cond-mat0512500/pms.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %%%  May 2005
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: 
5: \documentclass{iopart}
6: 
7: \usepackage{fleqn}
8: 
9: % special 
10: \usepackage{ifthen}
11: \usepackage{ifpdf}
12: 
13: % fonts
14: \usepackage{latexsym}
15: %\usepackage{amsmath}
16: \usepackage{amssymb}
17: \usepackage{bm}
18: 
19: 
20: % figures
21: \ifpdf
22: \usepackage{graphicx}
23: \usepackage{epstopdf}
24: \else
25: \usepackage{graphicx}
26: \usepackage{epsfig}
27: \fi
28: 
29: % math symbols I
30: \newcommand{\sinc}{\mbox{sinc}}
31: \newcommand{\const}{\mbox{const}}
32: \newcommand{\trc}{\mbox{trace}}
33: \newcommand{\intt}{\int\!\!\!\!\int }
34: \newcommand{\ointt}{\int\!\!\!\!\int\!\!\!\!\!\circ\ }
35: \newcommand{\ar}{\mathsf r}
36: \newcommand{\im}{\mbox{Im}}
37: \newcommand{\re}{\mbox{Re}}
38: 
39: 
40: % math symbols II
41: \newcommand{\eexp}{\mbox{e}^}
42: \newcommand{\bra}{\left\langle}
43: \newcommand{\ket}{\right\rangle}
44: \newcommand{\mass}{\mathsf{m}}
45: 
46: 
47: % more math commands
48: \newcommand{\tbox}[1]{\mbox{\tiny #1}}
49: \newcommand{\bmsf}[1]{\bm{\mathsf{#1}}} 
50: \newcommand{\amatrix}[1]{\begin{matrix} #1 \end{matrix}} 
51: \newcommand{\pd}[2]{\frac{\partial #1}{\partial #2}}
52: 
53: 
54: % equations
55: \newcommand{\be}[1]{\begin{eqnarray}\ifthenelse{#1=-1}{\nonumber}{\ifthenelse{#1=0}{}{\label{e#1}}}}
56: \newcommand{\ee}{\end{eqnarray}} 
57: 
58: 
59: % graphics
60: \newcommand{\drawline}{\begin{picture}(500,1)\line(1,0){500}\end{picture}}
61: \newcommand{\hide}[1]{}
62: \newcommand{\Cn}[1]{\begin{center} #1 \end{center}}
63: \newcommand{\mpg}[2][\hsize]{\begin{minipage}[b]{#1}{#2}\end{minipage}}
64: \newcommand{\putgraph}[2][width=\hsize]{ \includegraphics[#1]{#2} }
65: 
66: 
67: \begin{document}
68: 
69: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
70: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
71: 
72: \title[Quantum pump in a closed circuit]
73: {Operating a quantum pump in a closed circuit}
74: 
75: \author{Itamar Sela and Doron Cohen}
76: 
77: \address{
78: Department of Physics, Ben-Gurion University, Beer-Sheva 84005, Israel
79: }
80: 
81: 
82: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
83: 
84: \begin{abstract}
85: During an adiabatic pumping cycle a conventional 
86: two barrier quantum device takes an electron 
87: from the left lead and ejects it to the right lead. 
88: Hence the pumped charge per cycle is naively 
89: expected to be $Q \le e$. This zero order adiabatic 
90: point of view is in fact misleading. For a closed device 
91: we can get ${Q > e}$ and even ${Q \gg e}$. 
92: In this paper a detailed analysis of the 
93: quantum pump operation is presented. 
94: Using the Kubo formula for the geometric 
95: conductance, and applying the Dirac chains picture,  
96: we derive practical estimates for~$Q$. 
97: \end{abstract}
98: 
99: 
100: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
101: \section{Introduction}
102: 
103: 
104: Understanding of charge transport in mesoscopic 
105: and molecular size devices is essential 
106: for the future realization of `quantum circuits' \cite{rev}.   
107: Of particular interest are adiabatic processes 
108: that take electrons and move them one by one 
109: via a device. The simple minded peristaltic point 
110: of view of such process is misleading:    
111: such picture looks valid in the case of an open circuit, 
112: but breaks down once the pump is integrated into 
113: a closed circuit \cite{pmp,pmc,pme}.
114: %
115: It is the purpose of this paper to further 
116: elaborate on the physics of quantum pumping 
117: in {\em closed circuits}, and to provide a detailed 
118: analysis of a prototype pump. 
119: %
120: The interest and the feasibility of realizing 
121: experiments with closed circuits is discussed 
122: in Section~1.3 of \cite{pml}.
123: 
124: 
125: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
126: \subsection{The prototype pumping device}
127: 
128: The prototype example for a quantum pumping 
129: device is the two barrier model (Fig.1). 
130: The one particle Hamiltonian is   
131: %
132: \be{1}
133: \mathcal{H}(X_1(t),X_2(t)) = \frac{1}{2\mass}\hat{p}^2 +
134: X_1(t)\delta(\hat{x}-x_1)+X_2(t)\delta(\hat{x}-x_2)
135: \ee
136: %
137: where $\mass$ is the mass of the particle 
138: and $(\hat{x},\hat{p})$ are the position 
139: and the momentum operators. 
140: The region $x<x_1$ is the left lead, 
141: and the region $x>x_2$ is the right lead.
142: We refer to the segment $x_1<x<x_2$ 
143: as the ``dot region". 
144: The driving is performed by 
145: changing $X_1$ and $X_2$ in time. 
146: In an actual experiment the control 
147: parameters $X_1$ and $X_2$ represent gate voltages.
148: %
149: %
150: In order to talk about charge transfer we 
151: have to define also a current operator. In what 
152: follows we use the conventional definition 
153: %
154: \be{2}
155: \mathcal{I}=\frac{e}{2\mass}
156: \left(\hat{p}\delta(\hat{x}-x_0) + \delta(\hat{x}-x_0)\hat{p}\right) 
157: \ee 
158: %
159: where $e$ is the charge of the particle 
160: and $x_0$ is an arbitrary section point. 
161: The momentary current via different sections 
162: is in general not the same. But if we integrate 
163: it over a whole pumping cycle the result becomes 
164: independent of $x_0$.
165: 
166: 
167: The pumping device can be used in 
168: two different configuration. 
169: In the case of an {\em open geometry} (Fig.1) 
170: the leads are attached to reservoirs 
171: that have the same chemical potentials.
172: For simplicity one assumes 
173: a zero temperature Fermi occupation.
174: 
175:  
176: In the case of a {\em closed geometry} (Fig.1)
177: the leads are detached from 
178: the reservoirs and are connected together 
179: so as to have a ring.    
180: This means periodic boundary conditions 
181: over a large space interval $(-L/2)<x<(L/2)$.
182: Furthermore, the closed system is assumed to be 
183: strictly isolated from any environmental influences. 
184: %
185: The closed system can be regarded 
186: as a network with two nodes that 
187: are connected by two bonds one 
188: of length $L_{\tbox{D}}$ (dot region) 
189: and the other of length $L_{\tbox{W}}$ (wire region). 
190: The total length of the ring 
191: is $\mathit{L}=L_{\tbox{D}}+L_{\tbox{W}}$.
192: We assume that $L_{\tbox{D}} \ll L$.
193: 
194: 
195: 
196: 
197: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
198: \subsection{Open (leads) geometry}
199: 
200: The {\em open version} of the two barrier model 
201: has been considered in Ref.\cite{tbm} 
202: using the scattering matrix formalism 
203: of B\"{u}ttiker Pr\'{e}tre and Thomas (BPT) \cite{BPT1,BPT2}. 
204: A typical pumping cycle is illustrated in Fig.2c.
205: In the 1st half of the cycle an electron 
206: is taken from the left lead into the dot region 
207: via the left barrier,  while in the second 
208: half of the cycle an electron is transfered 
209: from the dot region to the right lead  
210: via the right barrier. 
211: Naively, by this peristaltic picture, 
212: it seems that at most one electron 
213: is pumped through the device per cycle. 
214: This expectation is supported by the formal 
215: calculation. Using the BPT formula   
216: one obtains \cite{SAA}
217: %
218: \be{1000}
219: Q \ \ = \ \ (1-\overline{g}_{\tbox{T}}) e
220: \ee
221: %
222: where $0<\overline{g}_{\tbox{T}}<1$ characterizes 
223: the transmission of the device during the charge transfer. 
224: In the limit $\overline{g}_{\tbox{T}} \rightarrow 0$, which 
225: is a pump with {\em no leakage}, indeed one gets $Q=e$. 
226: Otherwise one gets $Q<e$. 
227: 
228: 
229: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
230: \subsection{Closed (ring) geometry}
231: 
232: Our interest is in the {\em closed version} 
233: of the two barrier model. 
234: A major observation is that 
235: the pumped charge~$Q$ is {\em not} ``quantized" 
236: even if the device is closed and isolated from 
237: any environmental influences. 
238: Moreover, it can be larger than unit charge. 
239: In fact we can have $Q \gg e$.   
240: 
241: The analysis that we are going to present 
242: demonstrates and refines general results 
243: that were obtained in Ref.\cite{pmc}. 
244: There we have worked out an artificial circuit 
245: which has been modeled as a $3$~site system. 
246: In the present paper we would like to work out 
247: a major prototype model that allows 
248: the desired comparison between results 
249: for closed circuit as opposed to 
250: that of open geometry [Eq.(\ref{e1000})].  
251: 
252: 
253: We are going to use the Kubo approach 
254: to quantum pumping \cite{pmp,pmc,pme}.    
255: The ``Dirac chains picture" 
256: which we further review in the next subsection   
257: makes a distinction between  
258: ``near field" and ``far field" 
259: pumping cycles. The near field 
260: result has been further considered 
261: in Ref.\cite{pMB} using an extension 
262: of the BPT scattering approach to quantum pumping. 
263: 
264: 
265: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
266: \subsection{The Dirac chains picture}
267: 
268: In order to analyze an {\em adiabatic} \cite{berry1} pumping cycle 
269: we have first to understand the geometry of the parameter space. 
270: In fact the parameter space of the two barrier model  
271: is three dimensional $(X_1,X_2,X_3)$ where $X_3=\Phi$ 
272: is the flux via the ring. In practice we assume 
273: a planar $\Phi=0$ pumping cycle, but for the 
274: theoretical discussion it is convenient 
275: to regard $\Phi$ as a free parameter. 
276: 
277: We ask what is the amount of charge which is transported 
278: via a section of the ring per cycle. For this 
279: purpose we have to calculate the 
280: current $I=\langle\mathcal{I}\rangle$ at each moment. 
281: If we were changing the flux we would have 
282: by Ohm law $I=-G^{33}\dot{\Phi}$ where $G^{33}$ 
283: is called the Ohmic (dissipative) conductance. 
284: But if we change (say) the parameter $X_1$ 
285: then $I=-G^{31}\dot{X_1}$, where  $G^{31}$ 
286: is called the geometric (non-dissipative) 
287: conductance \cite{thouless,AvronNet,berry2}.
288: In general we can write $dQ = -G^{31}dX_1 -G^{32}dX_2$ 
289: and hence     
290: %
291: \be{3}
292: Q \ = \ \oint I dt \ = \ \oint \bm{G} \cdot d\bm{X}
293: \ee 
294: %
295: where $\bm{X}=(X_1,X_2)$ and $\bm{G} = (G^{31},G^{32})$.
296: 
297: 
298: The elements of the conductance matrix $G^{kj}$ 
299: can be calculated using the Kubo formula. 
300: It turns out \cite{berry1,berry2} that in the adiabatic 
301: limit $G^{31}=B_2$  and $G^{32}=-B_1$ 
302: where $\vec{B}$ is the ``magnetic" field (2-form) 
303: which appears in the theory of Berry phase \cite{berry1}.  
304: The sources of this field are Dirac monopoles 
305: that are located at the points of degeneracy. 
306: For the double barrier model a given level $n$ 
307: can have a degeneracy provided $X_1=X_2$,  
308: and either $\Phi=0$ or $\Phi=\pi\hbar/e$ 
309: modulo $2\pi\hbar/e$.
310: In fact we have for each level (excluding the ground state) 
311: two Dirac chains of degeneracies as in Fig.2d. 
312: 
313: 
314: From the above observation one easily 
315: draws the following conclusions:
316: \ \ {\bf (i)} We can get $Q \gg e$ for  
317: a tight cycle around a Dirac chain 
318: if the degeneracy is in the pumping plane. 
319: \ \ {\bf (ii)} We can get $Q \ll e$ for  
320: a tight cycle around a Dirac chain 
321: if the degeneracy is off the pumping plane. 
322: \ \ {\bf (iii)} We can get $Q \sim e$ for 
323: a cycle which is located in the far field 
324: of a Dirac chain. 
325: The existence of a far field region 
326: is not self evident. This constitutes 
327: a major motivation for the present study. 
328:   
329: 
330: 
331: 
332: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
333: \subsection{outline}
334: 
335: The outline of this paper is as follows. 
336: In section~2 we clarify the starting point 
337: of the calculation, which is the Kubo formula.  
338: In section~3 we introduce a preliminary discussion 
339: of the expected results and their significance.
340: %
341: Then in sections~4 to~11 we analyze the 
342: pumping process in the two barrier model. 
343: In particular we find the dependence 
344: of~$Q$ on the ``radius" of the pumping cycle, 
345: and make a distinction between ``near field" 
346: and ``far field" results.    
347:  
348: 
349: 
350: 
351: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
352: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
353: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
354: \section{The Kubo formula}
355: 
356: Given a time dependent 
357: Hamiltonian $\mathcal{H}(X)$ 
358: with $X=X(t)$ we define 
359: $\mathcal{F}=-\partial \mathcal{H}/\partial X$
360: and would like to calculate 
361: the generalized conductance 
362: as defined by the expression
363: %
364: \be{14}
365: \langle \mathcal{I} \rangle =
366: \ \ = \ \ - G \ \dot{X}
367: \ee
368: %
369: We label by $n$ the adiabatic levels 
370: of the closed ring. The adiabatic states  
371: are defined by the equation 
372: $\mathcal{H}|n\rangle = E_n |n\rangle$ 
373: with implicit $X$ dependence.
374: Using these notations the Kubo formula 
375: for the {\em geometric conductance}  
376: can be written as 
377: %
378: \be{15}
379: G = \sum_{m(\ne n)}
380: \frac{ 2\hbar\mathcal{F}_{mn} }
381: {(E_m-E_n)^2}\im[\mathcal{I}_{nm}]
382: \ee
383: %
384: The above expression assumes that only one 
385: level ($n$) is occupied. If several levels 
386: are occupied we have to sum over them.
387: If we have more than one control variable, 
388: say $X_1$ and $X_2$, then we have 
389: to use the more elaborated notation $G^{31}$ 
390: and $G^{32}$ in order to distinguish between 
391: different elements of the conductance matrix.
392: 
393:  
394: If the pumping cycle crosses very close 
395: to a degeneracy, we can get from Eqs.(\ref{e14}-\ref{e15})
396: a very large current $I$, and upon integration 
397: we can find that the transported charge is $Q\gg e$. 
398: In the next sections we shall develop actual 
399: estimates for $Q$. But first we would like 
400: to further illuminate the significance of~$Q$.
401: 
402: 
403: The Schrodinger equation can be written in the {\em adiabatic basis}, 
404: where the {\em transformed} Hamiltonian matrix takes the form 
405: $H_{nm} = E_n \delta_{nm}  + \dot{X} A_{nm}$, 
406: where $E_n$ are called the adiabatic 
407: energy levels, and $A_{nm}$ is a matrix 
408: that can be calculated using a well known recipe 
409: (which is summarized in section~III of \cite{pmc}).   
410: One regards $\dot{X}$ as the ``small parameter". 
411: %
412: If the system is prepared an 
413: instantaneous {\em adiabatic} state $|n\rangle$,  
414: then the instantaneous current is $I=0$, 
415: whereas if it is prepared in an  
416: instantaneous {\em steady} state 
417: (an eigenstate of $H_{nm}$),   
418: then the instantaneous current is finite, say $I=I_0$.
419: Accordingly one can question the physical 
420: relevance of~$Q$: After all typically 
421: the initial preparation is an adiabatic state, 
422: and not a steady state. 
423: 
424: 
425: So let us consider an actual physical scenario. 
426: For simplicity we assume that only two adiabatic 
427: levels are involved: The occupied level $n$ 
428: and next (empty) level $m=n+1$. 
429: To make the dynamical picture simple we use 
430: an analogy with the dynamics of a spin $1/2$ particle,  
431: and consider the illustration in Fig.3.  
432: We regard the state $n$ as spin polarized 
433: in the~$z$ direction, and the state $m$
434: as spin polarized in the $-z$ direction. 
435: The instantaneous steady states of the Hamiltonian 
436: are polarized along an axis that has a small tilt 
437: relative to the~$z$~direction. 
438: 
439: 
440: Initially the spin is polarized in the $z$ direction 
441: and therefore $I(t=0)=0$. 
442: For some time we have $|X(t)-X(0)| \ll \delta X_c$ 
443: where $\delta X_c$ is the relevant parametric scale 
444: for the variation of the adiabatic levels.
445: During this time interval the tilt angle is 
446: approximately constant. The spin is performing 
447: a precession around the tilted axis. 
448: As a results we get $I(t) \ne 0$. 
449: In fact the maximum current is $I(t)=2I_0$.  
450: We get this current after half period of precession. 
451: 
452: 
453: So we have a modulated current $I(t)$ that equals 
454: upon averaging $I_0$. As long as $|X(t)-X(0)| \ll \delta X_c$ 
455: the precession goes on as described above. But 
456: on larger time scales we have to take into account 
457: the variation in the tilt angle. Consequently the 
458: modulation of the current is no longer in the 
459: range $0 < I(t) < 2I_0$, but rather it is shifted 
460: and may increase. Still the average stays approximately $I_0$. 
461: 
462: 
463: Thus we see that in the actual physical scenario 
464: the average over $I(t)$ is the same as 
465: the $I_0$ of the instantaneous steady state. 
466: The validity conditions for this statement 
467: are essentially the validity conditions of 
468: linear response theory, which are further explained  
469: in Ref.\cite{pmc}. The discussion above illuminates 
470: the justification for the use of the first order 
471: steady state solution of Kubo for the purpose of evaluating 
472: the pumped charge in an actual physical scenario.  
473: 
474: 
475: 
476: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
477: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
478: \section{Charge transfer during an avoided crossing}
479: 
480: 
481: The pumped charge $Q$ is obtained via the integral Eq.(\ref{e3}) 
482: using Eq.(\ref{e15}) for $G$. On the basis of the naive heuristic 
483: picture of the Introduction (and see Fig.2) we expect 
484: that most of the contribution to the integral would 
485: come from a small segments in $\bm{X}$ space where the {\em last} 
486: occupied level has an avoided crossing with the first 
487: unoccupied level. Later we define precisely 
488: the region in $\bm{X}$ space where this assumption 
489: is a valid approximation. 
490: 
491: Let us try to sketch what might come out from the 
492: calculation. Later on we are going to do a proper job. 
493: But before we dive into the detailed analysis 
494: (which is quite lengthy) it would be nice to gain 
495: some rough expectation.  
496: %
497: %
498: Given that our interest is focused 
499: in a small energy window such that $E_n \sim E$ 
500: we define $v_{\tbox{E}}=(2E/\mass)^{1/2}$. 
501: The mean level spacing in the energy 
502: range of interest is  
503: %
504: \be{0}
505: \Delta = v_{\tbox{E}} \frac{\hbar\pi}{L}
506: \ee 
507: %
508: while the energy splitting $\Delta_s$  
509: at the avoided crossing might be much smaller.
510: We define the following notations
511: %
512: \be{0}
513: |\mathcal{F}_{nm}| &\equiv& \sigma_0 \\
514: |\mathcal{I}_{nm}| &\equiv& \frac{ev_{\tbox{E}}}{L} \sqrt{g_{\varphi}} \\ 
515: \Delta_s/\Delta &\equiv& \sqrt{1-g_0}
516: \ee
517: %
518: where both $0<g_{\varphi}<1$ and $0<g_0<1$ 
519: are dimensionless. Note that $g_0$ is related 
520: to the overall transmission $\overline{g}_{\tbox{T}}$ 
521: of the device. The adiabaticity condition is 
522: %
523: \be{0}
524: |\dot{X}| \ll \frac{\Delta_s^2}{\hbar\sigma_0} 
525: \ee
526: %
527: and from Eq.(\ref{e14}) with (\ref{e15}) we get the current
528: %
529: \be{0}
530: \langle \mathcal{I} \rangle  = 
531: 2\left(\frac{\hbar\sigma_0}{\Delta_s^2}\dot{X}\right) 
532: \left(\frac{ev_{\tbox{E}}}{L}\right) \sqrt{g_{\varphi}}
533: \ee
534: %
535: The time of the Landau-Zener transition 
536: via the avoided crossing is  
537: %
538: \be{0}
539: \delta t \approx (\Delta_0/\sigma_0)/\dot{X}
540: \ee
541: %
542: and hence the transported charge is 
543: %
544: \be{23}
545: Q  
546: \ \ \approx \ \   
547: \langle \mathcal{I} \rangle \delta t 
548: \ \ \approx \ \  
549: \left(\frac{g_{\varphi}}{1-g_0}\right)^{1/2} e
550: \ee
551: %
552: where $g_0$ and $g_{\varphi}$ should be estimated 
553: in the region of the avoided crossing.  
554: We note that $g_{\varphi}/(1-g_0)$ 
555: is like the Thouless conductance, 
556: and can be regarded as a measure for 
557: the sensitivity of the energy levels 
558: to a test flux. We have pointed out 
559: and discussed this issue in Refs.\cite{pmp,pmc}, 
560: and later it was derived \cite{pMB} in the context 
561: of the scattering formalism.
562: 
563: 
564: In Fig.4 we display the numerically determined $Q$ 
565: for various path segments. The horizontal axis is 
566: the $|X_1-X_2|$ distance of the path segment from 
567: the degeneracy point. As $|X_1-X_2|$ becomes small 
568: $g_0 \rightarrow 1$  and we get $Q \gg e$. 
569: But the asymptotic value $Q \approx e$ which is 
570: observed for large $|X_1-X_2|$ cannot be explained 
571: by such a simple minded calculation. 
572: A major objective of the detailed analysis 
573: is to illuminate the observed crossover. 
574: 
575: 
576: 
577: 
578: 
579: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
580: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
581: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
582: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
583: \section{The model, basic equations}
584: 
585: The one particle Hamiltonian of the two barriers model
586: depends on set of parameters $(X_1, X_2, \Phi)$.
587: From now on we use units such that $e=\hbar=1$, and    
588: characterize the geometry by the dimensionless parameter 
589: %
590: \be{0}
591: b \ \ = \ \ L_{\tbox{W}} / L_{\tbox{D}} \ \ \gg \ \ 1
592: \ee
593: %
594: We write the wavefunction on the two bonds as: 
595: %
596: \be{0}
597: \psi_{\tbox{dot}}(x)
598: &=& 
599: \left({2}/{\mathit{L}}\right)^{1/2} 
600: \sqrt{q_{\tbox{D}}} 
601: \ \sin({\varphi_{\tbox{D}}(x))}
602: \\
603: \psi_{\tbox{wire}}(x)
604: &=& 
605: \left({2}/{\mathit{L}}\right)^{1/2} 
606: \sqrt{q_{\tbox{W}}} 
607: \ \sin({\varphi_{\tbox{W}}(x))}
608: \ee
609: %
610: where $\varphi(x) = k x + \const$. 
611: Given $(X_1,X_2)$ and assuming $\Phi=0$,    
612: the eigenstates of this Hamiltonian can be found by 
613: searching $k_n$ values for which the 
614: matching conditions on the log-derivatives 
615: are satisfied. This leads to the following 
616: system of equations:
617: % 
618: \be{27}
619: && \cot(\varphi_{\tbox{W1}}) + \cot(\varphi_{\tbox{D1}}) = \frac{2\mass}{k} X_1 
620: \\ \label{e28}
621: \label{MatchingEquation2}
622: && \cot(\varphi_{\tbox{W2}}) + \cot(\varphi_{\tbox{D2}}) = \frac{2\mass}{k} X_2 
623: \\
624: && \varphi_{\tbox{D2}} - \varphi_{\tbox{D1}} = kL_{\tbox{D}} 
625: \\
626: && \varphi_{\tbox{W2}} - \varphi_{\tbox{W1}} = kL_{\tbox{W}}
627: \ee
628: %
629: where it should be clear 
630: that $\varphi_{\tbox{D1}} \equiv \varphi_{\tbox{D}}(x_1)$ etc.
631: The corresponding eigenenergies are $E_n=k_n^2/2\mass$. 
632: A similar system of equation can be written for $\Phi=\pi$.
633: We can find the $q_{\tbox{W}}/q_{\tbox{D}}$ ratio for 
634: a given eigenstate via the matching condition 
635: on the wavefunction at either of the two nodes:
636: % 
637: \be{31}
638: \sqrt{q_{\tbox{W}}} \ \sin({\varphi_{\tbox{W}})} 
639: = \sqrt{q_{\tbox{D}}} \ \sin({\varphi_{\tbox{D}})} 
640: \ee
641: %
642: where ${(\varphi_{\tbox{D}}, \varphi_{\tbox{W}})}$ mean 
643: either ${(\varphi_{\tbox{D1}},\varphi_{\tbox{W1}})}$
644: or ${(\varphi_{\tbox{D2}}, \varphi_{\tbox{W2}})}$. 
645: The normalization condition implies that 
646: %
647: \be{32}
648: \frac{L_{\tbox{D}}}{\mathit{L}} q_{\tbox{D}} 
649: + \frac{L_{\tbox{W}}}{\mathit{L}} q_{\tbox{W}} 
650: \ \approx \ 1
651: \ee
652: %
653: For an ``ergodic state" we have $q\approx1$ for both bonds. 
654: In general we characterize an eigenstate by a the mixing parameter 
655: %
656: \be{33}
657: \Theta  
658: \equiv
659: 2 \arctan\left(  \sqrt{ \frac{\mbox{Prob(wire)}}{\mbox{Prob(dot)}} }\right)    
660: =  2 \arctan\left( \sqrt{b}
661: \left(\frac{q_{\tbox{W}}}{q_{\tbox{D}}}\right)^{1/2}  \right)
662: \ee
663: %
664: such that $\Theta=0$ means a definite dot state, 
665: while $\Theta=\pi$ means a definite wire state. 
666: In practice we have ${0<\Theta<\pi}$. 
667: 
668: 
669: In the numerical analysis we use units 
670: such that $\mass=1$ and $L_{\tbox{D}}=1$. 
671: Given $X_1$ and $X_2$ we solve the system 
672: of equations for the $k_n$ and for 
673: the $\varphi^{(n)}$ at the nodes. 
674: Then we determine $q^{(n)}$ 
675: at each bond, and from the ratio we 
676: get $\Theta^{(n)}$ as well. 
677: 
678: 
679: 
680: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
681: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
682: \section{Regions in $\bm{X}$ space}
683: 
684: 
685: We focus on a small energy window 
686: such that the energy levels of interest 
687: are $E < E_n < E+dE$. We characterize 
688: a barrier $X\delta()$ using its transmission 
689: %
690: \be{0}
691: g(X) = \left[ 1 + \left(\frac{1}
692: {v_{\tbox{E}}}X \right)^2 \right]^{-1} 
693: \approx 
694: \left(\frac{1} {v_{\tbox{E}}}X \right)^{-2}
695: \ee  
696: %  
697: The last expression holds if $g(X) \ll 1$.
698: We can regard $g_1=g(X_1)$ and $g_2=g(X_2)$ 
699: as an alternate way to specify $X_1$ and $X_2$.  
700: % 
701: %
702: The $(X_1,X_2,\Phi)$ space is divided 
703: into various regions (Fig.5a). There is a region 
704: where $g_1$ and $g_2$ are of order one 
705: ($|1-g| \ll 1$). There the delta functions 
706: at the nodes can be treated as a small perturbation. 
707: There is a region where ${1/b \ll g_1,g_2 \ll 1}$. 
708: There each dot level mix with many wire 
709: levels. This intermediate region will 
710: be discussed in a future work \cite{pmm}. 
711: Finally there is the region in $\bm{X}$ space 
712: which is of interest in the present study:
713: %
714: \be{34}
715: g_1,g_2 \ \ \ll \ \ 1/b  
716: \ee
717: %
718: We shall argue that in this region 
719: the states are categorized into 
720: ``wire states" and ``dot state", 
721: which mix only whenever the energy 
722: level of the dot ``cross" 
723: an energy level of the wire (Fig.5bcd). 
724: This allows to use ``two level" 
725: approximation in the analysis of the mixing. 
726:     
727: 
728: The degeneracies of the Hamiltonian occur 
729: at points $(X^{(r)},X^{(r)})$ along the
730: symmetry axis of $\bm{X}$ space. They are divided 
731: into two groups: those that are located 
732: in the $\Phi=0 (\mbox{mod}(2\pi))$ planes and those 
733: that are located in the $\Phi=\pi(\mbox{mod}(2\pi))$ planes.
734: In Appendix~A we find the explicit expression 
735: for $X^{(r)}$. It should be clear that each degeneracy 
736: point is duplicated $\mbox{mod}(2\pi)$ 
737: in the $\Phi$ direction, hence creating 
738: what we call a ``Dirac chain".    
739: 
740: 
741: There are only two non-trivial  
742: degeneracies in $\bm{X}$ space which are 
743: associated with a given level $n$. 
744: One is with the neighboring level 
745: ``from above" and the other is with the 
746: neighboring level ``from below".  
747:  
748: 
749: Once we locate a degeneracy point  
750: we can make a distinction between ``near field" 
751: and ``far field" regions. 
752: The near field is defined as the region 
753: where we can use degenerate perturbation 
754: theory in order to figure out the 
755: splitting of the levels. 
756: In contrast to that, 
757: the far field is defined as the region 
758: where we can calculate the 
759: splitting of the levels 
760: by treating the dot-wire coupling 
761: as a first order perturbation. 
762: 
763: 
764: In Fig.5a we show one pumping path that 
765: crosses in the near field region, 
766: and a second pumping path that 
767: crosses in the far field region.
768:  
769: 
770: 
771: 
772: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
773: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
774: \section{Eigenstate analysis}
775: 
776: 
777: In the theoretical analysis it is illuminating 
778: to map the behavior of $k_n$ and $\Theta^{(n)}$ 
779: in $(X_1,X_2,\Phi)$ space. See Figs.6-7.
780: It is not difficult to realize that the 
781: variation of $\Theta$ is bounded as follows:
782: %
783: \be{35}
784: {\sqrt{b}}\left[\frac{1}{4}g\right]^{+1/2}   
785: \ \ < \ \ \tan(\Theta/2) \ \ < \ \  
786: {\sqrt{b}}\left[\frac{1}{4}g\right]^{-1/2} 
787: \ee
788: %
789: where $g$ is either $g_1$ or $g_2$.  
790: The derivation of this result is as follows: 
791: The matching conditions at a given node 
792: implies that $(\varphi_{\tbox{D}}, \varphi_{\tbox{W}})$ 
793: are constraint to be on one of two 
794: branches which are illustrated in Fig.8. 
795: With each point of a given branch we can 
796: associate a $\Theta$ value via Eq.(\ref{e33})
797: and either Eq.(\ref{e27}) or Eq.(\ref{e28}). 
798: It is a straightforward exercise to 
799: express $\Theta$ say as a function 
800: of $\varphi_{\tbox{D}}$, and then to find 
801: its minimum and maximum values. Assuming 
802: that $g\ll 1$ one obtains Eq.(\ref{e35}). 
803: 
804:  
805: 
806: 
807: Given $\Theta^{(n)}$ we can extract 
808: what are the $q^{(n)}$ at each bond, 
809: and what are the $\varphi^{(n)}$ at the nodes. 
810: By solving Eq.(\ref{e32}) with Eq.(\ref{e33}) we get  
811: %
812: \be{36}
813: q_{\tbox{D}} &\ \ = \ \ & b \ (\cos(\Theta/2))^2
814: \\
815: \label{e37}
816: q_{\tbox{W}} &\ \ = \ \ & (\sin(\Theta/2))^2
817: \ee
818: %
819: then from Eq.(\ref{e31}) 
820: with either Eq.(\ref{e27}) or Eq.(\ref{e28}) 
821: we find an expression for $\varphi$ at a given node. 
822: In particular we get 
823: %
824: \be{38}
825: \varphi_{\tbox{D}} = 
826: \left( 1 \pm \frac{1}{\sqrt{b}}\tan(\Theta/2) + ... \right) 
827: \left[\frac{1}{4}g\right]^{1/2} 
828: \ee
829: %
830: Note that $\varphi$ is well defined $\mbox{mod}(\pi)$. 
831: There are states which are dot-like ($\Theta \ll 1$),   
832: and there are states which are wire-like ($\Theta \sim \pi$). 
833: There are regions where a dot-like state mix with 
834: wire-like state leading to pair of states 
835: with $\Theta^{(+)} \sim \Theta^{(-)} \sim \pi/2$.     
836: From the above formula it follows that in the latter case 
837: %  
838: \be{39}
839: &&\varphi_{\tbox{D}}^{(+)} 
840: \ \ \approx \ \ \varphi_{\tbox{D}}^{(-)} 
841: \ \ \approx \ \ \left[\frac{1}{4}g\right]^{1/2} 
842: \\ \label{e40}
843: \label{varphi_D_Minus_varphi_D}
844: &&|\varphi_{\tbox{D}}^{(+)} - \varphi_{\tbox{D}}^{(-)}| 
845: \ \ = \ \ 
846: \frac{1}{\sqrt{b}} \left[\frac{1}{4}g\right]^{1/2}
847: \left( \tan(\Theta^{(+)}/2)   + 
848: \tan(\Theta^{(-)}/2) \right)
849: \ee
850: 
851: 
852: 
853: 
854: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
855: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
856: \section{Eigenenergies and "mixing" from perturbation theory}
857: 
858: In order to find the splitting and the mixing of the two levels 
859: we use perturbation theory once in the far field analysis 
860: and once in the near field analysis. 
861: In both cases we use the following notations.  
862: The unperturbed basis is $\vert i \rangle$ with a dot-like state 
863: $|1\rangle$ and a wire-like $|2\rangle$. 
864: The perturbed eigenstates $\vert n \rangle$  
865: are indicated as $|+\rangle$ and $|-\rangle$. 
866: The Hamiltonian in both cases has the general form
867: %
868: \be{0}
869: \mathcal{H}=\mathcal{H}_0+\bm{W}=\left(
870: \begin{array}{cc}
871: E_1 & 0   \\
872: 0   & E_2
873: \end{array}
874: \right)
875: +\left(
876: \begin{array}{cc}
877: W_{11} & W_{12}   \\
878: W_{21} & W_{22}
879: \end{array}
880: \right)
881: \ee
882: %
883: The eigenvectors are real so $W_{12}=W_{21}$.
884: The Hamiltonian can be written 
885: as a linear combination of Pauli matrices
886: %
887: \be{0}
888: \mathcal{H}=\left(\frac{E_1 + E_2}{2} + \frac{W_{11}+W_{22}}{2}\right){\bm 1}
889: +\left( \frac{E_1 - E_2}{2} + \frac{W_{11}-W_{22}}{2} \right)
890: {\bm \sigma}_3+W_{12}{\bm \sigma}_1
891: \ee
892: %
893: we define
894: %
895: \be{43}
896: \Delta_s &=& 2\sqrt{\left( \frac{E_1 - E_2}{2} + \frac{W_{11}-W_{22}}{2} \right)^2 + W_{12}^2} 
897: \\
898: \label{e44}
899: \tan(\theta) &=& \frac{2W_{12}}{(W_{11}-W_{22}) + (E_1 - E_2)}
900: \ee
901: %
902: The eigenenergies are
903: %
904: \be{45}
905: E_n =
906: \left(\frac{E_1 + E_2}{2} + \frac{W_{11}+W_{22}}{2}\right) 
907: \pm\frac{\Delta_s}{2} 
908: \ee
909: %
910: The eigenstates are found by rotating 
911: a spin half around the $y$ axis at an angle~$\theta$
912: %
913: \be{0}
914: |+\rangle
915: &\longrightarrow&
916: \left( \begin{array}{c} 
917: \cos{(\theta/2)} \\ \sin{(\theta/2)}
918: \end{array} \right)\\
919: \label{nVector}
920: |-\rangle
921: &\longrightarrow&
922: \left( \begin{array}{c} 
923: -\sin{(\theta/2)} \\ \cos{(\theta/2)}
924: \end{array} \right)
925: \ee
926: %
927: Assuming that the unperturbed basis 
928: consists of distinct dot-like 
929: and wire-like states, it follows that    
930: we can use the following approximation:
931: %
932: \be{48}
933: \Theta^{(+)} \ \ &=& \ \ \theta \\
934: \label{e49}
935: \Theta^{(-)} \ \ &=& \ \ \pi-\theta
936: \ee
937: %
938: Using Eq.(\ref{e36}) and Eq.(\ref{e40})
939: this implies
940: %
941: \be{50}
942: \sqrt{q_{\tbox{D}}^{(+)} q_{\tbox{D}}^{(-)}} 
943: \ \ &\approx& \ \ \frac{b}{2}|\sin(\theta)| 
944: \\ \label{e51}
945: %%%%%%%% \ \ = \ \  b\frac{|W_{12}|}{\Delta_s} 
946: |\varphi_{\tbox{D}}^{(+)} - \varphi_{\tbox{D}}^{(-)}|
947: \ \ &=& \ \ \frac{2}{\sqrt{b}} \left[\frac{1}{4}g\right]^{1/2}
948: \frac{1}{|\sin(\theta)|}
949: \ee
950: %
951: We note that in the strong mixing region we have $\theta\approx\pi/2$. 
952: 
953: In the next sections we obtain explicit expressions 
954: for the ``splitting" and the ``mixing" using the above 
955: scheme. From the numerics (see e.g. Fig.7) we see
956: that these are in fact very satisfactory approximations.  
957: 
958: 
959: 
960: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
961: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
962: \section{Far field analysis}
963: 
964: The eigenstates of the ring in the $\bm{X}$ region 
965: of interest as defined in Eq.(\ref{e34}) can be found 
966: using first order perturbation theory with respect 
967: to the wire-dot coupling. This is explained below. 
968: We have verified numerically that the approximation 
969: is remarkable unless we are very close to 
970: a degeneracy point. This defines our distinction 
971: between ``far" and ``near" field regions. 
972: In the latter case we present in the next section 
973: a complementary treatment using degenerate 
974: perturbation theory.  
975:  
976: The unperturbed Hamiltonian in the far field 
977: analysis corresponds to ${X_1=X_2=\infty}$. 
978: Using the notations as defined in the previous  
979: section we take $|2\rangle$ as the $n^{\tbox{th}}$ 
980: wire state, and $|1\rangle$ as the closest 
981: dot state from above. Hence 
982: %
983: \be{52}
984: E_1 &=&  \frac{1}{2\mass} \left(\frac{\pi}{L_{\tbox{D}}}(1+[n/b]_{\tbox{integer}})\right)^2  \\
985: \label{e53}
986: E_2 &=&  \frac{1}{2\mass} \left(\frac{\pi}{L_{\tbox{W}}}n\right)^2
987: \ee
988: %
989: Using the formula
990: %
991: \be{0}
992: W_{ij} \ \ = \ \ -\frac{1}{4\mass^2 X} [\partial \psi^{(i)}] [\partial \psi^{(j)}] 
993: \ee
994: % 
995: we get:
996: %
997: \be{0}
998: W_{11} &=& -\frac{v_{\tbox{E}}^2}{2L_{\tbox{D}}} 
999: \ \left[\frac{1}{X_1} + \frac{1}{X_2} \right] \\
1000: W_{22} &=& -\frac{v_{\tbox{E}}^2}{2L_{\tbox{W}}}
1001: \ \left[\frac{1}{X_1} + \frac{1}{X_2} \right] \\
1002: \label{e57}
1003: |W_{12}| &=& \frac{v_{\tbox{E}}^2}{2\sqrt{L_{\tbox{D}} L_{\tbox{W}}}} 
1004: \ \left|\frac{1}{X_1} \pm \frac{1}{X_2} \right| 
1005: \ee
1006: %
1007: where the $\pm$ sign in the expression for the dot-wire 
1008: coupling depends on whether the dot and the wire states 
1009: have the same parity or not. 
1010: 
1011: In case of a far field pumping cycle we start 
1012: (say) with very high barriers, and then lower 
1013: one of them, say $X_1$. If we neglected  
1014: the dot-wire coupling $W_{12}$, the dot level   
1015: would cross the wire level at a point $X_1=X^{(n)}$ 
1016: that can be determined from the equation $E_1+W_{11}=E_2$.
1017: At the vicinity of the avoided crossing we obtain 
1018: %
1019: \be{0} 
1020: |W_{12}| = \frac{1}{2\pi} (b g^{(n)})^{1/2} \Delta 
1021: \ee
1022: %
1023: where
1024: %
1025: \be{0}
1026: g^{(n)} = \left(\frac{1}{v_{\tbox{E}}}X^{(n)}\right)^{-2} 
1027: \ee
1028: %
1029: From the condition $|W_{12}|\ll \Delta$  
1030: we deduce Eq.(\ref{e34}) which defines 
1031: our $\bm{X}$ region of interest. 
1032: Furthermore, from the results of the previous 
1033: section we obtain expressions for the 
1034: splitting and for the mixing of the levels:  
1035: %
1036: \be{60}
1037: \Delta_s  &=&  bg^{(n)} \ \frac{1}{2L} ||\bm{X}-\bm{X}^{(r)}|| \\
1038: \sin(\theta)  &=&  \frac{2}{\sqrt{b g^{(n)})}} 
1039: \ \frac{v_{\tbox{E}}}{||\bm{X}-\bm{X}^{(r)}||}
1040: \ee
1041: %
1042: We use the following notation, which 
1043: we regard as a measure for the distance 
1044: of the pumping cycle from the nearby degeneracy: 
1045: %
1046: \be{0}
1047: ||\bm{X}-\bm{X}^{(r)}||  = \sqrt{(X_1 - X^{(n)})^2 
1048: + \frac{4}{b}\left(\frac{v_{\tbox{E}}}{\sqrt{g^{(n)}}}\right)^2}
1049: \ee
1050: % 
1051: The significance of this notation   
1052: will be further clarified in the next 
1053: section where we extend the analysis 
1054: into the near field region. 
1055: 
1056: 
1057: 
1058: 
1059: 
1060: 
1061: 
1062: 
1063: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1064: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1065: \section{Near field analysis}
1066: 
1067: The unperturbed Hamiltonian in the near field 
1068: analysis corresponds to ${X_1=X_2=X^{(r)}}$. 
1069: We can find a rough approximation for $X^{(r)}$ 
1070: using the analysis of the previous section, 
1071: but in fact we can find the exact expression  
1072: which is derived in Appendix~A, 
1073: where we also define the obvious 
1074: notations $k_r$ and $g^{(r)}$. 
1075: For later calculation the following 
1076: approximation is useful: 
1077: %
1078: \be{0}
1079: X^{(r)} \ \ \approx \ \ \frac{v_{\tbox{E}}}{\sqrt{g^{(r)}}}
1080: \ee
1081: % 
1082: The unperturbed basis consists of 
1083: the dot-like definite 
1084: parity state $|1\rangle$, 
1085: and the wire-like definite parity state $|2\rangle$. 
1086: We recall that for these states $\Theta$ 
1087: attains its extremal values as remarked 
1088: at the end of Appendix~A. 
1089: The energies of the unperturbed states are
1090: %
1091: \be{0}
1092: E_1 = E_2 = E_r = \frac{1}{2\mass}k_r^2
1093: \ee
1094: %
1095: and the perturbation matrix is    
1096: %
1097: \be{0}
1098: W_{11} &=& bg^{(r)} \frac{1}{2L} \left( \delta X_1 + \delta X_2 \right)\\
1099: W_{22} &=& g^{(r)} \frac{1}{2L} \left( \delta X_1 + \delta X_2 \right)\\
1100: |W_{12}| &=& \sqrt{b}g^{(r)} \frac{1}{2L} | \delta X_1 - \delta X_2 |
1101: \ee
1102: %
1103: where
1104: %
1105: \be{0}
1106: \delta X_1 &=& X_1 - X^{(r)} \\
1107: \delta X_2 &=& X_2 - X^{(r)} 
1108: \ee
1109: %
1110: Consequently we can determine both the 
1111: splitting and the mixing of the levels:
1112: %
1113: \be{0}
1114: \Delta_s  &=&  bg^{(r)}\frac{1}{2L} ||\bm{X}-\bm{X}^{(r)}|| \\
1115: \sin(\theta)  &=& 
1116: \frac{2}{\sqrt{b}} \  
1117: \frac{|X_1-X_2|}{||\bm{X}-\bm{X}^{(r)}||}
1118: \ee
1119: %
1120: In the above expression we have extended 
1121: the interpretation of the distance measure 
1122: as follows:
1123: %
1124: \be{0}
1125: ||\bm{X}-\bm{X}^{(r)}||  =
1126: \sqrt{
1127: \left( {X_1+X_2}-2X^{(r)}\right)^2 +
1128: \frac{4}{b} |X_1- X_2|^2}
1129: \ee
1130: %
1131: Now we realize that the far field results of the 
1132: previous section are formally a saturated 
1133: version of the near field results with
1134: %
1135: \be{0} 
1136: |X_1- X_2|_{\infty} =  \frac{v_{\tbox{E}}}{\sqrt{g^{(n)}}}
1137: \ee
1138: 
1139: 
1140: 
1141: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1142: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1143: \section{The expressions for $G$}
1144: 
1145: We are now ready to calculate $G$ from Eq.(\ref{e15}).
1146: There are of course $G^{31}$ and $G^{32}$ but 
1147: the expressions look alike so we concentrate, 
1148: say, on the case $X=X_1$, and suppress the 
1149: node indication subscript whenever possible. 
1150: By definition $\mathcal{F}
1151: =-\partial \mathcal{H}/\partial X
1152: =\delta(\hat{x}-x_1)$. 
1153: The current operator has been defined 
1154: in Eq.(\ref{e2}), but we still have the freedom 
1155: to set $x_0$ as we want. So the natural 
1156: choice for sake of calculation, 
1157: is obviously $x=x_1$. We shall expand later 
1158: on the results that would be obtained 
1159: if the current were measured via a different 
1160: section. The matrix elements of the operators 
1161: involved are 
1162: %
1163: \be{0}
1164: \label{F1Definition}
1165: \mathcal{F}_{nm} 
1166: &=& 
1167: -\frac{2}{L} 
1168: \sqrt{q_{\tbox{D}}^{(n)} q_{\tbox{D}}^{(m)} }
1169: \ \sin(\varphi_{\tbox{D}}^{(n)}) 
1170: \ \sin(\varphi_{\tbox{D}}^{(m)}) 
1171: \\
1172: \nonumber
1173: \mathcal{I}_{nm} 
1174: &=& 
1175: i\frac{e}{\mass L} 
1176: \sqrt{ q_{\tbox{D}}^{(n)} q_{\tbox{D}}^{(m)} } 
1177: \left[
1178: \frac{k_n{+}k_m}{2} \sin{(\varphi_{\tbox{D}}^{(m)} {-} \varphi_{\tbox{D}}^{(n)})}  
1179: + 
1180: \frac{k_n{-}k_m}{2} \sin{(\varphi_{\tbox{D}}^{(m)} {+} \varphi_{\tbox{D}}^{(n)})}
1181: \right]
1182: \ee
1183: %
1184: One should notice that $\mathcal{F}_{nm}$ is real and symmetric 
1185: with respect to $n\leftrightarrow m$ interchange, 
1186: while $\mathcal{I}_{nm}$ is antisymmetric and purely 
1187: imaginary as implied by time reversal symmetry.
1188: Once we sum  Eq.(\ref{e15}) over all the occupied levels, 
1189: $nm$ terms cancel with $mn$ terms.  
1190: Within the framework of the ``two level approximation" 
1191: the only remaining term involves the occupied level $n$ 
1192: and the next empty level $m=n+1$
1193: % 
1194: \be{0}
1195: \label{G1}
1196: G^{31}(X_1, X_2) =  2 \frac{ \mathcal{F}_{mn} \ \im[\mathcal{I}_{nm}] } { \Delta_s^2 }
1197: \ee
1198: %
1199: We recall the following expressions:  
1200: % 
1201: \be{0}
1202: \sqrt{q_{\tbox{D}}^{(+)} q_{\tbox{D}}^{(-)}} 
1203: \ \ &\approx& \ \ \frac{b}{2}|\sin(\theta)|
1204: \\  
1205: \varphi_{\tbox{D}}^{(+)} 
1206: \ \ \approx \ \ \varphi_{\tbox{D}}^{(-)} 
1207: \ \ &\approx& \ \ \left[\frac{1}{4}g\right]^{1/2} 
1208: \\
1209: |\varphi_{\tbox{D}}^{(+)} - \varphi_{\tbox{D}}^{(-)}| 
1210: \ \ &\approx& \ \ 
1211: \frac{2}{\sqrt{b}} \left[\frac{1}{4}g\right]^{1/2}
1212: \frac{1}{|\sin(\theta)|}
1213: \ee
1214: %
1215: Upon substitution we realize that both in the near 
1216: and in the far field we can neglect the second 
1217: term in the expression for $\mathcal{I}_{nm}$. Consequently  
1218: %
1219: \be{0}
1220: G^{31}(X_1, X_2) 
1221: \ = \
1222: -\frac{1}{4}\left(gb\right)^{3/2}
1223: \frac{ev_{\tbox{E}}}{L^2} 
1224: \ \frac{1}{\Delta_s^2}
1225: \ |\sin(\theta)| 
1226: \ = \
1227: -\frac{2 \ ev_{\tbox{E}}}{b\sqrt{g}} \ 
1228: \frac{|X_1 - X_2|}{||\bm{X}-\bm{X}^{(r)}||^3}
1229: \ee
1230: %
1231: We observe that in the near field, 
1232: where $||\bm{X}-\bm{X}^{(r)}||$ is essentially 
1233: the Euclidean measure of distance, 
1234: we get the field of a Dirac monopole 
1235: as expected. But as we go to the far field 
1236: the $|X_1-X_2|$ contribution saturates 
1237: as explained in the previous section.     
1238: 
1239: 
1240: 
1241: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1242: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1243: \section{The calculation of $Q$}
1244: 
1245: Having found $G^{31}$ and a similar 
1246: expression for $G^{32}$ we can perform 
1247: the integration of Eq.(\ref{e3}) in order 
1248: to obtain $Q$. We already pointed out 
1249: that most of the contribution 
1250: in the regions of our interest 
1251: comes from the avoided crossings. 
1252: In the near field  it is most convenient 
1253: to make the integration along 
1254: $X_1-X_2=\pm\const$ segments,  
1255: while in case of the far field 
1256: we make the integration along   
1257: $X_2=\infty$ and $X_1=\infty$ segments. 
1258: 
1259: 
1260: It is not difficult to realize that 
1261: in the near field calculation both 
1262: segments ($X_1-X_2=\pm\const$) give 
1263: the same contribution. This means that 
1264: we simply have to do one of the integral 
1265: and to multiply by~$2$. 
1266: But in the far field one should 
1267: be more careful. If we measure 
1268: the current in the {\em middle} of the dot 
1269: (as indeed done in the numerics) 
1270: then the same rule applies. But if 
1271: we measure the current (say) at node~$1$, 
1272: then the predominant contribution to $Q$    
1273: comes obviously from the $X_2=\infty$ segment, 
1274: so the result should not be multiplied by~$2$. 
1275: %
1276: %
1277: %
1278: Performing a straightforward calculation 
1279: of the $dX$ integral, using   
1280: %
1281: \be{0}
1282: \int_{-\infty}^{+\infty}
1283: \frac{dx}{\left( x^2 + a^2 \right)^{3/2}} = \frac{2}{a^2}
1284: \ee
1285: %
1286: and taking the above discussion into account, 
1287: we get in the near field 
1288: %
1289: \be{81}
1290: Q \ \ \approx \ \ \frac{X^{(r)}}{|X_1-X_2|} e
1291: \ee
1292: %
1293: One can show that this result is in agreement 
1294: with the rough estimate of Eq.(\ref{e23}). 
1295: However, in the far field we have to substitute 
1296: the saturated value of  $|X_1-X_2|$ 
1297: leading to the result $Q \approx e$. 
1298: 
1299: 
1300: The calculation in the far field does not care 
1301: whether the pumping cycle encircle a $\Phi=0$ 
1302: degeneracy or a  $\Phi=\pi$ degeneracy. 
1303: It is only in the near field of $X^{(r)}$ 
1304: that we see the difference. This is clearly 
1305: confirmed by the numerics (Fig.4). 
1306: But a closer look reveals that the far field 
1307: numerical result for $Q$ is somewhat 
1308: smaller than~$1$. This might look like 
1309: a contradiction with respect to the general expectations. 
1310: The resolution of this puzzle is related 
1311: to the limitation of the far field perturbative 
1312: treatment. Within the framework of this 
1313: treatment $\Theta^{(n)}$ changes 
1314: from $\Theta^{(n)}=\pi$ to $\Theta^{(n)}=0$ 
1315: as we lower (say) the $X_1$ barrier.
1316: But in fact we know from section~8 that $\Theta$ 
1317: is bounded. This means that not all the 
1318: the probability gets into the dot region. 
1319: Consequently,  if we integrate 
1320: along the $X_2=\infty$ segment, 
1321: we expect to get $Q<e$ as observed. 
1322: On the other hand, in case of a full 
1323: pumping cycle, we have to cross 
1324: from the $X_2=\infty$ segment  
1325: to the  $X_1=\infty$ segment. 
1326: This was neglected in our calculation. 
1327: Thus if we had a full cycle we would 
1328: expect to get $Q\approx e$ in the far field 
1329: as implied by the Dirac chains picture.
1330:    
1331: 
1332: 
1333: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1334: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1335: \section{Discussion}
1336: 
1337: We were able to derive an estimate for $Q$ 
1338: in the case where the pumping cycle 
1339: is dominated by a single degeneracy. 
1340: Within this framework 
1341: we can still distinguish between 
1342: near field (where $Q\gg e$) 
1343: and far field (where $Q\sim e$) regions.   
1344: %
1345: %
1346: Our results are in agreement with 
1347: those of Ref.\cite{pmc}. We note that 
1348: an optional derivation of the near field 
1349: limit has been introduced in Ref.\cite{pMB} 
1350: using an extension of the BPT scattering 
1351: formalism. But the latter derivation  
1352: was not suitable to reproduce the far field  
1353: result because it has been {\em assumed} there  
1354: that the charge cannot accumulate in the dot region.  
1355: 
1356: 
1357: It should be re-emphasized that the results 
1358: that we have obtained assume that the pumping 
1359: cycle is dominated by a single degeneracy. 
1360: In a follow up work \cite{pmm} we shall 
1361: discuss the case where the charge transport 
1362: involves many levels, such that the 
1363: contribution of neighboring levels (in Eq.(\ref{e15})) 
1364: is negligible compared with contribution 
1365: that comes from $|m-n|>1$ levels.
1366: 
1367: 
1368: The results that we obtain for a closed geometry are very 
1369: different from those that are obtained for an open 
1370: geometry. This is because the motion of the electron 
1371: is ``recycled". In a more technical language this means 
1372: that multiple rounds should be taken into account in 
1373: the calculation of correlation functions. 
1374: Refs.\cite{pmo,pme} further discuss how the Kubo formula 
1375: can be used in order to interpolate these two extreme 
1376: circumstances.  
1377:   
1378: 
1379: It should be clear that adiabatic transport becomes  
1380: counter-intuitive if one adopts a misleading 
1381: zero order point of view of the adiabatic process.
1382: Moreover, even within the ``two level approximation" 
1383: it would not be correct to say that $Q$ is determined 
1384: by {\em peristaltic} mechanism: 
1385: namely, it is not correct to say 
1386: that charge transfer is regulated   
1387: by the Landau-Zener transitions. 
1388: 
1389: 
1390: A {\em peristaltic} mechanism would imply $Q\sim e$.  
1391: In the near field we have  $Q\gg e$ so we do not have 
1392: such  mechanism for sure. This is also reflected 
1393: by having the same~$I$ at both nodes as discussed 
1394: in the paragraph of Eq.(\ref{e81}).
1395: However, even in the far field, 
1396: where the peristaltic picture seems natural, 
1397: we have realized that it is an over simplification:    
1398: also in the case of a far field cycle a finite 
1399: fraction of $Q$ is contributed during 
1400: the intermediate stages of the pumping cycle.
1401: 
1402: 
1403: \appendix
1404: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1405: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1406: \section{Finding the degeneracy points}
1407: 
1408: 
1409: If system is symmetric ($X_1=X_2=X$) then we can distinguish between 
1410: odd and even states leading to the following eigenvalue equations 
1411: %
1412: \be{101}
1413: \cot(kL_{\tbox{D}}/2) + \cot(kL_{\tbox{W}}/2) 
1414: &= -\frac{2\mass}{k}X   
1415: & \,\,\, \mbox{odd states} 
1416: \\ \label{e102} 
1417: \tan(kL_{\tbox{D}}/2) + \tan(kL_{\tbox{W}}/2) 
1418: &= +\frac{2\mass}{k}X   
1419: & \,\,\, \mbox{even states}
1420: \ee
1421: %
1422: As we lower $X$ we have an exact crossing whenever 
1423: a dot state crosses a wire state with the opposite parity. 
1424: We have an avoided crossing whenever a dot state 
1425: tries to cross a wire state with the same parity.  
1426: The later becomes an exact crossing  
1427: if the flux through the ring is half integer.  
1428: 
1429: 
1430: We can determine the degeneracy points by 
1431: equating (\ref{e101}) with (\ref{e102}).  
1432: This gives $\sin(kL_{\tbox{W}})=-\sin(kL_{\tbox{D}})$. 
1433: For half integer flux it is 
1434: convenient to use delta gauge on the middle of the wire, 
1435: so as to get there $\pi$ phase jump boundary conditions. 
1436: This implies that in the above equation we make 
1437: the replacement $(kL/2) \mapsto (kL/2) + (\pi/2)$, 
1438: hence getting the degeneracy condition  
1439: $\sin(kL_{\tbox{W}})=+\sin(kL_{\tbox{D}})$. 
1440: We therefore conclude that we have degeneracies for 
1441: %
1442: \be{0}
1443: k_r \ \ = \ \ \frac{\pi}{L_{\tbox{W}}-L_{\tbox{D}}} n_r
1444: \ee 
1445: %
1446: They are categorized into $\Phi{=}0$ degeneracies 
1447: for $n_r=1,3,5,\dots$ and $\Phi{=}\pi$ degeneracies 
1448: for $n_r=2,4,6,\dots$. 
1449: Their location is $(X^{(r)},X^{(r)})$ where   
1450: %
1451: \be{0}
1452: X^{(r)} \ \ = \ \ -\frac{k_r}{\mass}\cot(k_rL_{\tbox{D}}) 
1453: \ee
1454: % 
1455: Accordingly
1456: %
1457: \be{0}
1458: g^{(r)} \ \ = \ \ g(X^{(r)})
1459: \ \ = \ \ \sin^2(k_rL_{\tbox{D}}) 
1460: \ \ = \ \  \sin^2(k_rL_{\tbox{W}})
1461: \ee
1462: % 
1463: At a degeneracy point the mixing parameter $\Theta$ 
1464: that characterizes the odd and the even states 
1465: attains the extremal values which are allowed by Eq.(\ref{e35}). 
1466: This can be verified by deducing $q_{\tbox{W}}/q_{\tbox{D}}$ 
1467: from Eq.(\ref{e31}) with ${\varphi_{\tbox{D}}=k_rL_{\tbox{D}}/2}$ for the 
1468: odd state and ${\varphi_{\tbox{D}}=(\pi/2)+k_rL_{\tbox{D}}/2}$ 
1469: for the even state.
1470: 
1471: 
1472: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1473: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1474: \ack
1475: 
1476: D.C. thanks M.~Moskalets and M.~B{\"u}ttiker 
1477: for discussions that had motivated this work,  
1478: and Y.~Oreg for urging clarification of the 
1479: formal result. We are grateful to 
1480: T.~Kottos, H.~Schanz and G.~Rosenberg 
1481: for helpful suggestions and help with the 
1482: numerical procedure. The research was supported 
1483: by the Israel Science Foundation (grant No.11/02),
1484: and by a grant from the GIF, the German-Israeli Foundation 
1485: for Scientific Research and Development.
1486: 
1487: 
1488: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1489: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1490: \Bibliography{99}
1491: 
1492: \bibitem{rev}
1493: L.P. Kouwenhoven, C.M. Marcus, P.L. Mceuen,
1494: S. Tarucha, R. M. Westervelt and N.S. Wingreen,
1495: Proc. of Advanced Study Inst. on Mesoscopic
1496: Electron Transport, edited by L.L. Sohn,
1497: L.P. Kouwenhoven and G. Schon (Kluwer 1997).
1498: 
1499: 
1500: 
1501: \bibitem{pmp}
1502: D. Cohen, cond-mat/0208233 (2002); 
1503: Solid State Communications {\bf 133}, 583 (2005).
1504: 
1505: \bibitem{pmc}
1506: D. Cohen, Phys. Rev. B {\bf 68}, 155303 (2003).
1507: 
1508: \bibitem{pme}
1509: For a mini-review see: D. Cohen, Physica E 28, 308 (2005).
1510: 
1511: \bibitem{pml} 
1512: G.~Rosenberg and D.~Cohen, 
1513: cond-mat/0510289, J. Phys. A (2006, in press).
1514: 
1515: 
1516: 
1517: 
1518: \bibitem{tbm}
1519: Y. Levinson, O. Entin-Wohlman and P. Wolfle, cond-mat/0010494.
1520: 
1521: \bibitem{BPT1}
1522: M. Buttiker, H. Thomas and A Pretre, 
1523: Z.~Phys.~B-Condens.~Mat., {\bf 94}, 133 (1994).  
1524: 
1525: \bibitem{BPT2}
1526: P. W. Brouwer, Phys. Rev. B {\bf 58}, R10135 (1998).
1527: 
1528: \bibitem{SAA}
1529: T. A. Shutenko, I. L. Aleiner and B. L. Altshuler, 
1530: Phys. Rev. {\bf B61}, 10366 (2000).
1531: 
1532: \bibitem{pMB}
1533: M.~Moskalets and M.~B{\"u}ttiker, 
1534: Phys. Rev. B {\bf 68}, 161311(R) (2003).
1535: 
1536: 
1537: 
1538: 
1539: \bibitem{berry1}
1540: M.V. Berry, Proc. R. Soc. Lond. A {\bf 392}, 45 (1984).
1541: 
1542: \bibitem{thouless}
1543: D. J. Thouless, 
1544: Phys. Rev. B {\bf 27}, 6083 (1983).
1545: 
1546: \bibitem{AvronNet}
1547: J.E. Avron, A. Raveh and B. Zur,  
1548: Rev. Mod. Phys. {\bf 60}, 873 (1988).
1549: 
1550: \bibitem{berry2}
1551: M.V. Berry and J.M. Robbins, Proc. R. Soc. Lond. A {\bf 442}, 659 (1993).
1552: 
1553: \bibitem{pmm} 
1554: I.~Sela and D.~Cohen, in preparation. 
1555: 
1556: \bibitem{pmo}
1557: D. Cohen, Phys. Rev. B {\bf 68}, 201303(R) (2003).
1558: 
1559: 
1560:  
1561: \end{thebibliography}
1562: 
1563: 
1564: 
1565: 
1566: 
1567: \newpage
1568: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1569: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1570: %%% Figures
1571: 
1572: 
1573: 
1574: 
1575: 
1576: 
1577: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1578: \mpg{
1579: 
1580: % 2 delta model
1581: 
1582: \Cn{\putgraph[width=0.9\hsize]{ringnet}} 
1583: 
1584: {\footnotesize {\bf FIG. 1.} 
1585: Illustration of the model system. 
1586: The two barrier pumping device 
1587: is used in two different configurations. 
1588: Left panel: open lead geometry;  
1589: Right panel: closed ring geometry.
1590: The barriers are located 
1591: at the nodes $x_1$ and $x_2$ 
1592: while the current is measured 
1593: through the section at $x_0$. }
1594: 
1595: }
1596: 
1597: \ \\ \ \\ 
1598: 
1599: 
1600: 
1601: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1602: \mpg{
1603: 
1604: % pumping cycle
1605: 
1606: \Cn{
1607: \putgraph[height=0.3\hsize]{pme_2delta}
1608: \putgraph[height=0.3\hsize]{pmp_levels} 
1609: }
1610: 
1611: \Cn{
1612: \putgraph[width=0.92\hsize]{DiracChains}
1613: }
1614: 
1615: {\footnotesize {\bf FIG 2.}
1616: (a) Upper left: The energy levels of a ring 
1617: with two barriers, at the beginning of the pumping cycle. 
1618: It is assumed that the three lower levels are occupied. 
1619: (b) Upper right: The adiabatic levels as a function 
1620: of time during the pumping cycle. 
1621: (c) Lower Left: The $(X_1,X_2)$ locations of 
1622: the Dirac chains of the $3$ occupied levels. 
1623: Filled (hollow) circles imply that there 
1624: is (no) monopole in the pumping plane. 
1625: Note that for sake of illustration overlapping 
1626: chains are displaced from each other.
1627: The pumping cycle encircles $2+1$ Dirac chains 
1628: that are associated with the 3rd and 2nd levels respectively.   
1629: (d) Lower right: The $2$ Dirac chains 
1630: that are associated with the 3rd level.}
1631: 
1632: }
1633: 
1634: 
1635: \ \\ 
1636: 
1637: 
1638: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1639: \mpg{
1640: 
1641: % non stationary evolution
1642: 
1643: \putgraph[width=0.45\hsize]{adiabat} 
1644: \mpg[0.5\hsize]{
1645: {\footnotesize {\bf FIG. 3.} }
1646: Cartoon of the adiabatic evolution 
1647: within the framework of the two level approximation. 
1648: The system is prepared in level $n$, 
1649: and the nearby empty level is ${m=n+1}$. 
1650: The control parameter $X(t)$ is being changed 
1651: slowly, and therefore the system ``oscillates" 
1652: around the first order adiabatic solution. 
1653: The energy of the latter is illustrated 
1654: by a dashed line. Note that the identity 
1655: of the adiabatic state changes gradually, 
1656: and can be regarded as constant only  
1657: on scales $\delta X \ll \delta X_c$.}  
1658:    
1659: }
1660: 
1661: 
1662: \ \\ \ \\ 
1663: 
1664: 
1665: 
1666: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1667: \mpg{
1668: 
1669: % Q plot
1670: 
1671: \Cn{
1672: \mpg[0.4cm]{(a) \\ \vspace*{4cm} }
1673: \putgraph[width=0.5\hsize]{Qpumped}
1674: }
1675: 
1676: \Cn{
1677: \mpg[0.4cm]{(b) \\ \vspace*{2cm} }
1678: \putgraph[height=0.25\hsize]{QpumpedZoom}
1679: \hspace*{1cm}
1680: \mpg[0.4cm]{(c) \\ \vspace*{2cm} }
1681: \putgraph[height=0.25\hsize]{X1X2PumpingPath}
1682: }
1683: 
1684: {\footnotesize {\bf FIG. 4.} }
1685: Panels (a) and (b): 
1686: Numerical calculation of the pumped charge $Q$.
1687: The model parameters are ${L_{\tbox{D}}=\mass=e=1}$   
1688: and ${L_{\tbox{W}} = 3000.43}$. The current $I$ 
1689: is measured at the middle of the dot. 
1690: The numerical integration is carried out along 
1691: the segments which are indicated in panel (c),   
1692: and the results are multiplied by~2 so as to include 
1693: the equal contribution that comes from the second 
1694: half of the cycle.  There are two sets of data points.
1695: One set (filled circles) is for pumping cycles that encircle 
1696: an in-plane degeneracy point ($n_r=2993$).  
1697: A second set (hollow circles) is for pumping cycles 
1698: that encircle an off-plane degeneracy point ($n_r=2992$).
1699: The location of the avoided crossing for each 
1700: data set is indicated by the solid lines in panel (c).
1701: The near and that far field approximations 
1702: that we derive for~$Q$ are indicated by the 
1703: the solid lines in panel~(a). The zoom in panel~(b) 
1704: reveals that~$Q$ in the far field is in fact slightly 
1705: less then~$1$, which is explained in section~13.
1706: }
1707: 
1708: 
1709: 
1710: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1711: \mpg{
1712: 
1713: % X space regions
1714: 
1715: \Cn{
1716: \mpg[0.4cm]{(a) \\ \vspace*{5cm} }
1717: \putgraph[height=0.6\hsize]{X1_X2Plane}
1718: }
1719: 
1720: \Cn{
1721: \mpg[0.4cm]{(b) \\ \vspace*{3cm} }
1722: \putgraph[height=0.35\hsize]{X1_X2PlaneZoom}
1723: \hspace*{1cm}
1724: \mpg[0.4cm]{(c) \\ \vspace*{3cm} }
1725: \putgraph[height=0.33\hsize]{kDisconnectedRing}
1726: }
1727: 
1728: \Cn{
1729: \mpg[0.4cm]{(d) \\ \vspace*{3cm} }
1730: \putgraph[width=0.4\hsize]{kX1EqualX2}
1731: \hspace*{1cm}
1732: \mpg[0.4cm]{(e) \\ \vspace*{3cm} }
1733: \putgraph[width=0.4\hsize]{kX1EqualX2FluxPi}
1734: }
1735: 
1736: {\footnotesize {\bf FIG. 5.} 
1737: Regions in $\bm{X}$ space. The model parameters 
1738: are the same as in the previous figure. 
1739: Panel (a) displays the region of interest 
1740: as defined in Eq.(\ref{e34}). It is bounded 
1741: by the left and by the bottom solid lines 
1742: which are defined by $g(X) \sim 1/b$.
1743: In-plane and off-plane degeneracy points 
1744: are indicated by filled and hollow circles respectively.  
1745: We indicate by arrow one in-plane degeneracy 
1746: point ($n_r = 2993$). A zoom of its near 
1747: field is displayed in panel (b). The ellipse 
1748: in panel (b) indicates a level splitting 
1749: that equals $\Delta/10$. The dashed lines 
1750: in panels (a) and (b) indicate far field and near 
1751: field pumping cycles respectively.   
1752: In panels (c)-(e) we shows the energy levels ($k_n$) 
1753: along three paths in $\vec{X}$ space, 
1754: which are $(X_1{=}X,X_2{=}\infty, \Phi{=}0)$,  
1755: and $(X_1{=}X,X_2{=}X, \Phi{=}0)$,  
1756: and $(X_1{=}X,X_2{=}X, \Phi{=}\pi)$ respectively.
1757: The degeneracies ${n_r = 2992...2999}$ 
1758: are circled. The arrow indicates the 
1759: representative degeneracy point ${n_r = 2993}$. 
1760: In panels (d) the odd states are indicated 
1761: by dashed lines so as to distinguish 
1762: them from the even states. }
1763: 
1764: }
1765: 
1766: 
1767: 
1768: 
1769: 
1770: 
1771: 
1772: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1773: \mpg{
1774: 
1775: % two levels approximation
1776: % X space contors
1777: 
1778: \mpg[0.4cm]{(a) \\ \vspace*{4cm} }
1779: \putgraph[width=0.6\hsize]{X1X2PlaneLevelSplitting0Pi} 
1780: 
1781: \mpg[0.4cm]{(b) \\ \vspace*{4cm} }
1782: \putgraph[width=0.6\hsize]{X1X2PlaneLevelSplittingMixing0}
1783: 
1784: \mpg[0.4cm]{(c) \\ \vspace*{4cm} }
1785: \putgraph[width=0.6\hsize]{X1X2PlaneLevelSplittingMixingPi}
1786: 
1787: 
1788: {\footnotesize {\bf FIG. 6.} 
1789: The energy level splitting and the mixing 
1790: parameter $\Theta$ for two pairs of levels.
1791: The model parameters are ${L_{\tbox{D}}=\mass=e=1}$   
1792: and ${L_{\tbox{W}} = 160.43}$. 
1793: In panel (a) we show the contour lines 
1794: for the energy level splitting of 
1795: the first (even) dot level with an odd wire level ($n = 158$), 
1796: and for the energy level splitting of 
1797: the first (even) dot level with 
1798: an even wire level ($n = 157$).   
1799: The two cases are displayed again 
1800: in panels (b) and (c) respectively 
1801: where we plot both level splitting contours 
1802: (solid lines) and $\Theta$ contours (dashed lines).
1803: In the ``even-odd crossing" case 
1804: we have an in-plane degeneracy, which is 
1805: indicated by a filled circle,  
1806: while the inner most contour line 
1807: is for $\Delta/5$ splitting. 
1808: Note that within the white regions 
1809: the mixing is maximal ($\Theta\sim\pi/2$).   
1810: In the ``even-even avoided crossing" case 
1811: the projection of the off-plane degeneracy point  
1812: is indicated by a hollow circle.} 
1813: 
1814: % The mixing contour lines
1815: % correspond to levels mixing
1816: % such that $\frac{1}{\sqrt{b}}\tan(\Theta^{(\pm)}/2)$
1817: % is equal $0.15,\  0.5,\ 1,\ 5,\ 10,\ 20,\ 30,\ 40$.
1818: %
1819: % In this plot $\frac{1}{\sqrt{b}}\tan(\Theta^{(\pm)}/2)$
1820: % is equal $1/4,\  1/2,\ 1,\ 3/2,\ 2$.
1821: 
1822: }
1823: 
1824: 
1825: 
1826: 
1827: 
1828: 
1829: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1830: \mpg{
1831: 
1832: % near field analysis
1833: % k, q, Theta, phi
1834: 
1835: \putgraph[width=0.3\hsize]{KNearFieldApproximation}
1836: \putgraph[width=0.3\hsize]{ThetaNearFieldApproximation}
1837: \putgraph[width=0.3\hsize]{phiDotNearFieldApproximation}
1838: 
1839: 
1840: {\footnotesize {\bf FIG. 7.} 
1841: Tests of the perturbation theory based 
1842: approximations (dashed lines) against 
1843: the numerics (solid lines).
1844: The model parameters are ${L_{\tbox{D}}=\mass=e=1}$   
1845: and ${L_{\tbox{W}} = 3000.43}$, and we focus 
1846: on the degeneracy point $n_r = 2993$.
1847: For these parameters ${X^{(r)} \approx 465}$ 
1848: and ${g^{(r)} \approx 4.5 \times 10^{-5}}$.
1849: All the plots refer to the path $(X_2 - X_1) = 5$.
1850: In the left panel the dashed lines are derived 
1851: from Eq.(\ref{e45}). In the middle panel the dashed 
1852: lines are based on Eqs.(\ref{e48}-\ref{e49}) 
1853: with $\theta$ from Eq.(\ref{e44}).
1854: In the right panel the dashed lines are deduced 
1855: from Eq.(\ref{e38})}
1856: 
1857: 
1858: }
1859: 
1860: 
1861: \ \\ \ \\ 
1862: 
1863: 
1864: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1865: \mpg{
1866: 
1867: % phi vs phi branches
1868: 
1869: \putgraph[width=0.49\hsize]{phiDotphiWireBranches}
1870: 
1871: 
1872: {\footnotesize {\bf FIG. 8.} }
1873: The wire phase $\varphi_{\tbox{W}}/\pi$ versus 
1874: the dot phase $\varphi_{\tbox{D}}/\pi$ 
1875: at a node with delta barrier $g(X) = 0.225$.  
1876: The two branches are implied by the 
1877: matching condition Eq.(\ref{e27}). The ratio 
1878: $|{\sin(\varphi_{\tbox{D}})}/{\sin(\varphi_{\tbox{D}})}|$ 
1879: attains its extremal values (Eq.(\ref{e35})) 
1880: at the points which are indicated by circles.}
1881: 
1882: 
1883: 
1884: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1885: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1886: \end{document}
1887: 
1888: 
1889: 
1890: 
1891: 
1892: 
1893: 
1894: 
1895: 
1896: 
1897: 
1898: 
1899: 
1900: 
1901: 
1902: 
1903: 
1904: 
1905: 
1906: