, Ben-Gurion University, Beer-Sheva 84105, Israel
140: }
141: 
142: 
143: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
144: 
145: \begin{abstract}
146: We study the quantum analog of stirring
147: of water inside a cup using a spoon.
148: This can be regarded as a prototype example
149: for quantum pumping in closed devices.
150: The current in the device is induced by
151: translating a scatterer. Its calculation is done using the Kubo formula
152: approach.
153: The transported charge is expressed as a line integral that encircles chains
154: of Dirac monopoles. For simple systems
155: the results turn out to be counter intuitive:
156: e.g. as we move a small scatterer ``forward"
157: the current is induced ``backwards".
158: One should realize that the route towards
159: quantum-classical correspondence has to
160: do with ``quantum chaos" considerations,
161: and hence assumes greater complexity of the device. We also point out the
162: relation to the
163: familiar $S$ matrix formalism which is used
164: to analyze quantum pumping in open geometries.
165: \end{abstract}
166: 
167: 
168: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
169: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
170: \section{Introduction}
171: 
172: Consider a closed ring that contains particles (Fig.1a). 
173: Assume that one wants to create a current in this ring. 
174: If the particles are charged then one way to do it 
175: is by creating an electro motive force (EMF). 
176: This can be induced by varying an Aharonov-Bohm flux $\Phi$, 
177: such that by Faraday's law $\mbox{EMF}=-\dot{\Phi}$.
178: But there is another way to create a current 
179: that does not involve EMF, and hence does not assume 
180: charged particles. The idea is to change 
181: in time the scalar potential $V(\bm{r}; X_1(t),X_2(t))$.  
182: Here $\bm{r}$ is the coordinate of a representative particle 
183: in the ring, while $X_1$ and $X_2$ are some control 
184: parameters. By making a cycle in the $(X_1,X_2)$ space 
185: we can push non-zero net charge $Q$ through the system. 
186: Thus an ``AC driving" gives rise to a ``DC" component 
187: in the current.  This is known in the literature as ``quantum pumping". 
188: 
189: \vspace{3mm}
190: 
191: %%%%%%%%%%%%%%%%%%%%%%
192: \mpg{
193: \Cn{
194: \putgraph[height=0.2\hsize]{pmp_scat}
195: \\ 
196: \putgraph[height=0.2\hsize]{pmt_fig_d}
197: }
198: 
199: {\footnotesize 
200: {\bf Fig.1.} 
201: %
202: Models for the analysis of quantum stirring. 
203: %
204: (a) Upper panel: A scatterer (big black dot) 
205: is translated inside a Sinai billiard. 
206: A chaotic trajectory of a representative 
207: particle in this billiard is illustrated. 
208: %
209: (b) Lower panels: Network models for 
210: quantum stirring. The scatterer 
211: (big black dot) is translated along 
212: one of the bonds. The vertical 
213: dotted line is the section through 
214: which the current is measured. 
215: From left to right: chaotic network;  
216: double barrier model; triple barrier model. 
217: }
218: }
219: %%%%%%%%%%%%%%%%%%%%%%
220: 
221: \vspace{3mm}
222: 
223: %%%%%%%%%%%%%%%%%%%%%%
224: \mpg{
225: \putgraph[width=0.4\hsize]{ClosedGeometry}
226: \hspace{0.1\hsize}
227: \putgraph[width=0.4\hsize]{OpenGeometry3} 
228: 
229: {\footnotesize 
230: {\bf Fig.2.} 
231: %
232: (a) Left panel: A schematic representation  
233: of the network model. The vertical 
234: dotted line is the section through 
235: which the current is measured.  
236: The moving scatterer is indicated by its transmission $g_0$,   
237: while $X_1$ is its displacement along the bond. 
238: %
239: (b) Right panel: The corresponding 
240: open geometry where the left and right 
241: leads are connected to reservoirs 
242: with the same chemical potentials.
243: }
244: }
245: %%%%%%%%%%%%%%%%%%%%%%
246: 
247: \vspace{3mm}
248: 
249: 
250: In this paper we would like to consider a prototype 
251: pumping problem, which we call ``quantum stirring". 
252: It is the simplest scheme to create a current 
253: with a non-vanishing DC component. Referring to Fig.2   
254: we define $X_1$ as the location of a scatterer,  
255: while $X_2$ is its ``size". By ``size" we mean either 
256: the cross section or the reflection coefficient.
257: One can regard the scatterer as a ``piston" 
258: or as a ``spoon" with which it is possible 
259: to ``push" the particles.  A prototype example 
260: for a pumping cycle is illustrated in Fig.3. 
261: During the main stage of the cycle the 
262: scatterer is translated to the right 
263: a distance $\Delta X_1$. 
264: Consequently a charge $Q$ is transported.    
265: In the second stage the size of the scatterer
266: is ``lowered", and it is displaced back to 
267: its original location, where its original ``size" 
268: is restored. By repeating this cycle many times 
269: we can create a current with a DC component.   
270: 
271: 
272: In the following analysis we assume that the system 
273: consists of non-interacting spinless particles.  
274: All the particles have (formally) charge~$e$, 
275: even if they are not actually charged. 
276: We assume that there is no magnetic field in the system. 
277: Still, for the sake of a later mathematical 
278: formulation, it is convenient to introduce 
279: a third parameter $X_3=\Phi$, 
280: where $\Phi$ is an Aharonv-Bohm flux. 
281: The pumping cycle in the ${(X_1, X_2, X_3)}$ space 
282: is illustrated in Fig.3.    
283: 
284: 
285: \vspace{3mm}
286: 
287: %%%%%%%%%%%%%%%%%%%%%%
288: \mpg{
289: \Cn{
290: \putgraph[width=0.45\hsize]{PumpingCycle2}
291: \putgraph[width=0.4\hsize]{pmp_cyc_b3}
292: }
293: 
294: {\footnotesize 
295: {\bf Fig.3.} 
296: %
297: A prototype example for a pumping cycle. 
298: During the main stage of the cycle the 
299: scatterer is translated to the right 
300: a distance $\Delta X_1$. Consequently 
301: a charge $Q$ is transported.    
302: % 
303: (a) Left panel:  
304: The pumping cycle in the 2-dimensional $(X_1,X_2)$ plane. 
305: %
306: (b) Right panel: The same pumping cycle in 
307: the three dimensional ${(X_1, X_2, X_3)}$ space, 
308: where $X_3=\Phi$ is the Aharonov Bohm flux 
309: via the ring. 
310: }
311: }
312: %%%%%%%%%%%%%%%%%%%%%%
313: 
314: 
315: \vspace{3mm}
316: 
317: %%%%%%%%%%%%%%%%%%%%
318: \mpg{
319: \Cn{
320: \putgraph[width=0.3\hsize]{pmp_cyc_a4}
321: \hspace{0.1\hsize}
322: \putgraph[width=0.3\hsize]{pmp_cyc_a3}
323: }
324: 
325: {\footnotesize 
326: {\bf Fig.4.}
327: %
328: (a) Left panel: The calculation of the charge $Q$ is a line integral 
329: over $G$ that can be regarded as a calculation 
330: of the flux of $\bm{B}$ via a two dimensional curve.  
331: $\vec{ds}$ is a normal vector to the pumping cycle.
332: The black dot in the middle symbolizes the 
333: presence of ``magnetic charge" which is characterized 
334: by a density $\sigma(X_1,X_2)$. In the quantum mechanical 
335: analysis this should be understood as the density 
336: of ``Dirac chains". 
337: %
338: (b) Right panel: In the embedding ${(X_1, X_2, X_3)}$ space 
339: the magnetic charge is organized as vertical charged chains.
340: Each chain consists of ``Dirac monopoles" which are located 
341: at $\vec{X}$ points where an occupied level has a degeneracy 
342: with a nearby level. The ellipse represents a possible 
343: pumping cycle that may encircle either one or many chains. 
344: }
345: }
346: %%%%%%%%%%%%%%%%%%%%
347: 
348: 
349: \vspace{3mm}
350: 
351: 
352: 
353: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
354: \subsection{Linear response theory and the Dirac chains picture} 
355: 
356: We are going to analyze the stirring problem 
357: within the framework of linear response theory. 
358: If we have EMF then we expect to get in the DC limit 
359: Ohm law $\mathcal{I}=-G\dot{\Phi}$, 
360: while if we change slowly either $X_1$ or $X_2$ 
361: we expect to get in the DC limit  $\mathcal{I}=-G^1\dot{X}_1$ 
362: or  $\mathcal{I}=-G^2\dot{X}_2$ respectively. 
363: So in general we can write 
364: %
365: \be{1}
366: Q = \oint_{\tbox{cycle}} Idt 
367: = -\oint (G^1 dX_1 + G^2 dX_2) 
368: =  \oint \bm{B} \cdot \vec{ds} 
369: =  \iint  \sigma(X_1,X_2) dX_1 dX_2 
370: \ee     
371: %
372: In the second expression we define the normal 
373: vector $\vec{ds}=(dX_2,-dX_1)$ and use the 
374: notations ${\bm{B}_1=-G^2}$ and ${\bm{B}_2=G^1}$. 
375: See Fig.4a for an illustration. 
376: The third expression is obtained via the 
377: two dimensional version of the divergence 
378: theorem. If we regard  $\bm{B}$ as a fictitious magnetic 
379: field, then $\sigma$ is the two dimensional density 
380: of magnetic charge.  
381:  
382: 
383: 
384: It turns out that in the strict adiabatic 
385: limit the vector field $\bm{B}$ is 
386: related to the theory of Berry phase \cite{berry,BR}. 
387: The formulation of this relation is as 
388: follows. Assume that the system 
389: is adiabatically cycled in the ${(X_1, X_2, X_3)}$ space. 
390: In such case the Berry phase can be 
391: calculated as a line integral over a 
392: ``vector potential" (also called ``1~form") $\bm{A}$. 
393: This can be converted by the Stokes theorem into 
394: a surface integral over a ``magnetic field" 
395: (also called ``2~form") $\bm{B}$. 
396: The  $\bm{B}$ field is defined as the ``rotor" of $\bm{A}$. 
397: It is a divergence-less field but it can have 
398: singularities  which are known as ``Dirac monopoles".
399: These monopoles are located at $\vec{X}$ points 
400: where an occupied energy level has a degeneracy 
401: with a nearby level.  
402: Because of ${\Phi\mapsto\Phi+(2\pi\hbar/e)}$ gauge invariance    
403: the Dirac monopoles form vertical chains 
404: as illustrated in Fig.4b. Hence we have a distribution  
405: of what we call ``Dirac chains" \cite{pmc,pme}, 
406: which is characterized by a density $\sigma(X_1, X_2)$. 
407: 
408: 
409: \newpage
410: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
411: \subsection{Background and objectives} 
412:   
413: Most of the literature about quantum pumping 
414: deals with the open geometry of Fig.2b. 
415: The most popular approach is the $S$-scattering  
416: formalism which leads 
417: to the B{\"u}ttiker, Pr\^{e}tre and Thomas (BPT) formula \cite{BPT,brouwer}
418: for the generalized conductance $G$. 
419: The BPT formula, is essentially a generalization 
420: of the Landauer formula. In previous publications \cite{pmo,pme}  
421: we have demonstrated that the BPT formula can be 
422: regarded as a  special limit of the Kubo formula. 
423: Our Kubo formula approach to pumping \cite{pmc,pmp} 
424: leads to ``level by level" understanding of the 
425: pumping process, and allows to incorporate 
426: easily non-adiabatic and environmental effects.
427: In the strict adiabatic limit   
428: it reduces in a transparent way 
429: to the theory of adiabatic transport \cite{thouless,avronRev},  
430: also known as ``geometric magnetism \cite{BR}. 
431: On the other hand, in the non-adiabatic(!) 
432: ``DC limit" of an open geometry it reduces 
433: to the $S$ matrix picture, hence resolving 
434: some puzzles that had emerged in older publications.
435:  
436: 
437: The question ``how much charge is pushed 
438: by translating a scatterer" has been addressed 
439: in Ref.\cite{avron} in the case of an open 
440: geometry using the BPT formula.  We have 
441: addressed the corresponding problem 
442: of quantum stirring in closed geometry 
443: in a previous short publication \cite{pmt}, 
444: but the connection with the Dirac chains  
445: picture has not been illuminated. 
446: Furthermore, in \cite{pmt} only the quantum 
447: chaos limit was considered. 
448: 
449: 
450: In the present publication we put an emphasis on clarifying 
451: the route towards quantum-classical correspondence (QCC). 
452: We shall see that quantum mechanical 
453: effects are pronounced in {\em simple} systems. 
454: As the system becomes more chaotic QCC emerges. 
455: The Dirac chains picture leads to new  
456: insights regarding the route towards QCC.
457: These insights are easily missed if we stick 
458: to the formal Green function calculation of 
459: our earlier work \cite{pmt}. 
460: %
461: From the above it should be clear that the main objectives 
462: of the present study are:
463: %
464: \begin{itemize}
465: %
466: \item Derivation of a classical formula for $Q$ (assuming a stochastic picture).
467: %
468: \item Derivation of a quantum result for $Q$ using the Dirac chains picture.
469: %
470: \item Exposing some counter-intuitive results for $Q$ in the case of the simplest models.
471: %
472: \item Illuminating the route towards QCC as we go from ``simple" to ``chaotic" systems.
473: %
474: \end{itemize}
475: %
476: We note that in \cite{pmt} we have presented  
477: the classical formula for $Q$ without the derivation. 
478: 
479: 
480: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
481: \subsection{Physical motivation and experimental feasibility} 
482: 
483: In the previous section we have explained 
484: the theoretical motivations for dealing with 
485: the stirring problem. In the present section 
486: we would like to further discuss the practicality  
487: of this line of study, and the feasibility 
488: of actual experiments. 
489: 
490: It is quite clear that the main focus of today's 
491: experiments is on {\em open} devices (with leads), 
492: whereas our interest is in {\em closed} devices. 
493: Our believe is that {\em ``wireless"} mesoscopic 
494: or molecular size devices are going to be  
495: important building blocks of future ``quantum electronics". 
496: This is of course a vision that people may doubt.    
497: However, on the scientific side our task is  
498: to analyze its feasibility. 
499: 
500: 
501: It is possible to fabricate closed mesoscopic rings, 
502: and to measure the persistent or the induced currents. 
503: Experiments with closed devices have been performed 
504: already~10 years ago. As an example we mention Ref.\cite{orsay} 
505: where a large array of rings has been fabricated. 
506: The current measurement has been achieved by coupling 
507: the rings to a highly sensitive electromagnetic 
508: superconducting micro-resonator. 
509: 
510: 
511: The conceptually simplest way to drive a current 
512: is by inducing an electro motive force (EMF). 
513: In the setup of Ref.\cite{orsay} the EMF 
514: has been induced by a ``wire" that spirals on top 
515: of the array. In our view an attractive alternative option 
516: would be to induce currents by changing gate voltages 
517: so as to induce stirring. The advantage of such 
518: a possibility for the purpose of integrating 
519: wireless devices in future quantum electronics 
520: is quite obvious: It is much easier to control 
521: gate voltages than fluxes of magnetic field.
522: 
523: 
524: As far as {\em electronic devices}  
525: are concerned there is no question about the 
526: feasibility of realizing quantum stirring by 
527: manipulating gate voltages, and measuring 
528: the electrical currents.
529: But we would like to argue that such possibility 
530: is open also in case of {\em neutral atoms}.  
531: It is well known that ``billiards" that confine 
532: cold atoms can be realized and manipulated \cite{raiz,nir}.
533: Furthermore, there is no question regarding the 
534: possibility of creating a ``moving" optical  
535: barrier so as to create a stirring effect. 
536: There are variety of techniques to measure the 
537: induced neutral currents. 
538: For example one can exploit the Doppler effect 
539: at the perpendicular direction, which is known as  
540: the rotational frequency shift \cite{doppler}.
541: 
542:     
543: There is one more issue which might be of relevance 
544: in case of an actual experiment. The Kubo formalism 
545: assumes that the system settles into a steady state, 
546: whereas the preparation in case of an actual experiment 
547: is not very well controlled. We would like to 
548: argue that the results of the linear response analysis  
549: are quite robust. This issue is discussed 
550: in section~4 of Ref.\cite{pms}:  What we get for~$Q$ 
551: in the Kubo analysis is not merely a formal result, 
552: but rather a prediction that has an actual physical significance. 
553: 
554: 
555: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
556: \subsection{Outline} 
557: 
558: 
559: In the first part of this paper  
560: we review the result for $G$ 
561: in the case of an open system using the BPT formula.  
562: %
563: Then we present two equivalent derivations 
564: of the corresponding {\em classical} result 
565: in the case of a closed geometry. 
566: %
567: We use the term ``classical" in the Boltzmann 
568: sense. This means that interference within the 
569: ring is neglected, while the reflection by 
570: the scatterers (``cross section") is calculated  
571: quantum mechanically. 
572: The first derivation is based on a direct solution 
573: of a master equation, while the second 
574: is a straightforward application of the Kubo formula.
575: %
576: The classical calculation implies an expression 
577: for the density $\sigma(X_1,X_2)$ of the monopoles. 
578: The BPT formula implies $\sigma(X_1,X_2)$ that 
579: can be regarded as a special case of this calculation. 
580: 
581: 
582: In the second part of this paper  
583: we turn to the quantum mechanical analysis. 
584: As a preliminary stage we discuss the general 
585: conditions for having a degeneracy point $\vec{X}$ 
586: in the case of a one dimensional ring.   
587: Then we review how the pumped charge $Q$ 
588: can be estimated by calculating a line 
589: integral that encircles ``Dirac chains". 
590: Thus we realize that we have to figure out 
591: what` the distribution $\sigma(X_1,X_2)$ of these 
592: chains looks like. 
593: Specifically, we consider the model systems 
594: that are illustrated in Fig.1, 
595: and schematically in Fig.2.  
596: The simplest is a ring where both $g_1$ 
597: and $g_0$  are modeled as delta barriers. 
598: The result for $Q$ is quite remote from 
599: the classical expectation. 
600: Consequently we try to figure out what happens 
601: to  $\sigma(X_1,X_2)$  as the system becomes 
602: more complex: First we add a second fixed barrier, 
603: and finally we consider what happens 
604: in the case of a ``chaotic" barrier which 
605: is modeled using random matrix theory.
606: We make it clear that the route to 
607: the classical limit is intimately related 
608: to so called ``quantum chaos" considerations.   
609:  
610: 
611: 
612: 
613: 
614: 
615: \newpage
616: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
617: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
618: \section{Pushing particles in an open geometry}
619: 
620: 
621: Let us consider the model of Fig.2b, where 
622: we have a scatterer within a single mode wire 
623: which is connected to two reservoirs with 
624: the same chemical potential.
625: In this section we assume non-interacting 
626: spinless electrons and zero temperature Fermi 
627: occupation.  The scatterer is described by   
628: %
629: \be{0}
630: V(r;X_1,X_2) = X_2\delta(r-X_1)
631: \ee
632: %
633: Hence, for some fixed values of $X_1$ and $X_2$ 
634: its transmission is 
635: %
636: \be{0}
637: g_0(X_2)=
638: \left[ 1 
639: + \left( \frac{\mathsf{m}}{\hbar^2k_{\tbox{F}}}  X_2 \right)^2 
640: \right]^{-1}
641: \ee
642: %
643: where $\mathsf{m}$ is the mass of the particle 
644: and $k_{\tbox{F}}$ is the Fermi momentum. 
645: From now on we work with units such that $\hbar=1$. 
646: The $S$ matrix of the scattering region 
647: can be written in the general form 
648: %
649: \be{0}
650: \mathbf{S} = 
651: \eexp{i\gamma}
652: \left(\matrix{i\sqrt{1-g}\eexp{i\alpha} & \sqrt{g}\eexp{-i\phi} \cr
653: \sqrt{g}\eexp{i\phi} &  i\sqrt{1-g}\eexp{-i\alpha}}\right)
654: \ee
655: %
656: where $\gamma$ is the total phase shift, 
657: $\alpha$ is the reflection phase shift, 
658: and $\phi=e\Phi/\hbar$ represents 
659: the flux which we assume to be zero.   
660: %
661: In the setup of Fig.~2b the length 
662: of the right lead is $L_A-X_1$ 
663: and the length of the left lead 
664: is  $L_B+X_1$. Hence 
665: %
666: \be{0}
667: g &=& g_0
668: \\
669: \gamma &=& k_{\tbox{F}} (L_A+L_B) 
670: -\arctan\left( \frac{\mathsf{m}}{\hbar^2k_{\tbox{F}}}  X_2 \right)  
671: \\
672: \alpha &=& k_{\tbox{F}} (L_A-L_B) 
673: - 2k_{\tbox{F}} X_1 
674: \ee
675: %
676: Now that we know the dependence of the $\mathbf{S}$ matrix 
677: on the parameters $(X_1,X_2)$, the calculation of $G$ 
678: is quite straightforward. We use the BPT formula    
679: %
680: \be{0}
681: G^{j} = \frac{e}{2\pi i}
682: \trc\left(\bm{P}_{\tbox{lead}}\frac{\partial \bm{S}}{\partial X_j}
683: \bm{S}^{\dag}\right)
684: \ee
685: %
686: where $\bm{P}_{\tbox{lead}}$ projects on the channels of 
687: the lead where the current is measured. As indicated in Fig.2b  
688: the current is measured via a section which is located on 
689: the right lead. Using the BPT formula we get
690: %
691: \be{0}
692: G^{1} &=& -(1-g_0) \ \frac{e}{\pi}k_{\tbox{F}}\\
693: G^{2} &=& - g_0  \ \frac{e}{4\pi\hbar v_{\tbox{F}}} 
694: \ee
695: %
696: where $v_{\tbox{F}}$ is the Fermi velocity 
697: corresponding to $k_{\tbox{F}}$. The result 
698: for $G^1$ is our main interest. It has been 
699: discussed in Ref.\cite{avron}, 
700: where the term ``snow plow" has been coined 
701: in order to describe its physical interpretation. 
702: Namely, for zero temperature Fermi occupation  
703: the density of electrons in the wire is ${k_{\tbox{F}}/\pi}$. 
704: Therefore the number of electrons that are pushed 
705: by the scatterer is ${dN=(k_{\tbox{F}}/\pi) \times dX_1}$. 
706: If the transmission of the scatterer is not zero, 
707: some of the electrons pass through it and consequently 
708: we have to multiply $dN$ by the reflection probability ${1-g_0}$. 
709: 
710:  
711: 
712: 
713: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
714: \section{Stirring of particles in a closed geometry}
715: 
716: Let us consider the model of Fig.2a, where the system is closed. 
717: We assume that the transmission of the ring without 
718: the moving scatterer is $g^{cl}_1$, while the transmission 
719: of the scatterer itself is $g_0$. In the following two subsections 
720: we shall present two optional derivations of the ``classical" 
721: result for $G$. We use the term ``classical" in the Boltzmann 
722: sense. Namely, we regard the scattering from 
723: either $g^{cl}_1$ or $g_0$ as a stochastic process. 
724: Thus interference within the arms of the ring is 
725: not taken into account. 
726: For sake of comparison with the BPT-based result 
727: we still assume zero temperature Fermi occupation 
728: (while in later sections we shall allow any arbitrary occupation).   
729: Within this framework we obtain:
730:  %
731: \be{0}
732: G^{1} &=& 
733: - \left[\frac{ (1-g_0) g^{cl}_1 }{ g_0 + g^{cl}_1 - 2g_0 g^{cl}_1 } \right]
734: \ \frac{e}{\pi}k_{\tbox{F}}\\
735: G^{2} &=& 
736: - \left[\frac{ (1-g^{cl}_1) g_0 }{ g_0 + g^{cl}_1 - 2g_0 g^{cl}_1 } \right]
737: \ \frac{e}{4\pi\hbar v_{\tbox{F}}} 
738: \ee
739: % 
740: We note that the amount of charge which is pushed 
741: by translating a scatterer a distance $\Delta X_1$   
742: can also be written as \cite{pmt}
743: %  
744: \be{13}
745: Q = -G^{1} \Delta X_1 = 
746: \left[ \frac{1-g_0}{g_0}\right]
747: \left[ \frac{g_T}{1-g_T}\right]
748: \ \frac{e}{\pi}k_{\tbox{F}} \times \Delta X_1
749: \ee
750: %
751: where $g_T$ is the overall transmission of 
752: the ring (including the moving scatterer) 
753: if it were opened: 
754: %
755: \be{0}
756: \left[ \frac{1-g_T}{g_T}\right] = 
757: \left[ \frac{1-g_0}{g_0}\right] + 
758: \left[ \frac{1-g^{cl}_1}{g^{cl}_1}\right]
759: \ee
760: %
761: As expected the charge $Q$ which is 
762: transported as a result of an $X_1$ displacement 
763: depends in a monotonic way on 
764: the reflection coefficient ${1-g_0}$. 
765: It monotonically increases from zero, 
766: and attains half of its maximal value 
767: for $g_0=g^{cl}_1$.  
768: A plot of $Q$ versus the ``size" of the 
769: scatterer is presented in Fig.~5 for three 
770: representative values of $g^{cl}_1$.   
771: We also plot $Q$ against $X_2$, 
772: assuming that the scatterer is modeled as 
773: a delta function.      
774: 
775: \vspace{10mm}
776: 
777: 
778: %%%%%%%%%%%%%%%%%%%%%%
779: \mpg{
780: \Cn{
781: \putgraph[width=0.4\hsize]{ClassicalChargeAllBW}
782: \putgraph[width=0.4\hsize]{ClassicalChargelogX2BW}
783: }
784: 
785: {\footnotesize 
786: {\bf Fig.5.}
787: Plots of $Q$ as a function of the ``size" 
788: of the scatterer. We use arbitrary units 
789: such that $Q=1$ in the maximum.  
790: %
791: (a) Left panel:  $Q$ is plotted against the reflection 
792: coefficient ${(1-g_0)}$ for ${g^{cl}_1=0.1}$, 
793: for ${g^{cl}_1=0.5}$, and for ${g^{cl}_1=0.9}$. 
794: The dotted lines highlight that $Q$ 
795: for $g_0=g^{cl}_1$ is half its maximum value.
796: Note that the BPT based result corresponds 
797: to ${g^{cl}_1=0.5}$. 
798: %  
799: (b) Right panel: Here $Q$ is plotted against $X_2$ 
800: assuming that the scatterer is a delta function, 
801: and setting~$\mathsf{m} / (\hbar^2 k_{\tbox{F}}) = 1$.   
802: }
803: }
804: %%%%%%%%%%%%%%%%%%%%%%
805: 
806: 
807: \vspace{3mm}
808: 
809: It is important to realize that the result for an {\em open geometry} 
810: is formally a special case corresponding to $g^{cl}_1=1/2$. 
811: This value of $g^{cl}_1$ means that memory is completely 
812: lost once a particle is scattered by the ``surroundings". 
813: Namely, if $g^{cl}_1=1/2$ then after a collision a particle 
814: has equal probability to go in either direction, and any information  
815: about its initial direction is lost. This observation 
816: generalizes our discussion in Ref.\cite{kbf} regarding  
817: the relation between the Kubo and the Landauer conductance.
818: 
819:   
820: The classical expression for $G$ implies  
821: the following result for the density $\sigma(X_1,X_2)$,   
822: which is illustrated in Fig.6. 
823: %
824: \be{15}
825: \sigma(X_1,X_2) 
826: = \frac{d \bm{B}_2 }{dX_2}
827: = -\frac{e\mathsf{m}}{\pi\hbar^2} 
828: \frac
829: { 2 (1-g^{cl}_1) g^{cl}_1}
830: { \left[ 1 
831: + \left( 
832: \left( \frac{\mathsf{m}}{\hbar^2k_{\tbox{F}}} X_2 \right)^2
833: -1 \right) 
834: g^{cl}_1 \right]^2 }
835: \left( \frac{\mathsf{m}}{\hbar^2k_{\tbox{F}}}  X_2 \right)
836: \ee
837: 
838: 
839: 
840: \vspace{3mm}
841: 
842: %%%%%%%%%%%
843: \mpg{
844: \Cn{
845: \putgraph[width=0.5\hsize]{DensityOfDegeneracies_normalX2allBW}
846: }
847: 
848: {\footnotesize 
849: {\bf Fig.6.}
850: The classically deduced density $\sigma$ 
851: as a function of ${X_2}$ 
852: for ${g^{cl}_1=0.1}$, 
853: for ${g^{cl}_1=0.5}$, 
854: and for ${g^{cl}_1=0.9}$. 
855: We use arbitrary units for $\sigma$, 
856: and set~$\mathsf{m} / (\hbar^2 k_{\tbox{F}}) = 1$.
857: The dotted vertical lines correspond  
858: to the median $X_2$ values  
859: which are determined by the 
860: equation ${g_0(X_2) = g^{cl}_1}$. 
861: }
862: }
863: %%%%%%%%%%%
864: 
865: \vspace{3mm}
866: 
867: 
868: In the following sections we give 
869: two optional derivations of the classical result. 
870: The first derivation is based on a physically 
871: appealing master equation approach, 
872: in the spirit of the Boltzmann equation.   
873: The second derivation is a straightforward  
874: application of the Kubo formula. 
875: The calculation is done for $G^1$ and can be 
876: easily modified in order to get $G^2$. 
877: The advantage of the Kubo formula approach 
878: is that it can be generalized 
879: to the quantum mechanical case, 
880: and it allows the incorporation 
881: of non-adiabatic and environmental effects.  
882: 
883: 
884: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
885: \section{Classical derivation using a master equation}
886: 
887: We consider a ring with two scatterers: a moving 
888: scatterer $g_0$ whose velocity is $\dot{X}$,  
889: and a fixed scatterer $g_1$. A collision of a particle 
890: with the moving scatterer implies that its velocity 
891: is changed $v \mapsto v \pm 2\dot{X}$, 
892: where the sign depends on whether the collision 
893: is from the right or from the left. 
894: The associated change in the kinetic energy 
895: is ${E\mapsto E \pm 2\mathsf{m}v\dot{X} + \mathcal{O}(\dot{X}^2) }$ 
896: respectively. There are two regions ($x<0$ and $x>0$) 
897: on the two sides of the $g_0$ scatterer. 
898: Accordingly we have four distribution functions  
899: that satisfy the following balance equations: 
900: %
901: \be{0}
902: \frac{\partial\rho_{+}^{\rightarrow}}{\partial t} 
903: &=& 
904: - \left[\rho_{+}^{\rightarrow}v\right]  
905: + g_0 \left[\rho_{-}^{\rightarrow}v\right]
906: + (1-g_0) \left[\rho_{+}^{\leftarrow}v\right]_{E-2\mathsf{m}v\dot{X}}  
907: \\  
908: \frac{\partial\rho_{+}^{\leftarrow}}{\partial t} 
909: &=& 
910: - \left[\rho_{+}^{\leftarrow}v\right]  
911: + g_1 \left[\rho_{-}^{\leftarrow}v\right]
912: + (1-g_1) \left[\rho_{+}^{\rightarrow}v\right] 
913: \\
914: \frac{\partial\rho_{-}^{\rightarrow}}{\partial t} 
915: &=& 
916: - \left[\rho_{-}^{\rightarrow}v\right]  
917: + g_0 \left[\rho_{+}^{\rightarrow}v\right]
918: + (1-g_0) \left[\rho_{-}^{\leftarrow}v\right] 
919: \\
920: \frac{\partial\rho_{-}^{\leftarrow}}{\partial t} 
921: &=& 
922: - \left[\rho_{-}^{\leftarrow}v\right]  
923: + g_1 \left[\rho_{+}^{\leftarrow}v\right]
924: + (1-g_1) \left[\rho_{-}^{\rightarrow}v\right]_{E+2\mathsf{m}v\dot{X}}  
925: \ee
926: %
927: The zero order solution in $\dot{X}$ is to have all 
928: the four distribution functions equal 
929: to some arbitrary function $f(E)$.  
930: In the presence of driving, assuming that the 
931: system has reached a steady state, 
932: we still have to satisfy the two $\dot{X}$-free 
933: equations, leading to 
934: %
935: \be{0}
936: \rho_{+}^{\leftarrow} &=& 
937: g_1 \rho_{-}^{\leftarrow} + 
938: (1-g_1) \rho_{+}^{\rightarrow}
939: \\
940: \rho_{-}^{\rightarrow} &=& 
941: g_1 \rho_{+}^{\rightarrow} + 
942: (1-g_1) \rho_{-}^{\leftarrow}
943: \ee 
944: %
945: Substitution into the two other equations 
946: leads after linearization to 
947: %
948: \be{0}
949: \rho_{+}^{\rightarrow}(E)-\rho_{-}^{\leftarrow}(E) = 
950: -2\mathsf{m}v\dot{X} \left(\frac{1-g_0}{g_0+g_1-2g_0g_1}\right) 
951: \frac{\partial f(E)}{\partial E}  
952: \ee
953: %
954: and for the current we get
955: %
956: \be{0}
957: I &=& \int_0^{\infty} \frac{dp}{2\pi} 
958: \ (\rho_{\pm}^{\rightarrow}-\rho_{\pm}^{\leftarrow}) ev
959: =   \int_0^{\infty} \frac{dp}{2\pi} 
960: g_1 (\rho_{+}^{\rightarrow}-\rho_{-}^{\leftarrow}) ev
961: \\
962: &=&  -\dot{X} 
963: \int_0^{\infty} 
964: \left[\frac{e}{\pi}\left(\frac{(1-g_0)g_1}{g_0+g_1-2g_0g_1}\right) \mathsf{m}v \right] 
965: \frac{\partial f(E)}{\partial E} dE 
966: \ee 
967: %
968: With the assumption of zero temperature Fermi occupation 
969: this gives the cited result for $G^1$. 
970: 
971: 
972: 
973: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
974: \section{Classical derivation using the Kubo formula}
975: 
976: The generalized fluctuation-dissipation version of 
977: the Kubo formula (see Ref.\cite{pme} and further references therein)  
978: relates the generalized conductance to the the cross   
979: correlation function of the current ${\mathcal{I}}$ and the 
980: generalized force ${\mathcal{F}}=-\partial\mathcal{H}/\partial X$.
981: If $X$ is the displacement $X_1$ of the scatterer then 
982: %
983: \be{25}
984: \mathcal{F} 
985: = -\frac{\partial\mathcal{H}}{\partial X_1}
986: = X_2 \delta'(x-X_1)
987: \ee
988: %    
989: For the sake of comparison with previous results 
990: we assume zero temperature Fermi occupation.
991: Then the Kubo formula takes the form 
992: %
993: \be{26}
994: G = \mathsf{g}(E_{\tbox{F}})  \int_0^{\infty} 
995: \langle {\cal I}(\tau) {\cal F}(0) \rangle  d\tau
996: = \frac{L}{\pi \hbar v_{\tbox{F}}}  \langle {\cal Q} {\cal F} \rangle
997: \ee
998: %
999: where $\mathsf{g}(E)=L/(\pi \hbar v_{\tbox{F}})$ 
1000: is the density of states. This density of states 
1001: is proportional to the total ``volume" of the network 
1002: which is $L$.  In the second expression 
1003: we got rid of the time by introducing the notation 
1004: %
1005: \be{27}
1006: \mathcal{Q} = \int_0^{\infty} {\cal I}(\tau)  d\tau
1007: \ee 
1008: %
1009: It should be clear that both the generalized force $\mathcal{F}$
1010: and the transported charge $\mathcal{Q}$ are functions 
1011: in phase space, and that ${\langle...\rangle}$ 
1012: stands for phase space average over position and velocity.  
1013: For $\mathcal{F}$ we already have an explicit expression Eq.(\ref{e25}).  
1014: Now we have to figure out what is $\mathcal{Q}$. 
1015: 
1016: 
1017: On the ring there are two scatterers, 
1018: and one point $x=x_0$  where the current 
1019: is measured. Hence the ring is divided 
1020: into 3 segments. In addition, there are two possible 
1021: directions of motion (clockwise, anticlockwise). 
1022: Hence the phase space is divided into 6 regions. 
1023: It is obvious that the outcome from Eq.(\ref{e27}) 
1024: depends merely on which region the classical 
1025: trajectory had started its journey in.
1026: In fact we need to consider only the 4 regions 
1027: where the particle starts in the vicinity of the 
1028: moving scatterer, else $\mathcal{F}$ vanishes.
1029: So we have the $``+"$ region between the moving 
1030: scatterer and $x_0$, and the  $``-"$ 
1031: region on the other side between the two scatterers.   
1032: Accordingly the four possible outcomes from Eq.(\ref{e27}) are: 
1033: %
1034: \be{28}
1035: Q_{+}^{\rightarrow} &=& e\left[\frac{1}{2(1-g_T)}\right]
1036: \\ \label{e29}
1037: Q_{+}^{\leftarrow} &=& - e\left[\frac{1}{2(1-g_T)} - 1 \right]
1038: \\ \label{e30}
1039: Q_{-}^{\rightarrow} &=& \left[\frac{g_0}{1-(1-g_1)(1-g_0)} 
1040: -\frac{g_1(1-g_0)}{1-(1-g_1)(1-g_0)}\right] Q_{+}^{\rightarrow} 
1041: =\frac{g_0-g_1+g_0g_1}{g_0+g_1-g_0g_1} Q_{+}^{\rightarrow}
1042: \\ \label{e31}
1043: Q_{-}^{\leftarrow} &=& -\frac{g_1-g_0+g_0g_1}{g_0+g_1-g_0g_1} Q_{+}^{\rightarrow}
1044: \ee
1045: %
1046: The derivation of the above expressions is as follows.
1047: It is simplest if the particle starts in the $``+"$ region, 
1048: because then we can regard the two scatterers as one 
1049: effective scatterer $g_T$.  Assume that at time $t=0$ the 
1050: particle approach $x=x_0$ from the left. 
1051: The charge that goes through the section after a round trip 
1052: is suppressed by a factor $(2g_T-1)$ 
1053: due to the scattering (we sum the clockwise 
1054: and the anticlockwise contributions). 
1055: Thus we find that the total charge that 
1056: goes through the section due to multiple 
1057: reflections is a geometric sum that leads 
1058: to Eq.(\ref{e28}). If we start in the $``+"$ 
1059: region in the opposite direction, then we have 
1060: the same sequence but with the opposite sign 
1061: and without the first term. Hence we get Eq.(\ref{e29}).   
1062: Next assume that at $t=0$ the particle starts in the $``-"$ 
1063: region, and approaches $g_0$ from the left. 
1064: Then we can have at a later time 
1065: a positive pulse of current. The probability 
1066: for that is the geometric summation over 
1067: $g_0 ((1-g_0)(1-g_1))^{\tbox{integer}}$. 
1068: Otherwise, we get a negative pulse of 
1069: current, with a complementary probability 
1070: that can be regarded as a geometric summation over 
1071: $g_1 ((1-g_0)(1-g_1))^{\tbox{integer}} (1-g_0)$. 
1072: Thus the total current through the section,  
1073: taking into account all subsequent multiple reflections 
1074: (rounds) is given by Eq.(\ref{e30}). 
1075: A similar calculation leads to Eq.(\ref{e31}). 
1076: 
1077: 
1078: 
1079: Since there are only four possible values for $\mathcal{Q}$ 
1080: the calculation of the phase space average becomes trivial: 
1081: %
1082: \be{-1}
1083: \langle {\cal Q} {\cal F} \rangle 
1084: = \frac{1}{2L} \left[ \int_{+}\mathcal{F}dr \right] Q_{+}^{\rightarrow}
1085: + \frac{1}{2L} \left[ \int_{+}\mathcal{F}dr \right] Q_{+}^{\leftarrow}
1086: + \frac{1}{2L} \left[ \int_{-}\mathcal{F}dr \right] Q_{-}^{\rightarrow}
1087: + \frac{1}{2L} \left[ \int_{-}\mathcal{F}dr \right] Q_{-}^{\leftarrow} 
1088: \ee
1089: %
1090: The integral over $\mathcal{F}$ is taken either 
1091: within the $``+"$ or within the $``-"$ region. It is 
1092: trivially related to the momentum impact 
1093: and yields the result
1094: %
1095: \be{0}
1096: \int_{\pm}\mathcal{F}dr = \mp \mathsf{m}v_{\tbox{F}}^2
1097: \ee
1098: %   
1099: Putting everything together we get the desired result for $G^1$. 
1100: With some minor modifications we can calculate $G^2$ using 
1101: the same procedure. 
1102: 
1103: 
1104: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1105: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1106: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1107: \section{The quantum mechanical picture}
1108: 
1109: The Kubo formula holds also in the quantum mechanical case. 
1110: But now $\mathcal{I}$ and $\mathcal{F}$ are operators,
1111: so it is more convenient to express the Kubo formula 
1112: using their matrix elements. After some algebra one 
1113: obtains the result: 
1114: %
1115: \be{33}
1116: G = \sum_{m(\ne n)}
1117: \frac{ 2\hbar  \im[\mathcal{I}_{nm}] \mathcal{F}_{mn} }
1118: {(E_m-E_n)^2 + (\Gamma/2)^2}
1119: \ee
1120: %
1121: For more details see Ref.\cite{pme} and further references therein.  
1122: In the above formula it is assumed that only one energy level ($n$) 
1123: is occupied. If we have zero temperature Fermi occupation, 
1124: then we have to sum over all the occupied levels. 
1125: The Kubo formula incorporates a parameter~$\Gamma$ that 
1126: reflects either the non-adiabaticity of the driving, 
1127: or environmentally induced ``memory loss" due to decoherence.   
1128: For a strictly isolated system in the strict adiabatic limit  
1129: we have $\Gamma=0$. Then we identify $G$ as an element 
1130: of Berry's field $\bm{B}$,  
1131: as explained in the introduction. The effect of $\Gamma$ 
1132: on $\bm{B}$ will be discussed below. 
1133: 
1134: 
1135: 
1136: 
1137: We would like to see how the classical result can emerge 
1138: in some limit from the above quantum expression. 
1139: It turns out that this does not require a detailed calculation. 
1140: We can use some topological properties of $\bm{B}$ 
1141: in order to figure out the answer!  The main observations 
1142: that we further explain below are: 
1143: %
1144: \begin{itemize}
1145: \item[\bf{(1)}] The $\bm{B}$ is divergence-less with the exception of Dirac monopoles  
1146: \item[\bf{(2)}] The monopoles are arranged in $\vec{X}$ space as vertical chains  
1147: \item[\bf{(3)}] The far field of $\bm{B}$ is like a two-dimensional electrostatic problem 
1148: \item[\bf{(4)}] Only non-compensated chains give net contribution   
1149: \end{itemize}
1150: %
1151: As long as the occupied level $n$ does not have a degeneracy with 
1152: a nearby level,  $\bm{B}$ is finite and divergence-less. 
1153: Only at degeneracies can it become singular. It can be argued     
1154: that these singularities must have their charge quantized in units 
1155: of $\hbar/2$ else the Berry phase would be ill defined.
1156: We have defined $X_3=\Phi$ as the Aharonov-Bohm flux through the ring. 
1157: This means that if we change $X_3$ by $2\pi\hbar/e$  then by 
1158: gauge invariance we have another degeneracy.  This means 
1159: that the Dirac monopoles are arranged as vertical chains, 
1160: and that the average charge per unit length is $e/(4\pi)$.
1161: Thus the far field of a Dirac chain is as in a two dimensional 
1162: electrostatic problem. If we calculate the line integral  
1163: of Eq.(\ref{e1}) then we get, within the framework of the far field 
1164: approximation, $Q=1$. Thus we conclude that if we have several 
1165: Dirac chains of the same ``sign", then $Q$ simply counts how 
1166: many are encircled. 
1167: 
1168: We have to notice that if we have 
1169: Fermi occupation, then the {\em net} contribution comes only 
1170: from degeneracies of the last occupied level with the first 
1171: unoccupied level. This is what we meant above (item 4) 
1172: by ``non-compensated". In order to avoid 
1173: misunderstanding of the ``compensation" issue 
1174: let us discuss with some more details what happens 
1175: if two neighboring levels $n$ and $m$ are occupied.
1176: With the level $n$ we associate a field $\bm{B}^{(n)}$, 
1177: while with $m$ we associate a field $\bm{B}^{(m)}$. 
1178: In general $\bm{B}^{(m)} \ne -\bm{B}^{(n)}$. If we are near 
1179: a degeneracy than we may say that $\bm{B}^{(n)}$ 
1180: emerges from a Dirac chain which is associated 
1181: with level $n$, while $\bm{B}^{(m)}$ emerges from 
1182: a Dirac chain which is associated with level $m$. 
1183: By inspection of Eq.(\ref{e33}), taking into account 
1184: that ${\im[\mathcal{I}_{nm}]=-\im[\mathcal{I}_{mn}]}$, 
1185: we realize that the two Dirac chains have opposite 
1186: charge. Their corresponding fields do not cancel  
1187: each other, but the total field is no longer singular, 
1188: implying that the {\em net} charge is zero.
1189: 
1190: 
1191: In the quantum stirring problem we shall see that the 
1192: $X_1$ distance between non-compensated chains is simply 
1193: half the De-Broglie wavelength $\lambda_E=2\pi/k_E$. 
1194: From this it follows that the amount of charge which is pushed 
1195: by a very ``large" scatterer is  
1196: %
1197: \be{34}
1198: Q \approx 
1199: e \frac{\Delta X_1}{\lambda_{\tbox{E}}/2}
1200: = e \frac{k_{\tbox{E}}}{\pi} \times \Delta X_1 
1201: \ee
1202: %
1203: What happens if the cycle is not in the ``far field" 
1204: but rather passes through the distribution of the monopoles? 
1205: To be more specific let us consider what happens 
1206: to $Q$ if we displace the scatterer a distance $\Delta X_1$. 
1207: What is the dependence on $X_2$?
1208: Do we get the classical result as in Fig.5?
1209: Obviously, in order to get the classical result 
1210: the distribution $\sigma(X_1,X_2)$ 
1211: should be in accordance with Eq.(\ref{e15}). 
1212: Strictly speaking this is {\em not} the case 
1213: because we have a discrete set of monopoles rather 
1214: than a smooth distribution of ``magnetic charge". 
1215: Still we can hope that $\sigma(X_1,X_2)$
1216: would be classical-like upon course graining. 
1217: We discuss further this issue in the next paragraphs.
1218: 
1219: 
1220: 
1221: If we make a pumping cycle in the vicinity of a monopole 
1222: then it is obvious that the result would be 
1223: very different from the classical prediction. 
1224: What we expect to get in the quantum mechanical case is 
1225: illustrated in the upper panel of Fig.7.
1226: For a cycle that goes very close to a monopole 
1227: the charge can be huge. In reality it is very difficult 
1228: to satisfy the adiabatic condition near a degeneracy, 
1229: or else there are always environmental effects. 
1230: Either way, once we have a finite $\Gamma$, the result 
1231: that we get for $Q$ is smoothed.  
1232:  
1233: 
1234: If the pumping cycle passes through a distribution 
1235: of many monopoles then what we expect to get  
1236: (as we deform or shift the cycle) are huge 
1237: fluctuations as illustrated in the lower panel of Fig.7.   
1238: Again, the effect of either non-adiabaticity or environmental 
1239: effects is to smooth away these fluctuations. 
1240: The interested reader can find some further discussion 
1241: of this point including a numerical example in \cite{pmt}.  
1242: 
1243: 
1244: 
1245: \vspace{3mm}
1246: 
1247: %%%%%%%%%%%
1248: \mpg{
1249: \Cn{
1250: \putgraph[width=0.3\hsize]{OneDegeneracyRoutesDiagram}
1251: \hspace{0.05\hsize}
1252: \putgraph[width=0.4\hsize]{QuantumChargeDiagram} 
1253: \\ \vspace{5mm}
1254: \putgraph[width=0.3\hsize]{NDegeneraciesRoutesDiagram} 
1255: \hspace{0.05\hsize}
1256: \putgraph[width=0.4\hsize]{QuantumChargeNDegeneraciesDiagram} 
1257: }
1258: 
1259: {\footnotesize 
1260: {\bf Fig.7.}
1261: Several pumping cycles are indicated  
1262: in the left panels: It is implicit that 
1263: each segment is closed as in Fig.3. 
1264: The black points represent degeneracies. 
1265: For each pumping cycle one can calculated $Q$. 
1266: The qualitative expectation for the outcome 
1267: is illustrated in the right panels. 
1268: In the upper illustration we assume 
1269: that the pumping cycle encircles only 
1270: one degeneracy, while in the lower 
1271: illustration we assume that it encircles $N$ 
1272: degeneracies. In a later section we 
1273: display numerical results that support 
1274: the illustrated expectations.
1275: }
1276: }
1277: %%%%%%%%%%%
1278: 
1279: \vspace{3mm} 
1280: 
1281: 
1282: 
1283: Coming back to the quantum-classical correspondence (QCC) 
1284: issue, we realize that at best QCC can be satisfied  
1285: in a statistical sense.  So we ask whether the 
1286: coarse grained $\sigma(X_1,X_2)$ agrees 
1287: with the classical expectation Eq.(\ref{e15}).
1288: The answer which we give in the following sections, 
1289: is that QCC is not realized in the case of simple 
1290: non-chaotic models. In the ``simple" cases we 
1291: get a non-classical $\sigma(X_1,X_2)$ 
1292: and hence a different dependence of $Q$ on $X_2$.
1293:  
1294: 
1295: 
1296: 
1297: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1298: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1299: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1300: \section{The degeneracies in $X$ space}
1301: 
1302: We can use the scattering approach in order to find the 
1303: energy levels of a ring. In this approach the ring is opened 
1304: at some arbitrary point and the $S$ matrix of the 
1305: open segment is specified. It is more convenient to  
1306: use the row-swapped matrix, such that the transmission amplitudes 
1307: are along the diagonal:
1308: %
1309: \be{0}
1310: \tilde{\mathbf{S}}(E;X_1,X_2) = 
1311: \eexp{i\gamma}
1312: \left( \amatrix{\sqrt{g}\eexp{i\phi} &  i\sqrt{1-g}\eexp{-i\alpha} \cr
1313: i\sqrt{1-g}\eexp{i\alpha} & \sqrt{g}\eexp{-i\phi}} \right)
1314: \ee
1315: %
1316: The periodic boundary conditions imply the following 
1317: secular equation      
1318: %
1319: \be{0}
1320: \det(\tilde{\mathbf{S}}(E;X_1,X_2) -\bm{1}) = 0
1321: \ee
1322: %
1323: Using
1324: %
1325: \be{0}
1326: && \det(\tilde{\mathbf{S}}-I) = \det(\tilde{\mathbf{S}})-\trc{(\tilde{\mathbf{S}})}+1 \\
1327: && \det(\tilde{\mathbf{S}}) = (\eexp{i\gamma})^2  \\
1328: && \trc{(\tilde{\mathbf{S}})} = 2\sqrt{g}\eexp{i\gamma}\cos{\phi}
1329: \ee
1330: %
1331: we get 
1332: %
1333: \be{0}
1334: \cos(\gamma(E)) = \sqrt{g(E)} \cos(\phi)
1335: \ee
1336: %
1337: In order to find the eigen-energies we plot both sides as 
1338: a function of $E$. The left hand side oscillates between $-1$ 
1339: and $+1$,  while the right hand side may have a smaller 
1340: amplitude. It is not difficult to realize that the only way 
1341: to have two eigen-energies coincide is to get  
1342: %
1343: \be{0}
1344: \begin{array}{cc}
1345: \left\{ \begin{array}{ccc}
1346: \phi &=& 0 \ \mbox{mod}(2\pi)\\
1347: g &=& 1 \\ 
1348: \gamma &=& n_{\tbox{even}}\pi
1349: \end{array}
1350: \right\}
1351: \hspace{0.05\hsize} \mbox{or} \hspace{0.05\hsize}
1352: \left\{ \begin{array}{ccc}
1353: \phi &=& \pi \ \mbox{mod}(2\pi)\\
1354: g &=& 1 \\ 
1355: \gamma &=& n_{\tbox{odd}}\pi
1356: \end{array}
1357: \right\}
1358: \end{array}
1359: \ee
1360: %
1361: where $n$ is either even or odd integer that can 
1362: be exploited (if we keep track over $\gamma$) as a level counter.
1363: 
1364: 
1365: Both $g$ and $\gamma$ depend on $(E;X_1,X_2)$. 
1366: Since we want $g$ to be maximal the condition 
1367: for having a degeneracy involves 4 rather than 3 equations 
1368: as we are going to see below. 
1369: An immediate conclusion is that we have 
1370: two types of Dirac chains: those that have monopoles 
1371: in the plane of the pumping cycle ($X_3=\Phi=0$), 
1372: and the others that have monopoles off 
1373: the plane of the pumping cycle.
1374: 
1375: In our model system we have two scatterers. 
1376: One is the moving scatterer and the other 
1377: is the rest of the network. The two are connected 
1378: by arms of length $L_A-X_1$ and  $L_B+X_1$. 
1379: The constants $L_A$ and $L_B$ can be absorbed into the definition 
1380: of the surrounding network. Each scatterer is fully characterized 
1381: by the set of parameters ${ \{ g_i, \gamma_i, \alpha_i, \phi_i \} }$.
1382: Note that we do not absorb $X_1$ into the definition 
1383: of $\alpha_0$.  After some algebra we find the following expressions 
1384: for the transmission coefficient and for the total phase shift: 
1385: %
1386: \be{0}
1387: g &=& 
1388: \frac{g_0 g_1}
1389: {2 - g_0 - g_1  +  g_0 g_1 + 2\sqrt{(1-g_0)(1-g_1)}
1390: \cos(\gamma_0 + \gamma_1 + \alpha_0 + \alpha_1 - 2k_E X_1)}
1391: \\
1392: \gamma &=& \gamma_0+\gamma_1 
1393: \ee
1394: %
1395: where $k_E$ is the wavenumber that corresponds to the energy $E$.  
1396: Thus the conditions for having a degeneracy take the form   
1397: %
1398: \be{43}
1399: \left\{ \begin{array}{l}
1400: X_3 = \mbox{integer flux} \\
1401: g_0(X_2) = g_1 \\ 
1402: \alpha_0 + \alpha_1 - 2 k_E X_1 = \pi \ \mbox{mod}(2\pi) \\
1403: \gamma_0 + \gamma_1 =  n_{\tbox{even}}\pi
1404: \end{array}
1405: \right.
1406: \left\{ \begin{array}{l}
1407: X_3 = \mbox{half integer flux} \\
1408: g_0(X_2) = g_1 \\ 
1409: \alpha_0 + \alpha_1 - 2 k_E X_1 = 0 \ \mbox{mod}(2\pi) \\
1410: \gamma_0 + \gamma_1 = n_{\tbox{odd}}\pi
1411: \end{array}
1412: \right.
1413: \ee
1414: %
1415: We have highlighted the dependence on the parameters $(X_1,X_2,X_3)$. 
1416: There is of course also an implicit dependence 
1417: of ${ \{ g_i, \gamma_i, \alpha_i \} }$ on the energy $E$. 
1418: The conditions that are listed above are very intuitive: 
1419: The system should have time reversal symmetry; 
1420: The barriers should ``balance"  each other;  
1421: The phases which are associated 
1422: with the reflections should lead to destructive interference;  
1423: And the total phase shift should respect the periodic boundary conditions.   
1424:  
1425: From Eq.(\ref{e43}c) we see that in general 
1426: the $X_1$ distance between degeneracies that belong to the same level  
1427: is roughly half the De-Broglie wavelength as stated previously. 
1428: The question that we would like to address is how these degeneracies are 
1429: distributed with respect to $X_2$. 
1430:  
1431: 
1432: 
1433: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1434: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1435: \section{Quantum stirring in simple rings}
1436: 
1437: We would like to find the distribution 
1438: of degeneracies with respect to $X_2$ 
1439: in the simplest model: a ring with two 
1440: delta scatterers (see Fig.1). The arms 
1441: that connect the two scatterers are 
1442: of length $L_A+X_1$ and $L_B-X_1$. 
1443: For the $S$ matrix that represents 
1444: the fixed scatterer (including the arms) we have 
1445: %
1446: \be{0}
1447: g_1(E) &=& 
1448: \left[ 1 + \left( \frac{\mathsf{m}}{\hbar^2k_{\tbox{F}}} V \right)^2 \right]^{-1}
1449: \\ 
1450: \gamma_1(E) &=& k_E (L_A+L_B)
1451: -\arctan\left( \frac{\mathsf{m}}{\hbar^2k_{\tbox{F}}}  X_2 \right)
1452: \\ 
1453: \alpha_1(E) &=& k_E (L_A-L_B)
1454: \ee
1455: %
1456: Since the dependence of $g_0$ and $g_1$ on the 
1457: barrier ``size" has the same functional form,  
1458: the condition Eq.(\ref{e43}c) implies $X_2=V$ 
1459: irrespective of $E$.
1460: Thus we get that all the degeneracies  
1461: are concentrated at the same $X_2$. 
1462: This is clearly very different from 
1463: the classically expected distribution. 
1464: 
1465: 
1466: In Fig.8 we display an example. The degeneracies 
1467: that are associated with the first 7 levels are 
1468: indicated. Filled circles stand for $\phi=0$ degeneracies, 
1469: while hollow circles stand for $\phi=\pi$ degeneracies. 
1470: Only the last (7th) level contributes non-compensated monopoles. 
1471: The $X_1$ distance between the non-compensated monopoles 
1472: is roughly half De-Broglie wavelength. 
1473: 
1474: 
1475: In Fig.9 we show what happens to the degeneracies 
1476: if we add a second fixed scatterer. We have chosen 
1477: an additional scatterer that can be treated 
1478: as a perturbation. The calculation was done 
1479: using perturbation theory. 
1480: We shall not present the details 
1481: of this lengthy calculation here. 
1482: For larger perturbations (not presented) we had 
1483: to solve the secular equation numerically. 
1484: This was done using an efficient
1485: algorithm \cite{holger}. In any case, the purpose
1486: of Fig.9 is merely to demonstrate that once the
1487: symmetry of the system is broken the degeneracies
1488: spread out in the $X_2$ direction. 
1489: 
1490: 
1491: \vspace{3mm}
1492: 
1493: %%%%%%%%%%%%%%%%%%%%
1494: \mpg{
1495: \Cn{
1496: \putgraph[width=0.49\hsize]{Diagram7LevelsDegeneraciesFlat}
1497: \putgraph[width=0.49\hsize]{Diagram7LevelsDegeneraciesAndAvoidedCrossings2}
1498: }
1499: 
1500: {\footnotesize 
1501: {\bf Fig.8.}
1502: The degeneracies in the double delta model of Fig.1.
1503: We set $L_A=10.23$ and $L_B=0$, so that $X_1$ measures 
1504: the distance from the fixed scatterer. 
1505: The ``size" of the fixed delta scatterer is $V=1274.56$.   
1506: We use units such that $\mathsf{m}=\hbar=1$.
1507: We assume that only the lower~7~levels are occupied.
1508: The filled circles are degeneracies on the flux zero plane 
1509: and the empty circles are degeneracies on the 
1510: flux $\pi$ plane. The left graph shows the actual 
1511: arrangement in the $(X_1,X_2)$ plane. Namely, 
1512: all the degeneracies are on the line ${X_2=V}$. 
1513: In the right graph the degeneracies were displaced 
1514: for the sake of clarity. Only the~7th occupied level 
1515: contributes non-compensated monopoles.  
1516: }
1517: }
1518: %%%%%%%%%%%%%%%%%%%%
1519: 
1520: \vspace{3mm}
1521: 
1522: 
1523: %%%%%%%%%%%%%%%%%%%%
1524: \mpg{
1525: \Cn{
1526: \putgraph[width=0.6\hsize]{Diagram7LevelsDegeneraciesPerturbation}
1527: }
1528: 
1529: {\footnotesize 
1530: {\bf Fig.9.}
1531: The degeneracies in the triple delta model of Fig.1.
1532: Namely, to the model of Fig.8 we have added a delta
1533: barrier of ``size" $V_P=10^{-5}$, located at ${x=7.61}$.
1534: This additional delta barrier can be treated as a small
1535: perturbation. As a result of this perturbation the degeneracies  
1536: shift and spread out in the $X_2$ direction. 
1537: Degeneracies that belong to the same level 
1538: are connected by a line. As in the previous figure 
1539: only the~7th occupied level contributes non-compensated monopoles. 
1540: }
1541: }
1542: %%%%%%%%%%%%%%%%%%%%
1543: 
1544: \vspace{3mm}
1545: 
1546: 
1547: The distribution $\sigma(X_1,X_2)$ in the case of a ring with 
1548: a single fixed scatterer is very different from the 
1549: classical prediction. Consequently also $Q$ comes
1550: out very different from Eq.(\ref{e15}) [and see also Fig.6].
1551: The reader might be curious to know how $Q$ depends 
1552: on the ``size" ($X_2$) of the scatterer in the case of   
1553: Fermi occupation. So we have calculated $G$ 
1554: numerically using Eq.(\ref{e33}), and integrated 
1555: over it to get $Q$. The numerical results are displayed in Fig.10.     
1556: Further analysis of the crossover 
1557: from ``near field" to ``far field" cycles 
1558: will be published in a separate work \cite{pms}.
1559:   
1560: 
1561: \vspace{3mm}
1562: 
1563: %%%%%%%%%%%%%%%%%%%%
1564: \mpg{
1565: \Cn{
1566: \putgraph[width=0.49\hsize]{QuantumChargeRoutesBW}
1567: \putgraph[width=0.49\hsize]{QuantumChargeBW}
1568: }
1569: 
1570: {\footnotesize 
1571: {\bf Fig.10.}
1572: Several pumping routes are displayed 
1573: in the left panel. For each of them $Q$ 
1574: has been calculated numerically.  
1575: The results are displayed in the right panel. 
1576: Note the agreement with the qualitative 
1577: expectation that has been expressed in Fig.7.
1578: The calculation is done for the double 
1579: delta model of Fig.1 with $L_A=1000.23$  
1580: and $L_B=0$. The ``size" of the fixed 
1581: barrier is $V=810.56$. The energy level 
1582: involved are $n=998$ and $m=999$. 
1583: We use units such that $\mathsf{m}=\hbar=1$.
1584: }
1585: }
1586: %%%%%%%%%%%%%%%%%%%%
1587: 
1588: 
1589: 
1590: 
1591: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1592: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1593: \section{Quantum stirring in chaotic rings}
1594: 
1595: We would like to find the distribution 
1596: of degeneracies with respect to $X_2$ 
1597: in case of a chaotic network (see an example in Fig.1).
1598: Let us try to extend the approach that 
1599: has been used in the previous section.
1600: A hypothetical illustration of $g_1(E)$ in the chaotic    
1601: case is displayed in Fig.11. The universal conductance 
1602: fluctuations of $g_1$ are characterized by a one parameter  
1603: probability distribution $P(g_1;\bar{g}_1)$ which 
1604: we discuss below. This probability distribution 
1605: depends on one parameter, which we choose to be 
1606: the average transmission$\bar{g}_1$.
1607: 
1608: \vspace{3mm}
1609: 
1610: %%%%%%%%%%%%%%%%%%%%
1611: \mpg{
1612: \Cn{
1613: \putgraph[width=0.4\hsize]{OneDeltaOneComplexBarrierGraphBW}
1614: }
1615: 
1616: {\footnotesize  
1617: {\bf Fig.11.}
1618: A hypothetical illustration of $g_1(E)$ 
1619: in the case of a complex ``chaotic" barrier.
1620: Such a barrier can be modeled as a network (Fig.1a), 
1621: or it can be characterized using random matrix theory. 
1622: The smooth curves are the transmission $g_0(E;X_2)$ 
1623: of the delta scatterer for 3 different values of $X_2$.     
1624: }
1625: }
1626: %%%%%%%%%%%%%%%%%%%
1627: 
1628: \vspace{3mm}
1629: 
1630: In order to get a degeneracy, a necessary but insufficient 
1631: condition is that the transmission of the two barriers 
1632: is equal (${g_0(E;X_2) = g_1(E)}$). The solution 
1633: of this equation can be determined graphically via Fig.11. 
1634: In fact in most practical applications we can 
1635: assume that our interest is restricted to some 
1636: small energy window such that the smooth $E$ 
1637: dependence of $g_0$ can be neglected. So the 
1638: equation is in fact ${g_0(X_2) = g_1(E)}$. 
1639: For a given $E$ we can find an ${X_2^{(E)}}$ such that this 
1640: equation is satisfied. By playing with $X_1$ we can 
1641: satisfy the $\alpha$ related phase condition 
1642: for having a degeneracy. 
1643: But we still have to satisfy also the $\gamma$ related 
1644: phase condition, which leads to the quantization 
1645: of the energy $E$. Hence the erratic ${X_2^{(E)}}$ 
1646: is sampled. Still it is reasonable to assume that 
1647: the the distribution of the so-obtained $X_2$ values 
1648: is not affected by this random-like sampling. 
1649: We therefore conclude the following relation:   
1650: %
1651: \be{48}
1652: \mbox{Prob}\Big[X_2 < X_2^{(E)} < X_2+dX_2\big] 
1653: \ \ = \  \
1654: \mbox{Prob}\Big[g_0(X_2) < g_1 < g_0(X_2+dX_2)\big] 
1655: \ee
1656: %
1657: This implies a simple relation between $\sigma(X_1, X_2)$ 
1658: and the probability function $P(g_1;\bar{g}_1)$
1659: %
1660: \be{49}
1661: \sigma(X_1,X_2) = \const \times \frac{dg_0(X_2)}{dX_2} P(g_0(X_2)) 
1662: \ee
1663: %
1664: Thus the problem of finding $\sigma(X_1, X_2)$ has reduced 
1665: to the problem of finding $P(g_1;\bar{g}_1)$. 
1666: 
1667: 
1668: We can now proceed in three directions: 
1669: {\bf (A)} To determine $P()$ from simple heuristic 
1670: quantum chaos considerations; 
1671: {\bf (B)} To determine $P()$ from formal random  
1672: matrix theory considerations; 
1673: {\bf (C)} To use reverse engineering in order 
1674: to determine what is $P()$ that would give the 
1675: classical result. 
1676: It should be clear that universality can be 
1677: expected only if $\bar{g}_1 \ll 1$.  
1678: In Fig.12 we make a comparison between 
1679: the outcomes of these three procedures 
1680: for $\bar{g}_1 = 0.001$.  In the following 
1681: paragraph we give the details of the calculation. 
1682: 
1683: \vspace{3mm}
1684: 
1685: %%%%%%%%%%
1686: \mpg{
1687: \Cn{
1688: \putgraph[width=0.49\hsize]{DensityLowTransmissionLogBW}
1689: }
1690: 
1691: {\footnotesize 
1692: {Fig.12.}
1693: A plot of the distribution $P(g_1;\bar{g}_1)$ according to several different 
1694: expressions. In this calculation we assume that the average  
1695: transmission is ${\bar{g}_1=0.001}$, which is represented in the 
1696: figure by a vertical dashed line. The ``heuristic" result is based on 
1697: sampling of the random variable $g_1=\bar{g}_1\eta_1\eta_2$ where $\eta$ 
1698: is Porter-Thomas distributed. The ``RMT" result is based on Eq.(\ref{e51}).
1699: The ``classical" result is based on Eq.(\ref{e52}). 
1700: }
1701: }
1702: %%%%%%%%%%
1703: 
1704: \vspace{3mm}
1705: 
1706:  
1707: The heuristic approach is based on the idea that the transmission 
1708: via a chaotic network depends on the amplitudes of the wavefunctions 
1709: at the entrance and exit points. One might expect $g_1=\bar{g}_1\eta_1\eta_2$, 
1710: where $\eta$ has the Porter-Thomas distribution \cite{haake}
1711: ${P_{\tbox{GOE}}(\eta) = ({1}/{\sqrt{2\pi\eta}})\eexp{-{\eta}/{2}}}$. 
1712: This leads to the ``heuristic" result in Fig.12. 
1713: In fact this result should not be taken too 
1714: seriously. The formal RMT calculation \cite{BB} 
1715: of the probability distribution $P(g_1;\bar{g}_1)$
1716: leads to the following expressions:
1717: %
1718: \be{51}
1719: P_{\tbox{RMT}}(g_1;\bar{g}_1) = 
1720: \left\{ 
1721: \amatrix{
1722: ({2}/{\pi^2 \bar{g}_1}) \ g_1^{-1/2} & \ \ \mbox{for} \ g_1 \ll (\bar{g}_1)^2 \ll 1 
1723: \cr
1724: ({4\bar{g}_1}/{\pi^2}) \ g_1^{-3/2} & \ \ \mbox{for} \ (\bar{g}_1)^2 \ll g_1 \ll 1
1725: }\right.
1726: \ee
1727: %
1728: The small $g_1$ approximation is universal:   
1729: it merely assumes that the system has time reversal 
1730: symmetry. It has been confirmed \cite{KS} 
1731: that this universal behavior holds also for network 
1732: systems. But for larger values of $g_1$ there are 
1733: deviations that has to do with semiclassical considerations. 
1734: It is therefore in the latter region where one might 
1735: expect quantum-classical correspondence.    
1736: 
1737: 
1738: The probability distribution $P(g_1;\bar{g}_1)$
1739: that would reproduce the classical 
1740: result Eq.(\ref{e15}) via Eq.(\ref{e49}) is:
1741: %
1742: \be{52}
1743: P_{\tbox{CL}}(g_1;\bar{g}_1) =
1744: \frac{(1-g^{cl}_1)g^{cl}_1}{(g_1+g^{cl}_1-2g_1g^{cl}_1)^2}
1745: \ee
1746: %
1747: with ${g_1^{cl} \approx 0.12\bar{g_1}}$. 
1748: In order to compare with the RMT result we note that 
1749: %
1750: \be{53}
1751: P_{\tbox{CL}}(g_1;\bar{g}_1) \approx 
1752: \left\{ 
1753: \amatrix{
1754: (1/g^{cl}_1) \ (1-2g_1/g_1^{cl}) & \ \ \ \mbox{for} \ g_1 \ll g^{cl}_1 \ll 1 \cr
1755: g^{cl}_1 \ g_1^{-2} & \ \ \ \mbox{for} \ g^{cl}_1 \ll g_1 \ll 1 
1756: }\right.
1757: \ee
1758: %
1759: We see that in the large $g_1$ region, where one  
1760: might expect quantum-classical correspondence,  
1761: there is no agreement between  $P_{\tbox{CL}}()$ 
1762: and $P_{\tbox{RMT}}()$. 
1763: We suspect that $P_{\tbox{RMT}}()$ cannot 
1764: be trusted there. Otherwise we have to conclude  
1765: that Eq.(\ref{e49}) fails to take into account  
1766: strong correlations in the arrangement of Dirac monopoles.
1767: Either way it seems that RMT alone 
1768: is not enough in order to reproduce the classical result.   
1769:      
1770: 
1771: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1772: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1773: \section{The emergence of the classical limit}
1774: 
1775: With simple minded RMT reasoning we have failed to 
1776: get a quantitative correspondence with the classical result.   
1777: We therefore look for a different way to get 
1778: an estimate for either $\bm{B}_2$ or $\sigma(X_1,X_2)$
1779: in the case of a chaotic network. One obvious 
1780: way is to use the result of Ref.\cite{diabolic} 
1781: regarding the distribution of degeneracies (diabolic points).
1782: The perturbation term which is associated with $X_2$ is 
1783: %
1784: \be{530}
1785: \mathcal{W} 
1786: = \frac{\partial\mathcal{H}}{\partial X_2}
1787: = \delta(x-X_1)
1788: \ee
1789: %   
1790: and the density of the degeneracies 
1791: should be \cite{diabolic}  
1792: %
1793: \be{0} 
1794: \sigma(X_1,X_2) 
1795: \ \ = \ \ 
1796: \frac{\pi}{3} \mathsf{g}(E)^2 
1797: \ \mbox{RMS}[\mathcal{F}_{nm}] 
1798: \ \mbox{RMS}[\mathcal{W}_{nm}] 
1799: \ \ \propto \ \  \mbox{RMS}[\mathcal{W}_{nn}]
1800: \ee
1801: %
1802: where $\mathsf{g}(E)$ is the density of states.
1803: In the first equality it is implicit that 
1804: the root mean square (RMS) of {\em near~diagonal} 
1805: matrix elements should be estimated. 
1806: In fact only  $\mbox{RMS}[\mathcal{W}_{nm}]$ 
1807: is required in order to find the $X_2$ dependence.
1808: For a quantum chaos system with time reversal symmetry 
1809: the variance of the near diagonal elements 
1810: equals half the variance of the diagonal 
1811: elements \cite{agam}, leading to the second expression.  
1812: 
1813: 
1814: There is a well known semiclassical recipe \cite{mario1,mario2} 
1815: for calculating the variance of the near diagonal 
1816: matrix elements $\mathcal{W}_{nm}$. 
1817: One should find the classical correlation function 
1818: ${C(\tau)=\langle \mathcal{W}(t) \mathcal{W}(0) \rangle - \langle \mathcal{W} \rangle^2 }$,  
1819: and then integrate over $\tau$. 
1820: If $\mathcal{W}$ were the current operator 
1821: then $\langle \mathcal{W} \rangle$ would be equal to zero, 
1822: and we could proceed as in section~5. 
1823: But in case of Eq.(\ref{e530}) there is a problem: 
1824: The sign of $\mathcal{W}(t)$ does not fluctuate, 
1825: and it is essential to take into account 
1826: the distribution of the delay times 
1827: inside the network. Therefore there is no obvious 
1828: relation to the transmissions $g_0$ and $g_1$. 
1829: 
1830: 
1831: An optional possibility is to try to 
1832: evaluate  $\mbox{RMS}[\mathcal{W}_{nn}]$, 
1833: where $\mathcal{W}_{nn}=|\psi_{\tbox{barrier}}|^2$ 
1834: is the ``intensity" of the wavefunction 
1835: at the location of the scatterer. Obviously 
1836: the result depends on both $g_0$ and $g_1$, 
1837: and requires considerations which are at least 
1838: as difficult as estimating universal conductance 
1839: fluctuations. So it seems that we would run 
1840: into the same problems as in the previous section.
1841: 
1842:   
1843: Still there is the option to calculate $G^1=\bm{B}_2$ 
1844: from the Green function of the system. 
1845: This has been done in \cite{pmt}:  
1846: Writing the Green function as a sum over trajectories, 
1847: we have expressed $G^1$ as a double sum over paths.
1848: If this double sum is averaged over the energy 
1849: one obtains the diagonal approximation, 
1850: leading to the classical result. At first glance 
1851: the energy averaging is not quite legitimate, 
1852: because the energy is quantized. But one can justify 
1853: this procedure in the case of a ``quantum chaos system".  
1854: We have further supported this claim by the numerical 
1855: analysis of the chaotic network of Fig.1 \cite{pmt}.   
1856: We therefore conclude that for a chaotic network 
1857: the distribution of degeneracies should be in accordance 
1858: with Eq.(\ref{e15}). 
1859:      
1860: 
1861: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1862: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1863: \section{Conclusions}
1864: 
1865: As we translate a scatterer of ``size" $X_2$ 
1866: a distance $\Delta X_1$ along a single mode wire, 
1867: the amount of charge which is pushed is    
1868: %  
1869: \be{0}
1870: Q \ \ = \ \ r(X_2) \times \frac{e}{\pi}k_{\tbox{F}} \times \Delta X_1
1871: \ee
1872: %
1873: where $k_{\tbox{F}}$ is the Fermi momentum. If the 
1874: scatterer is very ``large" (${X_2\rightarrow\infty}$)
1875: then we expect to have $r(X_2)=1$. 
1876: This expectation is based on the ``snow plow" picture 
1877: that has been explained in the conclusion of section~2.  
1878: This result is also confirmed by the formal BPT based 
1879: calculation in the case of an {\em open} geometry. 
1880: It also can be formally derived for a {\em closed}  
1881: geometry using the ``Dirac chains picture". 
1882: In the latter case the key observation is 
1883: that the $X_1$ distance between contributing 
1884: degeneracies is roughly half the De-Broglie wavelength.
1885: See Eq.(\ref{e34}).
1886: 
1887: 
1888: Next we ask what happens to  $r(X_2)$ as $X_2$ becomes smaller. 
1889: In the case of an {\em open} geometry the intuitive naive 
1890: guess, which is based on the ``snow plow" picture,  
1891: turns out to be correct. 
1892: Namely, ${r(X_2)=1-g_0 }$ is simply the reflection 
1893: coefficient: Some of particles are not ``pushed" 
1894: by the scatterer because of its partial transparency.   
1895: In the case of a {\em closed} geometry we have 
1896: shown that the {\em classical} result for $r(X_2)$ is modified: 
1897: now it depends also on the overall transmission of 
1898: the device. See Eq.(\ref{e13}).
1899: 
1900: 
1901: It is important to realize that the {\em classical} result 
1902: for $r(X_2)$ is in complete agreement with the common  
1903: sense expectation. Namely, we have ${0 < r(X_2) < 1 }$, 
1904: and the dependence on the ``size" of the 
1905: scatterer is monotonic. 
1906: But once we go to the quantum mechanical analysis 
1907: we have a surprise. The results that we get  
1908: are counter-intuitive. They are most puzzling (Fig.10) 
1909: in the case of the simplest model, in which the ring contains 
1910: only one fixed delta barrier ($V$). 
1911: As we decrease $X_2$ the transported charge $Q$ 
1912: becomes larger(!). Moreover, once $X_2$ becomes 
1913: smaller than $V$, the coefficient $r(X_2)$ 
1914: changes sign. This means that as we push the particles 
1915: ``forward" the current is induced ``backwards".
1916: 
1917: 
1918: The reason for the failure of our intuition is 
1919: our tendency to regard ``adiabatic transport"  
1920: as a zero order adiabatic approximation, 
1921: while in fact it is based on a first order analysis 
1922: (for a detailed discussion see section~4 of \cite{pms}). 
1923: As a parameter in the system  is changed, 
1924: the induced current can be in either direction.    
1925: 
1926: 
1927: In order to understand the route towards quantum-classical 
1928: correspondence it is essential to figure out how 
1929: the degeneracies spread out in $\vec{X}$ space. 
1930: As the system becomes more complex, we get for $r(X_2)$   
1931: a result that resembles the classically implied one.  
1932: The resemblance is at best only on a coarse grained scale: 
1933: the quantum result has strong fluctuations.   
1934: These are related to universal conductance fluctuations. 
1935:    
1936: We have made an attempt to deduce from RMT considerations 
1937: the ``chaotic" distribution of the degeneracies, 
1938: and hence the dependence of $r(X_2)$ on $X_2$. 
1939: The quantitative results do not agree. We therefore 
1940: suspect that RMT considerations alone are not enough 
1941: in order to establish quantum-classical correspondence.
1942: Rather we had used \cite{pmt} semiclassical tools 
1943: in order to establish this correspondence. 
1944:  
1945: 
1946: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1947: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1948: 
1949: \vspace{3mm} 
1950: 
1951: \ack
1952: 
1953: We thank Nir Davidson for illuminating us regarding 
1954: the feasibility of measuring neutral currents in 
1955: cold atom systems. We also had the pleasure to have 
1956: very helpful discussions with Itamar Sela, 
1957: Tsampikos Kottos, Holger Schanz and Michael Wilkinson.  
1958: This research was supported by 
1959: the Israel Science Foundation (grant No.11/02),
1960: and by a grant from the GIF, the German-Israeli Foundation 
1961: for Scientific Research and Development.
1962: 
1963: 
1964: 
1965: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1966: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1967: 
1968: %\Bibliography{99}
1969: \section*{References}
1970: \begin{thebibliography}{99}
1971: 
1972: \bibitem{berry} 
1973: M.V. Berry, Proc. R. Soc. Lond. A {\bf 392} (1984) 45.
1974: 
1975: \bibitem{BR} 
1976: M.V. Berry and J.M. Robbins, 
1977: Proc. R. Soc. Lond. A {\bf 442} (1993) 659. 
1978: 
1979: \bibitem{pmc} 
1980: D. Cohen, Phys. Rev. B {\bf 68} (2003) 155303.
1981: 
1982: \bibitem{pme} 
1983: For a mini-review and further references see:   
1984: D. Cohen, Physica E {\bf 29}, 308 (2005).
1985: 
1986: \bibitem{BPT} 
1987: M.~B{\"u}ttiker et al, Z.~Phys.~B-Condens.~Mat., {\bf 94} (1994) 133.
1988: 
1989: \bibitem{brouwer} 
1990: P. W. Brouwer, Phys. Rev. B {\bf 58} (1998) R10135. 
1991: 
1992: \bibitem{pmo}
1993: D. Cohen, Phys. 
1994: Rev. B {\bf 68} (2003) 201303(R).
1995: 
1996: \bibitem{pmp} 
1997: D. Cohen, Solid State Communications {\bf 133}, 583 (2005).
1998: 
1999: \bibitem{thouless} 
2000: D. J. Thouless, Phys. Rev. {\bf B27} (1983) 6083.
2001: 
2002: \bibitem{avronRev}
2003: J.E. Avron et al, 
2004: Rev. Mod. Phys. {\bf 60} (1988) 873.
2005: 
2006: \bibitem{avron}
2007: J. E. Avron et al, 
2008: Phys. Rev. B {\bf 62}, R10 (2000) 618.
2009: 
2010: \bibitem{pmt}
2011: D. Cohen, T. Kottos and H. Schanz, 
2012: Phys. Rev. E {\bf 71}, 035202(R) (2005).
2013: 
2014: 
2015: \bibitem{orsay} 
2016: B. Reulet M. Ramin, H. Bouchiat and D. Mailly, 
2017: Phys. Rev. Lett. {\bf 75}, 124 (1995). 
2018: 
2019: \bibitem{raiz}
2020: C.F. Bharucha, J.C. Robinson, F.L. Moore, 
2021: K.W. Madison, S.R. Wilkinson, Bala Sundaram, and M.G. Raizen, 
2022: Atomic Physics {\bf 15}, 62 (World Scientific, Singapore, 1997). 
2023: \ V. Milner, J.L. Hanssen, W.C. Campbell, and M.G. Raizen,  
2024: Phys. Rev. Lett. {\bf 86}, 1514 (2001).
2025: 
2026: \bibitem{nir}
2027: N. Friedman, A. Kaplan, D. Carasso, and N. Davidson, 
2028: Phys. Rev. Lett. {\bf 86},  1518 (2001).
2029: \ M. Andersen, and N. Davidson, 
2030: Phys. Rev. Lett. {\bf 87}, 274101 (2001).
2031: 
2032: 
2033: \bibitem{doppler}
2034: I. Bialynicki-Birula and Z. Bialynicka-Birula, 
2035: Phys. Rev. Lett. {\bf 78}, 2539 (1997).
2036: 
2037: 
2038: \bibitem{pms} 
2039: I. Sela and D. Cohen, cond-mat/0512500.
2040: 
2041: 
2042: \bibitem{kbf}
2043: D. Cohen and Y. Etzioni,
2044: J. Phys. A {\bf 38}, 9699 (2005).
2045: 
2046: 
2047: \bibitem{holger} 
2048: We thank H. Schanz for suggesting the numerical algorithm.
2049: 
2050: 
2051: \bibitem{haake} 
2052: F. Haake, Quantum Signatures of Chaos (Springer 2001).
2053: 
2054: \bibitem{BB} 
2055: P.W. Brouwer and C.W.J. Beenakker, 
2056: Phys. Rev. B. {\bf 50}, 11263 (1994). 
2057: 
2058: \bibitem{KS} 
2059: T. Kottos and U. Smilansky, 
2060: J. Phys. A {\bf 36}, 3501-3524 (2003).
2061: 
2062: \bibitem{diabolic} 
2063: M. Wilkinson and E. J. Austin,  
2064: Phs. Rev. A {\bf 47}, 4 (1993).
2065: 
2066: \bibitem{agam}
2067: B. Eckhardt, S. Fishman, J. Keating, O. Agam, J. Main and K. Muller, 
2068: Phys. Rev. E {\bf 52}, 5893 (1995); and further references therein.
2069: 
2070: \bibitem{mario1}
2071: M. Feingold and A. Peres, Phys. Rev. A {\bf 34} 591, (1986).
2072: 
2073: \bibitem{mario2}
2074: M. Feingold, D. Leitner, M. Wilkinson, Phys. Rev. Lett. {\bf 66}, 986 (1991). 
2075:  
2076: 
2077: \end{thebibliography}
2078: 
2079: 
2080: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2081: \end{document}
2082: