1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %%% Cohen (March 2003)
3: %%% Figures:
4: %%% Fig.1: pmp_fig_ae.eps
5: %%% Fig.2: pmp_delta.eps; pmp_levels.eps
6: %%% Fig.3: pmp_bfield_en.eps
7: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8:
9: \documentclass[aps,pre,twocolumn,floats]{revtex4}
10: \usepackage{epsfig}
11: \usepackage{bm}
12:
13: %\documentclass[bm]{iopart}
14: %\usepackage{epsfig}
15:
16: \begin{document}
17:
18: \newcommand{\hide}[1]{}
19: \newcommand{\tbox}[1]{\mbox{\tiny #1}}
20: \newcommand{\half}{\mbox{\small $\frac{1}{2}$}}
21: \newcommand{\const}{\mbox{const}}
22: \newcommand{\ointt}{\int\!\!\!\!\int\!\!\!\!\!\circ\ }
23: \newcommand{\intt}{\int\!\!\!\!\int }
24: \newcommand{\ar}{\mathsf r}
25: \newcommand{\im}{\mbox{Im}}
26: \newcommand{\re}{\mbox{Re}}
27: \newcommand{\mbf}[1]{{\mathbf #1}}
28: \newcommand{\sinc}{\mbox{sinc}}
29: \newcommand{\trc}{\mbox{trace}}
30: \newcommand{\eexp}{\mbox{e}^}
31: \newcommand{\bra}{\left\langle}
32: \newcommand{\ket}{\right\rangle}
33:
34: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
35:
36: \title{Classical and quantum pumping in closed systems}
37:
38: \author{Doron Cohen}
39:
40: %\address{
41: \affiliation{
42: Department of Physics, Ben-Gurion University, Beer-Sheva 84105, Israel
43: }
44:
45: \date{Aug 2002, long published version \cite{pmp}, follow up \cite{pmo}}
46:
47: %\pacs{03.65.-w}{Quantum mechanics}
48: %\pacs{03.65.Vf}{Geometric phases etc}
49: %\pacs{73.23.-b}{Mesoscopics/transport}
50: %\pacs{05.45.Mt}{Quantum chaos}
51:
52:
53: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
54:
55: \begin{abstract}
56: Pumping of charge ($Q$) in a closed ring geometry is not
57: quantized even in the strict adiabatic limit. The deviation form
58: exact quantization can be related to the Thouless conductance.
59: We use the Kubo formalism as a starting point for the calculation
60: of both the dissipative and the adiabatic contributions to $Q$.
61: As an application we bring examples for classical dissipative
62: pumping, classical adiabatic pumping, and in particular we
63: make an explicit calculation for quantum pumping in case of the
64: simplest pumping device, which is a 3~site lattice model.
65: We make a connection with the popular $S$ matrix formalism
66: which has been used to calculate pumping in open systems.
67: \end{abstract}
68:
69: \maketitle
70: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
71:
72: Pumping of charge in mesoscopic \cite{marcus_rev}
73: and molecular size devices is regarded as a major
74: issue in the realization of future "quantum circuits"
75: or "quantum gates", possibly for the purpose of
76: "quantum computing". Of particular interest is
77: the possibility to realize a pumping cycle that transfers
78: {\em exactly} one unit of charge per cycle
79: \cite{thouless,avronI,aleiner,barriers}.
80: In open systems this "quantization" holds only
81: approximately. But it has been argued \cite{aleiner}
82: that the deviation from quantization is due to
83: "dissipative" effect, and that exact quantization
84: would hold in the strict adiabatic limit, if the
85: system were {\em closed}.
86: %
87: In this Letter we would like to show that the
88: correct picture is quite different.
89: In particular we would like to make
90: a proper distinction between "dissipative" and
91: "adiabatic" contributions to the pumping, and
92: to calculate the deviation from exact quantization
93: in the latter case. As a starting point we adopt
94: the traditional Kubo formula \cite{landau},
95: but we also point out the relation
96: to the "adiabatic" \cite{berry,robbins}
97: and to the "$S$-matrix" \cite{BPT} formulations.
98: %
99: The present formulation of the pumping problem
100: has few advantages: It is not restricted to the
101: adiabatic regime; It give a "level by level"
102: understanding of the pumping process;
103: It allows the consideration of any type of
104: occupation (not necessarily Fermi occupation);
105: It allows future incorporation of external
106: environmental influences such as that of noise;
107: It regards the "voltage" over the pump
108: as "electro motive force", rather than adopting
109: the conceptually complicated view \cite{ora}
110: of having a "chemical potential difference".
111: %
112: Finally, on the practical level, we give
113: a solution for the pumping
114: in a 3~site lattice model.
115: This is definitely the simplest pump circuit
116: possible, and we believe that it can be realized
117: as a molecular size device.
118: It also can be regarded as an approximation
119: for the closed geometry version of the
120: two delta potential pump \cite{barriers}.
121:
122:
123: The structure of this Letter is as follows:
124: We show how to get from the Kubo formalism
125: an expression for the pumped charge $Q$,
126: and explain the distinction between
127: "dissipative" and the "adiabatic" contributions.
128: Then we give illuminating examples
129: for classical dissipative pumping
130: and for classical adiabatic pumping.
131: Next we discuss the case of
132: quantum pumping, where the cycle is around
133: a chain of degeneracies. We show that
134: this can be understood as a special
135: case of "adiabatic transfer" scheme.
136: In order to get a quantitative estimate
137: for the pumped charge we consider
138: a 3~site lattice model. We get expressions
139: for $Q$, and express them in terms of the Thouless
140: conductance. We conclude by a short discussion
141: of the relation between the Kubo formalism,
142: the adiabatic formalism, and the $S$-matrix
143: formalism.
144:
145:
146:
147: %%%%%%%%%%
148: %%%%%%%%%%
149: %%%%%%%%%%
150: %%%%%%%%%%
151:
152:
153:
154:
155: Consider a system that has a ring geometry (Fig.1a).
156: The Hamiltonian is ${\cal H}(x_1(t),x_2(t),x_3(t))$,
157: where $x_1$ and $x_2$ are parameters that control
158: the shape of the ring, or the height of some barriers,
159: while $x_3{=}\Phi{=}\hbar \phi$ is the magnetic flux.
160: We use units such that the elementary charge is unity.
161: The "generalized forces" are conventionally defined as
162: $F^k \equiv -{\partial {\cal H}}/{\partial x_k}$.
163: In particular $\langle F^{3} \rangle$ is the current $I$ through
164: the ring (see remark \cite{rmrk}).
165: Consider for a moment the time independent Hamiltonian
166: ${\cal H}(x)$, with $x=\mbox{const}$,
167: and assume that the system is prepared in a
168: {\em stationary} state (either pure or mixed).
169: The expectation value $\langle F^k \rangle$
170: of a generalized force is known as the
171: "conservative force" or (in case of $k{=}3$)
172: as the "persistent current". The "fluctuations" of the
173: generalized forces are conventionally characterized
174: by the real functions:
175: %
176: \begin{eqnarray} \label{e1}
177: C^{ij}(\tau) &=& \langle \half (F^i(\tau)F^j(0)+F^j(0)F^i(\tau)) \rangle \\
178: K^{ij}(\tau) &=& \frac{i}{\hbar} \langle [F^i(\tau),F^j(0)]\rangle
179: \end{eqnarray}
180: %
181: Note that both functions have a well defined classical limit.
182: Their Fourier transform will be denoted by $\tilde{C}^{ij}(\omega)$
183: and $\tilde{K}^{ij}(\omega)$ respectively.
184:
185:
186:
187:
188:
189: Our interest in the following is in a {\em driving~cycle},
190: where $x=x(t)$ forms a loop in the 3~dimensional parameter space.
191: %
192: In linear response theory \cite{landau}
193: the non-conservative contribution to $\langle F^{k} \rangle$
194: is related to $x(t)$ by a causal
195: response kernel \mbox{$\alpha^{ij}(t-t')$}.
196: The Kubo expression for this response kernel is
197: $\alpha^{ij}(\tau) = \Theta(\tau) \ K^{ij}(\tau)$.
198: Its Fourier transform is the generalized
199: susceptibility $\chi^{ij}(\omega)$.
200: From here we can derive the expression
201: $\langle F^k \rangle =
202: -\sum_{j} \mbf{G}^{kj} \ \dot{x}_j$,
203: where $\mbf{G}^{kj}$ is the generalized
204: conductance matrix:
205: %
206: \begin{eqnarray} \label{e3}
207: \mbf{G}^{ij} \ = \ \lim_{\omega\rightarrow 0}
208: \frac{\im[\chi^{ij}(\omega)]}{\omega} \ = \
209: \int_0^{\infty} K^{ij}(\tau)\tau d\tau
210: \end{eqnarray}
211: %
212: Following Berry and Robbins \cite{robbins} we split
213: the conductance matrix
214: into symmetric and anti-symmetric parts.
215: Namely $\mbf{G}^{ij} = \bm{\eta}^{ij} + \mbf{B}^{ij}$.
216: The antisymmetric part $\mbf{B}$ can be regarded as a vector
217: $\vec{\mbf{B}}=(\mbf{B}^{23},\mbf{B}^{31},\mbf{B}^{12})$,
218: and the expression for the current can be written
219: in an abstract way as
220: $\langle F \rangle = -\bm{\eta} \cdot \dot{x} - \mbf{B}\wedge \dot{x}$.
221:
222:
223:
224:
225:
226:
227: The rate of dissipation, which is defined as the rate
228: in which energy is absorbed into the system, is given by
229: $\dot{{\cal W}} = -\langle F \rangle \cdot \dot{x} =
230: \sum_{kj} \bm{\eta}^{ij} \dot{x}_i\dot{x}_j$.
231: Only the symmetric part of $\mbf{G}^{ij}$
232: is responsible for dissipation of energy.
233: %
234: %
235: The adiabatic regime is defined by the condition
236: $|\dot{x}| \ll {\Delta^2}/{\hbar\sigma}$,
237: where $\Delta$ is the typical level spacing,
238: and $\sigma$ is the root mean square value of
239: the matrix element $(\partial{\cal H}/\partial x)_{nm}$
240: between neighboring levels.
241: %
242: %
243: In the adiabatic regime $\bm{\eta}^{ij}$ vanishes because
244: of the discreteness of the energy spectrum \cite{robbins}.
245: But outside of the adiabatic regime the levels
246: acquires an effective width
247: \mbox{$\Gamma/\Delta=(({\hbar\sigma}/{\Delta^2})V)^{2/3}>1$}
248: and therefore the smoothed version of $\tilde{K}^{ij}(\omega)$
249: should be considered. Consequently one can obtain
250: the fluctuation-dissipation (FD) relation:
251: $\bm{\eta}^{ij} \sim \tilde{C}^{ij}(\omega=0)$.
252: %
253: The formulation of the exact FD relation
254: depends on the assumptions regarding the
255: occupation $f(E_n)$ of the energy levels. See \cite{frc,wlf}.
256: Commonly one assumes a zero temperature Fermi
257: occupation, but this is not essential for the following analysis.
258: %
259: In order to derive the above expression for $\Gamma$
260: we have used the result of \cite{frc} (Sec.17) for the "core width"
261: at the breaktime $t=t_{\tbox{prt}}$ of perturbation theory.
262: Note that in the semiclassical limit (small $\hbar$) the
263: adiabaticity condition always breaks down.
264:
265:
266:
267: The antisymmetric part $\mbf{B}$ of $\mbf{G}^{ij}$
268: does not have to vanish in the adiabatic limit. It can be
269: obtained from the adiabatic equation by looking
270: for a first-order stationary-like solution
271: \cite{thouless,avronI,robbins}, but we prefer to regard it
272: as a term in the (full) Kubo expression Eq.(\ref{e3}).
273: In \cite{berry,robbins} it has been demonstrated that it can be written as
274: %
275: \begin{eqnarray} \nonumber
276: \mbf{B}^{ij} &=& -2\hbar\sum_n f(E_n) \
277: \im\left.\left\langle\frac{\partial}{\partial x_i} n(x) \right|
278: \frac{\partial}{\partial x_j} n(x) \right\rangle
279: \\ \label{e4}
280: &=&
281: 2\hbar \sum_{m \ne n} f(E_n)
282: \frac{\im\left[
283: \left(\frac{\partial {\cal H}}{\partial x_i}\right)_{nm}
284: \left(\frac{\partial {\cal H}}{\partial x_j}\right)_{mn}\right]}
285: {(E_m-E_n)^2}
286: \end{eqnarray}
287: %
288: Note that the "vertical" component of $\vec{\mbf{B}}$ vanishes
289: in the "horizontal" $x_3{=}0$ plane due to time reversal symmetry.
290:
291:
292:
293:
294: Disregarding a possible persistent current contribution
295: (that does not exist in the case of a planar $\Phi{=}0$~cycle),
296: the expression for the pumped charge is:
297: %
298: \begin{eqnarray} \label{e6}
299: Q \ = \ \oint I dt \ = \
300: -\left[ \oint \bm{\eta} \cdot dx + \oint \mbf{B} \wedge dx \right]_{k=3}
301: \end{eqnarray}
302: %
303: If we neglect the first term, which is associated
304: with the dissipation effect, and average the second
305: ("adiabatic") term over the flux, then we get
306: %
307: \begin{eqnarray} \label{e7}
308: \overline{Q|_{\tbox{adiabatic}}} \ = \
309: -\frac{1}{2\pi\hbar}\intt \mbf{B} \cdot \vec{dx} \wedge \vec{dx}
310: \ = \ \mbox{integer}
311: \end{eqnarray}
312: %
313: The integration should be taken over
314: a cylinder of vertical height $2\pi\hbar$,
315: and whose basis is determined by the projection of
316: the pumping cycle onto the $(x_1,x_2)$ plane.
317: %
318: The last equality is argued as follows:
319: The flux \mbox{$(1/\hbar)\intt \mbf{B} \cdot {dx} \wedge {dx}$}
320: through a surface that is enclosed by a cycle is the Berry phase \cite{berry}.
321: The result should be independent of the surface.
322: Therefore the flux through a {\em closed} surface
323: should equal $2\pi\times$integer.
324: Integrating over a cylinder, as in Eq.(\ref{e7}),
325: is effectively like integrating over a closed surface
326: (because of the $2\pi$~periodicity in the vertical
327: direction). This means that the flux averaged $Q$
328: of Eq.(\ref{e7}) has to be an integer.
329:
330:
331: %%%%%%%
332: %%%%%%%
333:
334:
335:
336:
337: Before we discuss the quantum mechanical pumping,
338: it is instructive to bring two simple examples for
339: {\em classical} pumping. In the following we consider
340: one particle ($\mbf{r}$)
341: in a two dimensional ring as in Fig.1a.
342:
343:
344:
345: The first example is for classical {\em dissipative} pumping.
346: The conductance $G=\mbf{G}^{33}$
347: can be calculated for this system \cite{wlf}
348: leading to a mesoscopic variation of the Drude formula.
349: The current is $I=-G\times\dot{\Phi}$, where $-\dot{\Phi}$
350: is the electro-motive-force.
351: %
352: Consider now the following pumping cycle:
353: Change the flux from $\Phi_1$ to $\Phi_2$,
354: hence pumping charge $Q = -G(1)\times (\Phi_2-\Phi_1)$.
355: Change the conductance from $G(1)$ to $G(2)$
356: by modifying the shape of the ring.
357: Change the flux from $\Phi_2$ back to $\Phi_1$,
358: hence pumping charge $Q(2) = -G(2) \times (\Phi_1-\Phi_2)$.
359: Consequently the net pumping
360: is $Q = (G(2)-G(1))\times(\Phi_2-\Phi_1)$.
361:
362:
363:
364:
365: The second example is for classical {\em adiabatic} pumping.
366: The idea is to trap the particle
367: inside the ring by a potential well
368: $U_{\tbox{trap}}(\mbf{r}_1{-}x_1(t),\mbf{r}_2{-}x_2(t))$.
369: Then make a translation of the trap along
370: a circle of radius $R$, namely
371: $x(t) = (R\cos(\Omega t),R\sin(\Omega t),\Phi{=}\const)$.
372: It is a-priori clear that in this
373: example the pumped charge per cycle is $Q=1$, irrespective
374: of $\Phi$. Therefore the $\vec{\mbf{B}}$ field must be
375: %
376: \begin{eqnarray} \label{e8}
377: \vec{\mbf{B}} = -\frac{(x_1,x_2,0)}{2\pi (x_1^2+x_2^2)}
378: \end{eqnarray}
379: %
380: This can be verified by calculation via Eq.(\ref{e4}).
381: The singularity along the $x_3$ axis
382: is not of quantum mechanical origin:
383: It is not due to degeneracies, but rather
384: due to the diverging current operator
385: ($\partial {\cal H}/\partial x_3\propto 1/\sqrt{x_1^2+x_2^2}$).
386:
387:
388:
389:
390: %%%%%%%%%%
391: %%%%%%%%%%
392:
393:
394:
395:
396: We turn now to the quantum mechanical case.
397: Consider an adiabatic cycle that
398: involves a particular energy level $n$.
399: This level is assumed to have a degeneracy
400: point at $(x_1^{(0)},x_2^{(0)},\Phi^{(0)})$.
401: It follows that in fact there is
402: a vertical "chain" of degeneracy points
403: that are located at
404: \mbox{$(x_1^{(0)},x_2^{(0)},\Phi^{(0)}+2\pi\hbar\times\mbox{\small integer})$}.
405: These degeneracy points are important for the geometrical
406: understanding of the $\mbf{B}$ field, as implied by Eq.(\ref{e4}).
407: Every degeneracy point is like a monopole charge.
408: The total flux that emerges from each monopole
409: must be $2\pi\hbar\times$integer for a reason that was
410: explained after Eq.(\ref{e7}).
411: Thus the monopoles are quantized in units of~$\hbar/2$.
412:
413:
414: The $\mbf{B}$ field which is created (so to say)
415: by a vertical chain of monopoles may have a different ``near field"
416: and ``far field" behavior, which we discuss below.
417: (Later we further explain that "near field" means regions in $x$ space,
418: in the vicinity of degeneracy points, where $g_T \gg 1$,
419: while "far field" means regions where $g_T \ll 1$).
420: The far field regions exist if the chains are well isolated.
421: The far field region of a given chain is obtained by regarding the
422: chain as a smooth line.
423: This leads {\em qualitatively} to the same field as in Eq.(\ref{e8}).
424: Consequently, for a "large radius" pumping cycle
425: in the $\Phi=0$ plane, we get $|Q|\approx1$.
426: In the following we are interested in
427: the deviation from "exact" quantization:
428: If $\phi^{(0)}=0$ we expect to have
429: $|Q| \ge 1$, while if $\phi^{(0)}=\pi$ we expect $|Q|\le 1$.
430: Only for the $\phi$ averaged $Q$ of Eq.(\ref{e7})
431: we get {\em exact quantization}.
432:
433:
434:
435: The deviation from $|Q|\approx 1$ is extremely large
436: if we consider a tight pumping cycle around
437: a $\phi^{(0)}=0$ degeneracy.
438: After linear transformation of the shape parameters,
439: the energy splitting $\Delta=E_n-E_m$ of the energy level~$n$
440: from its neighboring (nearly degenerated) level~$m$
441: can be written as
442: %
443: $\Delta=
444: ((x_1{-}x_1^{(0)})^2+(x_2{-}x_2^{(0)})^2+
445: c^2(\phi{-}\phi^{(0)})^2)^{1/2}$
446: %
447: where $c$ is a constant. The monopole field is accordingly
448: %
449: \begin{eqnarray} \label{e9}
450: \vec{\mbf{B}} = \pm\frac{c}{2} \
451: \frac{( x_1{-}x_1^{(0)}, x_2{-}x_2^{(0)}, x_3{-}x_3^{(0)}) }
452: {((x_1{-}x_1^{(0)})^2+(x_2{-}x_2^{(0)})^2+
453: (\frac{c}{\hbar})^2(x_3{-}x_3^{(0)})^2)^{3/2}}
454: \end{eqnarray}
455: %
456: where the prefactor is determined by the requirement
457: of having a single ($\hbar/2$) monopole charge.
458: Assuming a pumping cycle of radius $R$ in the $\Phi=0$ plane
459: we get from the second term of Eq.(\ref{e6}) that
460: the pumped charge is $Q = \mp \pi \sqrt{g_T}$,
461: where $g_T=({\partial^2\Delta}/{\partial\phi^2})/\Delta={c^2}/{R^2}$
462: is a practical definition for the Thouless conductance
463: in this context. It is used here simply as a measure
464: for the sensitivity of an energy level to the magnetic flux $\Phi$.
465:
466:
467: What we want to do in the following is to "interpolate"
468: between the "near field" result, which is $Q={\cal O}(\sqrt{g_T})$,
469: and the "far field" result, which is $Q={\cal O}(1)$.
470: For this purpose it is convenient to consider
471: a particular model that can be solved exactly.
472: %
473: %
474: We consider a ring with two barriers.
475: The model is illustrated in Fig.2.
476: A version of this model, where the two barriers
477: are modeled as "delta functions", has been analyzed
478: in \cite{barriers} in case of {\em open} geometry.
479: Below we are going to analyze a different version
480: of the two barrier model, that allows an exact
481: solution for {\em closed} geometry.
482:
483:
484:
485: We can classify the eigenstates of the {\em closed} ring
486: into two categories: wire states, and dot states (Fig.2a).
487: The latter are those states that are localized in the
488: "dot region" in the limit of infinitely
489: high barriers. In case of zero temperature Fermi
490: occupation we define $E_F$ as the energy of the last
491: occupied wire level in the limit of infinitely high barriers.
492: %
493: %
494: The two "shape" parameters are the the bias $x_1$,
495: and the dot potential $x_2$.
496: The bias determines whether the dot tends
497: to exchange particles via the {\em left}
498: or via the {\em right} barrier.
499: The dot potential is loosely defined
500: as the energy of the dot level (Fig.2a).
501: A model specific definition of these
502: parameters in the context of the
503: 3-site lattice Hamiltonian will be given later.
504:
505:
506:
507:
508: {\em The pumping cycle is assumed to be in the
509: $\Phi=0$ plane, so there is no issue
510: of "conservative" persistent current contribution}.
511: We start with a positive bias ($x_1>0$)
512: and lower the dot potential from a large $x_2>E_F$
513: value to a small $x_2<E_F$ value.
514: As a result, one electron is
515: transfered via the {\em left} barrier into the
516: dot region. Then we invert the bias
517: ($x_1<0$) and raise back $x_2$. As a result
518: the electron is transfered back into
519: the wire via the {\em right} barrier.
520: %
521: %
522: %
523: A closer look at the above scenario (Fig.2b)
524: reveals the following:
525: As we lower the dot potential across a wire
526: level, an electron is adiabatically transfered
527: once from left to right and then from right
528: to left. As long as the bias is
529: positive ($x_1>0$) the net charge being pumped
530: is very small ($|Q| \ll 1$). Only the lowest wire
531: level that participate in the pumping cycle
532: carries $Q={\cal O}(1)$ net charge:
533: It takes an electron from the left side,
534: and after the bias reversal it emits it
535: into the right side.
536: %
537: %
538: %
539: Thus the pumping process in this model can be regarded
540: as a particular example \cite{avronI} of an {\em adiabatic transfer scheme}:
541: The electrons are adiabatically transfered from
542: state to state, one by one, as in "musical chair game".
543:
544:
545:
546: For a single occupied level the net $Q$ is the
547: sum of charge transfer events that take place in
548: few avoided crossings. For many particle occupation
549: the total $Q$ is the sum over the net $Q$s
550: which are carried by individual levels.
551: For a dense zero temperature Fermi occupation
552: the summation over all the net $Q$s is a telescopic sum,
553: leaving non-canceling contributions only from the
554: first and the last adiabatic crossings. The latter
555: involve the last occupied level at the Fermi energy.
556:
557:
558:
559: In order to get a quantitative estimate
560: for the $Q$ in a given avoided crossing,
561: we consider the simplest version
562: of the "two barrier model" that still
563: contains all the essential ingredients:
564: This is a three site lattice system.
565: %
566: The middle site supports a single "dot state",
567: while the two other sites support
568: two "wire states". The Hamiltonian is
569: %
570: \begin{eqnarray} \label{e10}
571: {\cal H} \mapsto \left(
572: \matrix{
573: 0 & c_1 & \eexp{i\phi} \cr
574: c_1 & u & c_2 \cr
575: \eexp{-i\phi} & c_2 & 0} \right)
576: \end{eqnarray}
577: %
578: The three parameters are
579: the bias $x_1=c_1-c_2$,
580: the dot potential $x_2=u$,
581: and the flux $x_3 = \Phi = \hbar\phi$.
582: For presentation purpose we assume
583: that \mbox{$0 < c_1,c_2 \ll 1$}.
584: %
585: %
586: %
587: The eigenstates are $E_n$.
588: Disregarding the coupling between
589: the "wires" and the "dot" we have
590: two wire states with $E = \pm 1$,
591: and a dot state with $E = u$.
592: Taking into account the wire-dot
593: coupling we find that
594: there are two vertical chains
595: of degeneracies. The $u\approx-1$ chain is
596: \mbox{$(0,-1{+}c_1^2,2\pi\hbar\times\mbox{\small integer})$}
597: and the $u\approx1$ chain is
598: \mbox{$(0,+1{+}c_1^2,\pi+2\pi\hbar\times\mbox{\small integer})$}.
599:
600:
601:
602:
603: The eigenvalues $E_n$ are the solutions
604: of a cubic equation. Rather than
605: writing the (lengthy) analytical expressions
606: for them we give a numerical example for
607: their dependence on $u$ in the inset of Fig.3.
608: The eigenstates are
609: %
610: \begin{eqnarray} \label{e11}
611: |n(x)\rangle \mapsto
612: \frac{1}{\sqrt{S}}
613: \left(
614: \matrix{
615: c_2 \eexp{i\phi} + c_1 E_n \cr
616: 1 - E_n^2 \cr
617: c_1\eexp{-i\phi} + c_2 E_n} \right)
618: \end{eqnarray}
619: %
620: where $S$ is the normalization.
621: Note that for $E=\pm 1$ we have
622: $S=2(c_1 \pm c_2)^2$, while
623: for $E=0$ we have $S\approx 1$.
624: %
625: %
626: %
627: After some algebra we find that
628: the first component of the
629: $\vec{\mbf{B}}$ field in the
630: $\Phi=0$ plane is
631: %
632: \begin{eqnarray} \nonumber
633: {\mbf{B}}^1 =
634: -2\im\left.\left\langle\frac{\partial}{\partial u} n(x) \right|
635: \frac{\partial}{\partial \phi} n(x) \right\rangle =
636: -(c_1^2-c_2^2)
637: \frac{1}{S^2}
638: \frac{\partial S}{\partial u}
639: \end{eqnarray}
640: %
641: Which is illustrated in Fig.3.
642: For a pumping cycle around
643: the \mbox{$u\approx \mp 1$} vertical "chain"
644: the main contribution to $Q$ comes from
645: crossing the $u\approx \mp 1$ line.
646: Hence we get
647: %
648: \begin{eqnarray} \label{e13}
649: Q = \pm \frac{c_1 \pm c_2}{c_1 \mp c_2} =
650: \pm\sqrt{1 \pm 2g_T}
651: \end{eqnarray}
652: %
653: where the Thouless conductance in this context
654: is defined as
655: $g_T = {2c_1c_2}/{(c_1 \mp c_2)^2}$.
656: In both cases we have approximate quantization
657: $Q=\pm 1 + {\cal O}(g_T)$ for $g_T \ll 1$,
658: while for a tight cycle either $Q \rightarrow \infty$
659: or $Q \rightarrow 0$ depending on which
660: line of degeneracies is being encircled.
661: If the pumping cycle encircles both "chains" then
662: we get $Q = {4c_1c_2}/{(c_1^2-c_2^2)}$.
663: In the latter case $Q={\cal O}(g_T)$ for $g_T \ll 1$,
664: with no indication for quantization.
665:
666:
667:
668: %%%%%%%%%%%
669:
670:
671: For a pumping in a dot-wire system (see illustration in Fig.1b),
672: in the limit of a very long wire (many sites) we express the Kubo
673: formula for the conductance matrix using the $S$ matrix
674: of the dot region. The derivation assumes ``quantum chaos",
675: and leads to
676: %
677: \begin{eqnarray} \label{e14}
678: \mbf{G}^{3j} = \frac{1}{2\pi i}
679: \trc\left(P\frac{\partial S}{\partial x_j}
680: S^{\dag}\right)
681: \end{eqnarray}
682: %
683: This is easily identified as the B\"{u}ttiker-Pr\'{e}tre-Thomas formula~\cite{BPT},
684: which has been derived for quantum pumping in open systems (Fig.1e).
685: In particular we get
686: \mbox{$G^{33}=(1/(2\pi\hbar))\trc(PS(1{-}P)S^{\dag})$},
687: which is just the Landauer formula \cite{datta,fisher,stone}.
688:
689:
690:
691: In summary we have shown how the Kubo formalism can
692: be used in order to derive both classical and quantum
693: mechanical results for the pumped charge $Q$ in
694: a closed system. In this formulation the distinction
695: between dissipative and non-dissipative contributions is manifest.
696: %
697: The dissipative contribution to the pumping
698: can be neglected in the adiabatic regime.
699: However, if the adiabaticity condition is violated
700: it does not mean automatically that we have
701: a dissipative effect. Classical pumping by translation
702: is an obvious example.
703: %
704: For the derivation of the
705: dissipative part of the Kubo formula it is essential
706: to realize that in generic circumstances
707: (unlike the case of translations)
708: the adiabatic equation does not possess
709: a stationary solution.
710:
711:
712: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
713:
714: {\bf Acknowledgments:}
715: I thank Y. Avishai (Ben-Gurion University), Y. Avron (Technion)
716: and S. Fishman (Technion) for useful discussions.
717: This research was supported by the Israel Science Foundation (grant No.11/02).
718:
719:
720: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
721: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
722: \begin{thebibliography}{99}
723:
724: \bibitem{pmp}
725: D. Cohen, Phys. Rev. B 68, 155303 (2003).
726:
727: \bibitem{pmo}
728: D. Cohen, Phys. Rev. B 68, 201303(R) (2003).
729:
730: \bibitem{marcus_rev}
731: L.P. Kouwenhoven et al,
732: Proc. of Advanced Study Inst. on Mesoscopic
733: Electron Transport, edited by L.L. Sohn,
734: L.P. Kouwenhoven and G. Schon (Kluwer 1997).
735:
736: \bibitem{thouless}
737: D. J. Thouless, Phys. Rev. {\bf B27}, 6083 (1983).
738:
739: \bibitem{avronI}
740: J. E. Avron and L. Sadun,
741: Phys. Rev. Lett. {\bf 62}, 3082 (1989);
742: \ Ann. Phys. {\bf 206}, 440 (1991).
743: \ J.E. Avron, A. Raveh, and B. Zur Rev. Mod. Phys. {\bf 60}, 873 (1988).
744:
745: \bibitem{aleiner}
746: T. A. Shutenko, I. L. Aleiner and B. L. Altshuler,
747: Phys. Rev. {\bf B61}, 10366 (2000).
748:
749: \bibitem{barriers}
750: Y. Levinson et al, cond-mat/0010494. \\
751: M. Blaauboer and E.J. Heller, cond-mat/0110366.
752:
753: \bibitem{landau}
754: L.D. Landau and E.M. Lifshitz,
755: {\em Statistical physics}, (Buttrworth Heinemann 2000).
756:
757: \bibitem{berry}
758: M.V. Berry, Proc. R. Soc. Lond. A {\bf 392}, 45 (1984).
759:
760: \bibitem{robbins}
761: M.V. Berry and J.M. Robbins,
762: Proc. R. Soc. Lond. A {\bf 442}, 659 (1993).
763:
764: \bibitem{BPT}
765: M. B\"{u}ttiker et al, Z. Phys. {\bf B94}, 133 (1994). \\
766: P. W. Brouwer, Phys. Rev. {\bf B58}, R10135 (1998). \\
767: J. E. Avron et al, Phys. Rev. B {\bf 62}, R10 618 (2000).
768:
769: \bibitem{ora}
770: O. Entin-Wohlman et al, cond-mat/0201073.
771:
772: \bibitem{frc}
773: D. Cohen, Annals of Physics 283, 175 (2000).
774:
775: \bibitem{wlf}
776: A. Barnett et al, J. Phys. A {\bf 34}, 413 (2001).
777:
778:
779: \bibitem{datta}
780: S. Datta, {\em Electronic Transport in Mesoscopic Sytems}
781: (Cambridge University Press 1995).
782:
783: \bibitem{fisher}
784: D.S. Fisher and P.A. Lee, Phys. Rev. B {\bf 23}, 6851 (1981).
785:
786: \bibitem{stone}
787: H.U. Baranger and A.D. Stone,
788: Phys. Rev. B {\bf 40}, 8169 (1989).
789:
790: \bibitem{rmrk}
791: There is some freedom in the
792: definition of the current operator.
793: For example one may define the current
794: through a given section. However,
795: the pumped charge of Eq.(\ref{e6}) comes
796: out the same due to the continuity equation.
797:
798: \end{thebibliography}
799:
800: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
801: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
802:
803:
804: \clearpage
805: \onecolumngrid
806:
807: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
808: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
809: %begin{figure}
810: \centerline{\epsfig{figure=pmp_fig_ae,width=0.55\hsize}}
811: %caption{
812: {\footnotesize FIG1.
813: Illustration of a ring system (a).
814: The shape of the ring is controlled
815: by some parameters $x_1$ and $x_2$.
816: The flux through the ring is $x_3=\Phi$.
817: A system with equivalent topology,
818: and abstraction of the model are
819: presented in (b) and (c).
820: The "dot" can be represented by an $S$ matrix
821: that depends on $x_1$ and $x_2$. In (d) also the
822: flux $x_3$ is regarded as a parameter of the dot.
823: If we "cut" the wire in (d) we get the open
824: two lead geometry of (e).}
825: %end{figure}
826: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
827: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
828:
829: \ \\ \ \\ \ \\
830:
831: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
832: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
833: %begin{figure}[h]
834: \centerline{
835: \epsfig{figure=pmp_delta,height=0.2\hsize}
836: \epsfig{figure=pmp_levels,height=0.2\hsize}
837: }
838: %caption{
839: {\footnotesize FIG2.
840: Schematic illustration of quantum pumping
841: in a closed wire-dot system. The net charge via the third
842: level (thick solid line on the right) is vanishingly
843: small: As the dot potential is lowered an electron
844: is taken from the left side (first avoided crossing),
845: and then emitted back to the left side
846: (second avoided crossing).
847: Assuming that the bias is inverted before the
848: dot potential is raised back, only the second level
849: carry a net charge $Q={\cal O}(1)$.}
850: %end{figure}
851: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
852: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
853:
854:
855: \ \\ \ \\
856:
857: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
858: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
859: %begin{figure}[h]
860: \centerline{\epsfig{figure=pmp_bfield_en,width=0.5\hsize}}
861: \
862: %caption{
863: {\footnotesize FIG3.
864: The first component of the $\mbf{B}$ field
865: for a particle in the middle level of
866: the 3~site lattice model. It is plotted
867: as a function of the dot potential $x_2=u$.
868: The other parameters are $\phi=0$, and $c_1=0.1$,
869: while $c_2=0.04$ for the thick line and
870: $c_2=0.02$ for the thin line.
871: In the limit $c_2 \rightarrow 0$,
872: all the charge that is transfered from the
873: left side into the dot during the first avoided crossing,
874: is emitted back into the left side during the second
875: avoided crossing.
876: Inset: The eigenenergies $E_n(x)$ for
877: the $c_2=0.04$ calculation.}
878: %end{figure}
879: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
880: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
881:
882:
883:
884: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
885: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
886: \end{document}
887:
888:
889:
890:
891:
892:
893:
894:
895:
896:
897:
898:
899:
900: