1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %%% (March 2003 - July 2003) closed systems
3: %%% pmp_fig, pmp_delta + pmp_levels, pmp_bfield_en
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5:
6: \documentclass[aps,pre,twocolumn,floats]{revtex4}
7: \usepackage{epsfig}
8: \usepackage{bm}
9:
10: \begin{document}
11:
12: \newcommand{\hide}[1]{}
13: \newcommand{\tbox}[1]{\mbox{\tiny #1}}
14: \newcommand{\half}{\mbox{\small $\frac{1}{2}$}}
15: \newcommand{\const}{\mbox{const}}
16: \newcommand{\ointt}{\int\!\!\!\!\int\!\!\!\!\!\circ\ }
17: \newcommand{\intt}{\int\!\!\!\!\int }
18: \newcommand{\ar}{\mathsf r}
19: \newcommand{\im}{\mbox{Im}}
20: \newcommand{\re}{\mbox{Re}}
21: \newcommand{\mbf}[1]{#1}
22: \newcommand{\bmsf}[1]{\bm{\mathsf{#1}}}
23: \newcommand{\fullG}{G}
24: \newcommand{\symG}{\eta}
25: \newcommand{\sinc}{\mbox{sinc}}
26: \newcommand{\trc}{\mbox{trace}}
27: \newcommand{\eexp}{\mbox{e}^}
28: \newcommand{\bra}{\left\langle}
29: \newcommand{\ket}{\right\rangle}
30:
31: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
32:
33: \title{
34: Quantum pumping in closed systems, adiabatic transport, and the Kubo formula
35: }
36:
37: \author{Doron Cohen}
38:
39: \affiliation{
40: Department of Physics, Ben-Gurion University, Beer-Sheva 84105, Israel
41: }
42:
43: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
44:
45: \begin{abstract}
46: Quantum pumping in closed systems is considered.
47: We explain that the Kubo formula contains all
48: the physically relevant ingredients for the calculation
49: of the pumped charge ($Q$) within the framework
50: of linear response theory. The relation to
51: the common formulations of adiabatic transport
52: and ``geometric magnetism" is clarified.
53: We distinguish between adiabatic and dissipative
54: contributions to~$Q$.
55: On the one hand we observe that adiabatic pumping does
56: not have to be quantized. On the other hand we
57: define circumstances in which quantized adiabatic pumping
58: holds as an approximation. The deviation from
59: exact quantization is related to the Thouless conductance.
60: As an application we discuss the following examples:
61: classical dissipative pumping by conductance control,
62: classical adiabatic (non dissipative) pumping by translation,
63: and quantum pumping in the double barrier model.
64: In the latter context we analyze a 3~site lattice Hamiltonian,
65: which represents the simplest pumping device.
66: We remark on the connection with the popular $S$ matrix formalism
67: which has been used to calculate pumping in open systems.
68: \end{abstract}
69:
70: \maketitle
71:
72: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
73: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
74: \section{Introduction}
75:
76:
77: Linear response theory (LRT) \cite{landau,imry,datta}
78: is the leading formalism to
79: deal with driven systems. Such systems are described
80: by a Hamiltonian ${\cal H}(\bm{x})$ where $\bm{x}(t)$ is a set
81: of time dependent classical parameters ("fields").
82: The Kubo formula is the corner stone of LRT.
83: It allows the calculation of the response coefficients,
84: and in particular the {\em conductance matrix} ($\bmsf{G}$)
85: of the system. If we know $\bmsf{G}$, we can calculate
86: the charge ($Q$) which is transported through the
87: system during one cycle of a periodic driving.
88: This is called {\em pumping}.
89:
90:
91: Pumping of charge in mesoscopic \cite{marcus_rev}
92: and molecular size devices is regarded as a major
93: issue in the realization of future quantum circuits
94: or quantum gates, possibly for the purpose of
95: quantum computing.
96:
97:
98: %%%%
99: \subsection{Model system}
100:
101: In order to explain the motivation for the present work,
102: and its relation to the published literature,
103: we have to give a better definition of the problem.
104: For presentation purpose we focus on
105: a model system with a ring geometry (Fig.1).
106: The shape of the ring is controlled
107: by some parameters $x_1$ and $x_2$.
108: These parameters can be gate voltages that determine
109: the location of some boundaries, or the height of some barriers.
110: The third parameter is the flux through the ring:
111: %
112: \begin{eqnarray}
113: x_3 \ = \ \Phi \ \equiv \ (\hbar/e) \phi
114: \end{eqnarray}
115: %
116: We shall use units such that the
117: elementary charge is \mbox{$e=1$}.
118: Note that the Hamiltonian ${\cal H}(x_1(t),x_2(t),x_3(t))$
119: has gauge invariance for $\phi \mapsto \phi+2\pi$.
120: %
121: %
122: Another system with a ring topology is presents in Fig.~1b,
123: and its abstraction is represented in Fig.~2c.
124: The ``dot" can be represented by an $S$ matrix
125: that depends on $x_1$ and $x_2$.
126: In Fig.~1d also the flux $x_3$ is regarded as a parameter of the dot.
127: If we cut the wire in Fig.1d we get the open two lead geometry
128: of Fig.1e. Finally we can put many such units in series (no flux),
129: hence getting the periodic system of Fig.1f.
130: In the latter case the Hamiltonian is invariant for
131: unit translations, and therefore the quasi momentum $\phi$
132: is a constant of motion. It follows that the mathematical
133: treatment of a driven periodic structure reduces to
134: an analysis of a driven ring system with flux.
135:
136:
137:
138: %%%%
139: \subsection{Classification of pumps}
140:
141: {\em ``Pumping"} means that net charge (or maybe better
142: to say ``net integrated probability current") is transported
143: through the ring per cycle of a periodic driving.
144: Using the common jargon of electrical engineering
145: this can be described as AC-DC conversion.
146: We distinguish between \\
147: %
148: \begin{minipage}{\hsize}
149: \vspace*{0.1cm}
150: \begin{itemize}
151: \setlength{\itemsep}{0cm}
152: \item
153: pumping in open systems (such as in Fig.~1e).
154: \item
155: pumping in closed systems (such as in Fig.~1d)
156: \item
157: pumping in periodic systems (such as in Fig.~1f)
158: \end{itemize}
159: \vspace*{0.0cm}
160: \end{minipage}
161: %
162: For a reason that was explained at the end of the previous subsection
163: we regard the last category \cite{thouless} as mathematically equivalent
164: to the second category. We also regard the first
165: category \cite{BPT,aleiner,barriers,ora}
166: as a special (subtle) limit of the second category:
167: in a follow up paper \cite{pmo} we demonstrate
168: that in the limit of open geometry the Kubo formula
169: reduces to the $S$-matrix formula
170: of B\"{u}ttiker Pr\'{e}tre and Thomas \cite{BPT}.
171:
172:
173:
174: There are works in the literature regarding
175: ``rectification" and ``ratchets" \cite{ratch}.
176: These can be regarded as studies of
177: pumping in periodic systems
178: with the connotation of having {\em damped}
179: non-Hamiltonian dynamics. These
180: type of systems are beyond the scope
181: of the present Paper.
182: %
183: There is also a recent interest in Hamiltonian Ratchets\cite{ratchets},
184: which is again a synonym for pumping in periodic systems,
185: but with the connotation of having a {\em non-linear}
186: pumping mechanism.
187: We are going to clarify what are the conditions for having
188: a {\em linear} pumping mechanism. Only in case of linear pumping mechanism
189: the Kubo formula can be used, which should be distinguished
190: from the non-linear mechanism of Ref.\cite{ratchets}.
191:
192:
193:
194: %%%%%%%%%%%%
195: \subsection{Objectives}
196:
197: The purpose of this Paper is to explain and demonstrate
198: that the Kubo formula contains all
199: the physically relevant ingredients for
200: the calculation of the charge ($Q$) which is pumped
201: during one cycle of a periodic driving.
202: %
203: %
204: In the limit of a very slow time variation (small $\dot{x}$),
205: the emerging picture coincides with
206: the adiabatic picture of Refs.\cite{berry,thouless,avron,robbins}.
207: In this limit the response of the system
208: is commonly described as a non-dissipative
209: ``geometric magnetism" \cite{robbins} effect,
210: or as adiabatic transport.
211:
212:
213: A major objective of this Paper is to bridge between
214: the adiabatic picture and the more general LRT / Kubo picture,
215: and to explain how {\em dissipation} emerges
216: in the quantum mechanical treatment.
217: %
218: For one-parameter driving a unifying picture that bridges between
219: the quantum mechanical adiabatic picture and LRT has been presented
220: in \cite{vrn,dsp,crs,frc}. A previous attempt \cite{wilk} had ended
221: in some confusion regarding the identification of the linear
222: response regime, while \cite{robbins} had avoided the analysis of
223: the mechanism that leads to dissipation in the quantum mechanical case.
224:
225:
226: The presented (Kubo based) formulation of the pumping problem
227: has few advantages:
228: It is not restricted to the adiabatic regime;
229: It allows a clear distinction between {\em dissipative} and {\em adiabatic} contributions to the pumping;
230: The classical limit is manifest in the formulation;
231: It gives a level by level understanding of the pumping process;
232: It allows the consideration of any type of occupation (not necessarily Fermi occupation);
233: It allows future incorporation of external environmental influences such as that of noise;
234: It regards the voltage over the pump as electro motive force,
235: rather than adopting the conceptually more complicated view \cite{ora}
236: of having a chemical potential difference.
237:
238:
239: Of particular interest is the possibility to realize
240: a pumping cycle that transfers {\em exactly} one unit of charge
241: per cycle. In open systems \cite{aleiner,barriers}
242: this ``quantization" holds only approximately,
243: and it has been argued \cite{aleiner} that the deviation from
244: exact quantization is due to the dissipative effect.
245: Furthermore it has been claimed \cite{aleiner}
246: that exact quantization would hold in the strict adiabatic limit,
247: if the system were {\em closed}.
248: %
249: In this Paper we would like to show that the correct picture is quite different.
250: We shall demonstrate that the deviation from exact quantization is in fact
251: of adiabatic nature. This deviation is related
252: to the so-called ``Thouless conductance" of the device.
253:
254:
255:
256:
257: %%%%%%%%%
258: \subsection{Examples}
259:
260: We give several examples for the application
261: of the Kubo formula to the calculation of
262: the pumped charge~$Q$: \\
263: %
264: \begin{minipage}{\hsize}
265: \vspace*{0.1cm}
266: \begin{itemize}
267: \setlength{\itemsep}{0cm}
268: \item
269: classical dissipative pumping
270: \item
271: classical adiabatic pumping (by translation)
272: \item
273: quantum pumping in the double barrier model
274: \end{itemize}
275: \vspace*{0.0cm}
276: \end{minipage}
277: %
278: The last example is the main one.
279: In the context of open geometry it is known
280: as ``pumping around a resonance" \cite{barriers}.
281: We explain that this is in fact
282: an diabatic transfer scheme,
283: and we analyze a particular version
284: of this model which is represented
285: by a 3~site lattice Hamiltonian.
286: This is definitely the simplest pump circuit
287: possible, and we believe that it can be realized
288: as a molecular size device.
289: It also can be regarded as an approximation
290: for the closed geometry version of the
291: two delta potential pump \cite{barriers}.
292:
293:
294:
295:
296: %%%%%%%%%
297: \subsection{Outline}
298:
299: In Section~2 we define the main object of the study,
300: which is the conductance matrix $\bmsf{G}$ of Eq.(\ref{e_8}).
301: The conductance matrix can be written as the sum
302: of a symmetric ($\bm{\eta}$) and an anti-symmetric ($\bmsf{B}$)
303: matrices, which are later identified as the dissipative and
304: the adiabatic contributions respectively.
305:
306: In the first part of the paper (Sections~2-8)
307: we analyze the adiabatic equation (Section~3),
308: and illuminate the distinction
309: between its zero order solution (Section~4),
310: its stationary first order solution (Section~5),
311: and its non-stationary solution (Section~6).
312: %
313: %
314: The outcome of the analysis in Section~5 is Eq.(\ref{e_30})
315: for the conductance matrix $\bmsf{G}$.
316: This expression is purely adiabatic, and does
317: not give any dissipation. In order to get dissipation
318: we have to look for a non-stationary solution.
319:
320: The standard textbook derivation of the Kubo formula (Eq.(\ref{e_33}))
321: for the conductance matrix $\bmsf{G}$ {\em implicitly} assumes
322: a non-stationary solution. We show how to get from it
323: Eq.(\ref{e_34}) for $\bm{\eta}$ and Eq.(\ref{e_35}) for $\bmsf{B}$.
324: The latter is shown to be identical with the adiabatic result (Eq.(\ref{e_30})).
325: In Section~7 we further simplify the expression for $\bm{\eta}$
326: leading to the fluctuation-dissipation relation (Eq.(\ref{e43})).
327:
328:
329: The disadvantages of the standard textbook derivation
330: of Kubo formula make it is essential to introduce a different
331: route toward Eq.(\ref{e43}) for $\bm{\eta}$.
332: This route, which is discusssed in Section~8,
333: {\em explicitly} distinguishes the dissipative effect from the adiabatic effect,
334: and allows to determine the conditions for the validity
335: of either the adiabatic picture or LRT.
336: In particular it is explained that LRT is based,
337: as strange as it sounds, on perturbation theory to infinite order.
338:
339:
340: In Section~9 we clarify the general scheme of
341: the pumping calculation (Eq.(\ref{e6})).
342: Section~10 and Section~11 give two simple classical examples.
343: In Section~12 we turn to discuss quantum pumping,
344: where the cycle is around a chain of degeneracies.
345: The general discussion is followed by presentation
346: of the double barrier model (Section~13).
347: In order to get a quantitative estimate for the pumped charge
348: we consider a 3~site lattice Hamiltonian (Section~14).
349:
350: The summary (Section~15) gives some larger perspective on the subject,
351: pointing out the relation to the $S$-matrix formalism,
352: and to the Born-Oppenheimer picture.
353: In the appendices we give some more details regarding the
354: derivations, so as to have a self-contained presentation.
355:
356:
357:
358: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
359: \section{The conductance matrix}
360:
361: Consider the Hamiltonian ${\cal H}(\bm{x}(t))$, where $\bm{x}(t)$
362: is a set of time dependent parameters ("fields").
363: For presentation, as well as for practical reasons,
364: we assume later a set of three time dependent parameters
365: \mbox{$\bm{x}(t)=(x_1(t),x_2(t),x_3(t))$}.
366: We define generalized forces in the conventional way as
367: %
368: \begin{eqnarray}
369: F^k \ \ = \ \ -\frac{\partial {\cal H}}{\partial x_k}
370: \end{eqnarray}
371: %
372: Note that if $x_1$ is the location of a
373: wall element, then $F^1$ is the force in the
374: Newtonian sense. If $x_2$ is an electric field,
375: then $F^2$ is the polarization.
376: If $x_3$ is the magnetic field or the flux through a ring,
377: then $F^3$ is the magnetization or the current through the ring.
378:
379:
380:
381: In linear response theory (LRT) the response of the system
382: is described by a causal response kernel, namely
383: %
384: \begin{eqnarray} \label{e_3}
385: \langle F^k \rangle_t \ = \ \sum_j \int_{-\infty}^{\infty} \alpha^{kj}(t-t') \ x_j(t')dt'
386: \end{eqnarray}
387: %
388: where $\alpha^{kj}(\tau)=0$ for $\tau<0$.
389: The Fourier transform of $\alpha^{kj}(\tau)$
390: is the generalized susceptibility $\chi^{kj}(\omega)$.
391: %
392: The conductance matrix is defined as:
393: %
394: \begin{eqnarray}
395: \fullG^{kj} \ = \ \lim_{\omega\rightarrow 0}
396: \frac{\im[\chi^{kj}(\omega)]}{\omega}
397: \ = \ \int_0^{\infty} \alpha^{kj}(\tau) \tau d\tau
398: \end{eqnarray}
399: %
400: Consequently, as explained further in Appendix~B,
401: the response in the ``DC limit" ($\omega\rightarrow0$)
402: can be written as
403: %
404: \begin{eqnarray} \label{e_8}
405: \langle F^k \rangle =
406: -\sum_{j} \fullG^{kj} \ \dot{x}_j
407: \end{eqnarray}
408: %
409: As an example for the applicability of this formula
410: note the following standard examples for one parameter driving:
411: Let $x=$ wall or piston displacement, then $\dot{x}=$ wall or piston velocity,
412: $G=$ friction coefficient, and $F=-G \dot{x}$ is the friction force.
413: Another standard example is $x=$ magnetic flux,
414: $-\dot{x}=$ electro motive force, $G=$ electrical conductance,
415: and hence $F=-G \dot{x}$ is Ohm law.
416:
417:
418:
419:
420: It is convenient to write the conductance matrix
421: as $\fullG^{kj} \equiv \ \symG^{kj} + \mbf{B}^{kj}$,
422: where $\symG^{kj}=\symG^{jk}$ is the symmetric part
423: of the conductance matrix,
424: while $\mbf{B}^{kj}=-\mbf{B}^{jk}$ is the antisymmetric part.
425: In case of having three parameters we can arrange the
426: elements of the antisymmetric part
427: as a vector \mbox{$\vec{\bm{B}}=(B^{23},B^{31},B^{12})$}.
428: %
429: Consequently Eq.(\ref{e_8}) can be written in abstract notation as
430: %
431: \begin{eqnarray}
432: \langle \bm{F} \rangle \ = \ -\bm{\symG} \cdot \dot{\bm{x}} \ - \ \bmsf{B}\wedge \dot{\bm{x}}
433: \end{eqnarray}
434: %
435: where the dot product should be interpreted as matrix-vector
436: multiplication, which involves summation over the index~$j$.
437: The wedge-product also can be regarded as a matrix-vector
438: multiplication. It reduces to the more familiar cross-product
439: in case that we consider $3$ parameters.
440: The dissipation, which is defined as the rate in which energy
441: is absorbed into the system, is given by
442: %
443: \begin{eqnarray}
444: \dot{{\cal W}} = \frac{d}{dt}\langle {\cal H} \rangle
445: \ = \ -\langle \bm{F} \rangle \cdot \dot{\bm{x}} \ = \
446: \sum_{kj} \symG^{kj} \ \dot{x}_k\dot{x}_j
447: \end{eqnarray}
448: %
449: Only the symmetric part contributes
450: to the the dissipation. The contribution
451: of the antisymmetric part is identically zero.
452:
453:
454: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
455: \section{The adiabatic equation}
456:
457: The adiabatic equation is conventionally obtained from the
458: Schrodinger equation by expanding the wavefunction
459: in the $x$-dependent adiabatic basis:
460: %
461: \begin{eqnarray}
462: \frac{d}{dt}|\psi\rangle \ \ &=&
463: \ -\frac{i}{\hbar} {\cal H}(\bm{x}(t)) \ |\psi\rangle
464: \\
465: |\psi\rangle \ \ &=& \ \ \sum_n a_n(t) \ |n(\bm{x}(t))\rangle
466: \\
467: \frac{da_n}{dt} &=& -\frac{i}{\hbar} E_n a_n
468: +\frac{i}{\hbar} \sum_m \sum_j
469: \dot{x}_j \mbf{A}^{j}_{nm} a_m
470: \end{eqnarray}
471: %
472: where following \cite{berry} we define
473: %
474: \begin{eqnarray}
475: \mbf{A}^{j}_{nm}(\bm{x}) =
476: i\hbar \left\langle n(\bm{x}) \Big|
477: \frac{\partial}{\partial x_j} m(\bm{x})\right\rangle
478: \end{eqnarray}
479: %
480: Differentiation by parts of
481: $\partial_j \langle n(\bm{x})|m(\bm{x}) \rangle = 0$
482: leads to the conclusion that $\mbf{A}^{j}_{nm}$
483: is a hermitian matrix.
484: Note that the effect of gauge transformation is
485: %
486: \begin{eqnarray}
487: |n(\bm{x})\rangle \ & \ \mapsto \ & \ \eexp{-i\frac{\Lambda_n(\bm{x})}{\hbar}} \ |n(\bm{x})\rangle \\
488: \mbf{A}^{j}_{nm} \ & \ \mapsto \ & \
489: \eexp{i\frac{\Lambda_n{-}\Lambda_m}{\hbar}} \mbf{A}^{j}_{nm}
490: + (\partial_j \Lambda_n)\delta_{nm}
491: \end{eqnarray}
492: %
493: Note that the diagonal elements $\mbf{A}^{j}_{n} \equiv\mbf{A}^{j}_{nn}$
494: are real, and transform as
495: $\mbf{A}^{j}_{n} \mapsto \mbf{A}^{j}_{n} + \partial_j \Lambda_n$.
496:
497:
498:
499: Associated with $\bm{A}_n(\bm{x})$
500: is the gauge invariant 2-form, which is defined as:
501: %
502: \begin{eqnarray}
503: \mbf{B}^{ij}_{n} &=&
504: \partial_i \mbf{A}^{j}_{n} - \partial_j \mbf{A}^{i}_{n} \\
505: &=& -2\hbar \im\langle\partial_i n|\partial_j n\rangle \\
506: &=& -\frac{2}{\hbar}\im\sum_{m}
507: \mbf{A}^{i}_{nm}\mbf{A}^{j}_{mn}
508: \end{eqnarray}
509: %
510: This can be written in an abstract notation as $\bmsf{B}=\nabla\wedge\bm{A}$.
511: %
512: %
513: Using standard manipulations, namely via
514: differentiation by parts of
515: $\partial_j \langle n(\bm{x})|{\cal H}|m(\bm{x}) \rangle = 0$,
516: we get for $n\ne m$ the expressions:
517: %
518: \begin{eqnarray}
519: \mbf{A}^{j}_{nm}(\bm{x}) = \frac{i\hbar}{E_m{-}E_n}
520: \left\langle n \left|\frac{\partial {\cal H}}{\partial x_j}\right|m\right\rangle
521: \equiv-\frac{i\hbar F^j_{nm}}{E_m{-}E_n}
522: \end{eqnarray}
523: %
524: and hence
525: %
526: \begin{eqnarray} \label{e29}
527: \mbf{B}^{ij}_{n} = 2\hbar \sum_{m(\ne n)}
528: \frac{\im\left[
529: F^i_{nm}F^j_{mn}\right]}
530: {(E_m-E_n)^2}
531: \end{eqnarray}
532:
533:
534:
535: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
536: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
537: \section{The strictly adiabatic solution, and the Berry phase}
538:
539: We define the perturbation matrix as
540: %
541: \begin{eqnarray} \label{e_23}
542: \mbf{W}_{nm}=-\sum_j\dot{x}_j\mbf{A}^j_{nm}
543: \ \ \ \ \ \ \ \ \mbox{for $n\ne m$}
544: \end{eqnarray}
545: %
546: and $\mbf{W}^{j}_{nm}=0$ for $n=m$.
547: Then the adiabatic equation can be re-written
548: as follows:
549: %
550: \begin{eqnarray} \label{e_24}
551: \frac{da_n}{dt} = -\frac{i}{\hbar} (E_n{-}\dot{x}A_n) a_n
552: -\frac{i}{\hbar} \sum_m
553: \mbf{W}_{nm} a_m
554: \end{eqnarray}
555: %
556: If we neglect the perturbation $W$,
557: then we get the strict adiabatic solution:
558: %
559: \begin{eqnarray}
560: \eexp{
561: -\frac{i}{\hbar} \left(
562: \int_0^t E_n(\bm{x}(t'))dt'
563: -\int_{\bm{x}(0)}^{\bm{x}(t)} \mbf{A}_n(\bm{x})\cdot d\bm{x}
564: \right)
565: }
566: \ |n(\bm{x}(t))\rangle
567: \end{eqnarray}
568: %
569: Due to $\bm{A}_n(\bm{x})$, we have the so called geometric phase.
570: This can be gauged away unless we consider a closed
571: cycle. For a closed cycle, the gauge invariant
572: phase $(1/\hbar) \oint \bm{A}\cdot \vec{d\bm{x}}$ is called Berry phase.
573:
574:
575: With the above zero-order solution
576: we can obtain the following result:
577: %
578: \begin{eqnarray}
579: \langle F^k \rangle &=&
580: \left\langle n(\bm{x}) \left|-\frac{\partial {\cal H}}{\partial x_k} \right| n(\bm{x}) \right\rangle
581: \\ &=&
582: -\frac{\partial}{\partial x_k}
583: \left\langle n(\bm{x})|\ {\cal H}(\bm{x}) \ |n(\bm{x})\right\rangle
584: \end{eqnarray}
585: %
586: In case of the standard examples that were mentioned previously
587: this corresponds to conservative force or to persistent current.
588: From now on we ignore this
589: trivial contribution to $\langle F^k \rangle$ ,
590: and look for the a first order contribution.
591:
592:
593: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
594: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
595: \section{The stationary adiabatic solution: Adiabatic Transport or ``Geometric Magnetism"}
596:
597: For linear driving (unlike the case of a cycle) the
598: $\bm{A}_n(\bm{x})$ field can be gauged away.
599: Assuming further that the adiabatic equation
600: can be treated as parameter independent
601: (that means disregarding the parametric dependence
602: of $E_n$ and ${W}$ on $\bm{x}$) one realizes
603: that Eq.(\ref{e_24}) possesses stationary solutions.
604: To first order these are:
605: %
606: \begin{eqnarray} \label{e_28}
607: | \psi \rangle \ = \
608: |n\rangle +
609: \sum_{m(\ne n)}
610: \frac{\mbf{W}_{mn}}
611: {E_n-E_m} |m\rangle
612: \end{eqnarray}
613: %
614: Note that in a {\em fixed-basis representation}
615: the above stationary solution is in fact time-dependent.
616: Hence the notations $|n(\bm{x}(t))\rangle$, $|m(\bm{x}(t))\rangle$
617: and $| \psi(t) \rangle$ are possibly more appropriate.
618:
619:
620: With the above solution
621: we can write $\langle F^k \rangle$ as a sum of
622: zero order and first order contributions.
623: From now on we ignore the
624: zero order contribution,
625: and go on with the first order contribution:
626: %
627: \begin{eqnarray}
628: \langle F^k \rangle &=&
629: -\sum_{m(\ne n)}
630: \frac{\mbf{W}_{mn}}{E_n-E_m}
631: \left\langle n \Big|
632: \frac{\partial {\cal H}}{\partial x_k}
633: \Big| m \right\rangle + \mbox{CC}
634: \nonumber \\
635: &=& \sum_j \left(i\sum_{m}
636: \mbf{A}^k_{nm} \mbf{A}^{j}_{mn} + \mbox{CC}\right) \ \dot{x}_j
637: \nonumber \\
638: &=& -\sum_j \mbf{B}_n^{kj} \ \dot{x}_j
639: \end{eqnarray}
640: %
641: %
642: For a general {\em stationary} preparation,
643: either pure or mixed, one obtains Eq.(\ref{e_8}) with
644: %
645: \begin{eqnarray} \label{e_30}
646: \fullG^{kj} \ = \ \sum_n f(E_n) \ \mbf{B}_n^{kj}
647: \end{eqnarray}
648: %
649: where $f(E_n)$ are weighting factors,
650: with the normalization $\sum_n f(E_n)=1$.
651: For a pure state preparation $f(E_n)$ distinguishes
652: only one state $n$, while for canonical
653: preparation $f(E_n)\propto\exp(-E_n/T)$,
654: where $T$ is the temperature.
655: For a many-body system of non-interacting Fermions
656: $f(E_n)$ can be re-interpreted as
657: the Fermi occupation function,
658: so that $\sum_n f(E_n)$ is the total
659: number of particles.
660:
661:
662: Thus we see that the assumption of a stationary
663: first-order solution leads to a non-dissipative (antisymmetric)
664: conductance matrix This is know as either
665: Adiabatic Transport \cite{thouless,avron}
666: or ``Geometric Magnetism" \cite{berry}.
667: In the later sections we shall discuss
668: the limitations of the above result.
669:
670:
671: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
672: \section{The non-stationary solution: The Kubo formula}
673:
674:
675: The Kubo formula is an expression for the linear response
676: of a driven system that goes beyond the stationary
677: adiabatic solution of the previous section.
678: The Kubo formula has many type of derivations.
679: One possibility is to use the same procedure
680: as in Section~5 starting with
681: %
682: \begin{eqnarray} \nonumber
683: | \psi(t) \rangle \ &=& \
684: \eexp{-iE_nt}|n\rangle \\ \nonumber
685: &+& \sum_{m(\ne n)}
686: \left[-i\mbf{W}_{mn}\int_0^t\eexp{i(E_n{-}E_m)t'}dt'\right]
687: \eexp{-iE_mt} |m\rangle
688: \end{eqnarray}
689: %
690: For completeness we also give in Appendix~A
691: a simple version of the standard derivation, which is based
692: on a conventional fixed-basis first-order-treatment
693: of the perturbation. The disadvantages are: \\
694: %
695: \begin{minipage}{\hsize}
696: \vspace*{0.1cm}
697: \begin{itemize}
698: \setlength{\itemsep}{0cm}
699: \item
700: The standard derivation does not illuminate the underlaying
701: physical mechanisms of the response.
702: \item
703: The stationary adiabatic limit is not manifest.
704: \item
705: The fluctuation-dissipation relation is vague.
706: \item
707: The validity conditions of the derivation are not clear:
708: no identification of the {\em regimes}.
709: \end{itemize}
710: \vspace*{0.0cm}
711: \end{minipage}
712: %
713: For now we go on with the conventional approach,
714: but in a later section we refer to the more illuminating
715: approach of \cite{wilk} and \cite{vrn,dsp,crs,frc}.
716: Then we clarify what is the regime (range of $\dot{x}$)
717: were the Kubo formula can be trusted \cite{crs,frc},
718: and what is the sub-regime where the response can be
719: described as non-dissipative adiabatic transport.
720:
721:
722: In order to express the Kubo formula one
723: introduces the following definition:
724: %
725: \begin{eqnarray}
726: K^{ij}(\tau) = \frac{i}{\hbar} \langle [F^i(\tau),F^j(0)]\rangle
727: \end{eqnarray}
728: %
729: We use the common interaction picture notation
730: \mbox{$F^k(\tau)= \eexp{i{\cal H}t} F^k \eexp{-i{\cal H}t}$},
731: where ${\cal H}={\cal H}(\bm{x})$ with $x=\mbox{const}$.
732: %
733: The expectation value assumes that
734: the system is prepared in a {\em stationary} state
735: (see previous section).
736: It is also implicitly assumed that the result is
737: not sensitive to the exact value of $x$.
738: %
739: Note that $K^{ij}(\tau)$ has
740: a well defined classical limit.
741: Its Fourier transform will be denoted
742: $\tilde{K}^{ij}(\omega)$.
743:
744:
745:
746:
747: The expectation value $\langle F^{k} \rangle$
748: is related to the driving $\bm{x}(t)$ by
749: the causal response kernel \mbox{$\alpha^{ij}(t-t')$}.
750: The Kubo expression for this response kernel,
751: as derived in Appendix~A, is
752: %
753: \begin{eqnarray} \label{e_kubo}
754: \alpha^{ij}(\tau) \ \ = \ \
755: \Theta(\tau) \ K^{ij}(\tau)
756: \end{eqnarray}
757: %
758: where the step function $\Theta(\tau)$ cares
759: for the upper cutoff of the integration in Eq.(\ref{e_3}).
760: The Fourier transform of $\alpha^{ij}(\tau)$
761: is the generalized susceptibility $\chi^{ij}(\omega)$.
762: The conductance matrix is defined as:
763: %
764: \begin{eqnarray} \label{e_33}
765: \fullG^{ij} \ = \ \lim_{\omega\rightarrow 0}
766: \frac{\im[\chi^{ij}(\omega)]}{\omega} \ = \
767: \int_0^{\infty} K^{ij}(\tau)\tau d\tau
768: \end{eqnarray}
769: %
770: This can be split into symmetric
771: and anti-symmetric components
772: (see derivation in Appendix~C)
773: as follows:
774: %
775: \begin{eqnarray} \label{e_34}
776: \symG^{ij} = \half(\fullG^{ij}{+}\fullG^{ji}) &=&
777: \frac{1}{2}\lim_{\omega\rightarrow 0}
778: \frac{\im[\tilde{K}^{ij}(\omega)]}{\omega}
779: \\ \label{e_35}
780: \mbf{B}^{ij} = \half(\fullG^{ij}{-}\fullG^{ji}) &=&
781: {-}\int_{-\infty}^{\infty}\frac{d\omega}{2\pi}
782: \frac{\re[\tilde{K}^{ij}(\omega)]}{\omega^2}
783: \end{eqnarray}
784: %
785: The antisymmetric part is identified \cite{robbins} (Appendix~C)
786: as corresponding to the stationary solution Eq.(\ref{e_30})
787: of the adiabatic equation.
788:
789:
790:
791:
792: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
793: \section{The emergence of dissipative response, and the fluctuation-dissipation relation}
794:
795:
796: The Kubo formula for the symmetric part of the
797: conductance matrix ($\symG^{ij}$) can be further simplified.
798: If we take Eq.(\ref{e_34}) literally, then $\symG^{ij}=0$ due to
799: the simple fact that we have finite spacing between
800: energy levels (see \cite{ophir} for a statistical point of view).
801: But if we assume that the energy levels
802: have some finite width $\Gamma$, then the smoothed
803: version of $\tilde{K}^{ij}(\omega)$ should
804: be considered. In common textbooks the introduction
805: of $\Gamma$ is ``justified" by assuming
806: some weak coupling to an environment,
807: or by taking the limit of infinite volume.
808: But we are dealing with a strictly isolated finite
809: system, and therefore the meaning of $\Gamma$
810: requires serious consideration.
811: We postpone the discussion of this issue
812: to the next section.
813:
814:
815: If the smoothed version of $\tilde{K}^{ij}(\omega)$
816: should be used in Eq.(\ref{e_34}), then it is possible to obtain
817: $\symG^{ij}$ from power spectrum $\tilde{C}^{ij}(\omega)$
818: of the fluctuations. This is called the Fluctuation-Dissipation relation.
819: The spectral function $\tilde{C}^{ij}(\omega)$
820: is defined as the Fourier transform of the
821: symmetrized correlation function
822: %
823: \begin{eqnarray}
824: C^{ij}(\tau) = \langle \half (F^i(\tau)F^j(0)+F^j(0)F^i(\tau)) \rangle
825: \end{eqnarray}
826: %
827: We use again the interaction picture,
828: as in the definition of $K^{ij}(\tau)$.
829: Also this function has a well defined classical limit.
830:
831:
832:
833: There are several versions for the
834: Fluctuation-Dissipation relation.
835: The microcanonical version \cite{robbins}
836: has been derived using classical considerations,
837: leading to
838: %
839: \begin{eqnarray} \label{e43}
840: \symG^{ij}|_E \ = \
841: \frac{1}{2}\frac{1}{g(E)}
842: \frac{d}{dE}\left[g(E) \ \tilde{C}_E^{ij}(\omega\rightarrow0)\right]
843: \end{eqnarray}
844: %
845: In Appendix~C we introduce its quantum mechanical derivation.
846: The subscript emphasizes that we assume
847: a microcanonical state with energy $E$,
848: and $g(E)$ is the density of states.
849: %
850: The traditional version of the Fluctuation-Dissipation relation
851: assumes a canonical state. It can be obtained
852: by canonical averaging over the microcanonical version leading to:
853: %
854: \begin{eqnarray}
855: \symG^{ij}|_T \ = \
856: \frac{1}{2T}\tilde{C}_T^{ij}(\omega\rightarrow0)
857: \end{eqnarray}
858:
859:
860:
861: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
862: \section{The validity of linear response theory and beyond}
863:
864:
865: The standard derivation of the dissipative part of the Kubo formula,
866: leading to the Fluctuation-Dissipation relation Eq.(\ref{e43}),
867: is not very illuminating physically.
868: More troubling is the realization that one cannot tell
869: from the standard derivation what are the conditions for its validity.
870: An alternate, physically appealing derivation \cite{vrn,dsp,crs,frc},
871: is based on the observation that energy absorption is related
872: to having diffusion in energy space \cite{wilk}.
873: In Appendix~E we outline the main ingredients of this approach.
874:
875:
876: It should be clear that the diffusion picture of Appendix~E
877: holds only in case of {\em chaotic systems}. If this diffusion picture
878: does not hold, then also the Kubo formula for $\symG^{ij}$ does not hold!
879: Driven one-dimensional systems are the obvious example
880: for the failure of linear response theory (LRT).
881: As in the case of the kicked rotator (standard map) \cite{qkr}
882: there is a complicated route to chaos and stochasticity:
883: By increasing the driving amplitude the phase space
884: structure is changed. If the amplitude is smaller than
885: a threshold value, then the diffusion is blocked by Kolmogorov-Arnold-Moser curves,
886: and consequently there is not dissipation.
887: Therefore the Kubo formula is not applicable in such cases.
888:
889:
890: The following discussion of {\em dissipative response}
891: assumes that we deal with a quantized chaotic system.
892: We would like to discuss the reason and the consequences
893: of having an energy scale $\Gamma$. In the standard derivation
894: the assumption of having level broadening as if comes out of the blue.
895: As we already noted it is customary in textbooks to argue
896: that either a continuum limit, or some small coupling to an environment
897: is essential in order to provide $\Gamma$. But this is of course
898: just a way to avoid confrontation with the physical problem
899: of having a driven {\em isolated finite mesoscopic system}.
900: In fact the energy scale $\Gamma$ is related to the rate ($\dot{x}$)
901: of the driving:
902: %
903: \begin{eqnarray} \label{e_42}
904: \Gamma \ \ = \ \
905: \left(\frac{\hbar\sigma}{\Delta^2}|\dot{x}|\right)^{2/3} \times \Delta
906: \end{eqnarray}
907: %
908: where for simplicity of presentation we assume
909: one parameter driving. We use $\Delta$ to denote the
910: mean level spacing, and $\sigma$ is the root mean
911: square value of the matrix element $F_{nm}$
912: between neighboring levels.
913: %
914: In order to derive the above expression for $\Gamma$
915: we have used the result of \cite{frc} (Sec.17) for the ``core width"
916: at the breaktime $t=t_{\tbox{prt}}$ of perturbation theory.
917: %
918: The purpose of the present section is to give an optional
919: "pedestrian derivation" for $\Gamma$, and to discuss
920: the physical consequences.
921:
922:
923:
924: Looking at the first order solution Eq.(\ref{e_28}) of Section~5
925: one realizes that it makes sense provided $|\mbf{W}_{mn}| \ll \Delta$.
926: This leads to the adiabaticity condition
927: %
928: \begin{eqnarray} \label{e91}
929: |\dot{x}| \ \ \ll \ \ \frac{\Delta^2}{\hbar\sigma}
930: \end{eqnarray}
931: %
932: If this condition is not satisfied one should
933: go beyond first order perturbation theory (FOPT),
934: in a sense to be explained below.
935: Note that this adiabaticity condition can be
936: written as $\Gamma\ll\Delta$.
937:
938:
939: The adiabaticity condition Eq.(\ref{e91}) can be explained
940: in a more illuminating way as follows:
941: Let us assume that we prepare the system at time $t=0$
942: at the level $|n\rangle$.
943: Using {\em time dependent} FOPT we find out that
944: a stationary-like solution is reached after
945: the Heisenberg time $t_{\tbox{H}}=2\pi\hbar/\Delta$.
946: This is of course a valid description
947: provided we do not have by then a breakdown
948: of FOPT. The condition for this is
949: easily found to be $\dot{x}t_{\tbox{H}} \ll \delta x_c$,
950: where $\delta x_c=\Delta/\sigma$. This
951: leads again to the adiabaticity condition Eq.(\ref{e91}).
952:
953:
954:
955: Another assumption in the derivation of
956: Section~5 was that we can
957: ignore the parametric dependence
958: of $E_n$ and ${W}$ on $\bm{x}$.
959: The adiabaticity condition
960: $t_{\tbox{H}} \ll \delta x_c/\dot{x}$
961: manifestly justify such an assumption:
962: We should think of $t_{\tbox{H}}$ as
963: the transient time for getting
964: a stationary-like state, and we
965: should regard $\delta x_c$ as the parametric
966: correlation scale.
967:
968:
969:
970: As strange as it sounds, in order to
971: have dissipation, it is essential to have
972: a breakdown of FOPT. In the language
973: of perturbation theory this implies
974: a required summation of diagrams to {\em infinite order},
975: leading to an effective broadening of the energy levels.
976: By iterating FOPT, neglecting interference terms,
977: we get a {\em Markovian approximation}
978: for the energy spreading process.
979: This leads to the diffusion equation of Appendix~E.
980: This diffusion can be regarded as arising
981: from Fermi-golden-rule transitions between
982: energy levels.
983: %
984: A simple ad-hoc way to determine the
985: energy level broadening is to introduce $\Gamma$
986: as a lower cutoff in the energy distribution
987: which is implied by Eq.(\ref{e_28}):
988: %
989: \begin{eqnarray}
990: |\langle n | \psi \rangle|^2 \ = \
991: \frac{|\hbar \dot{x} \mbf{F}_{mn}|^2}
992: {(E_n-E_m)^4 + (\Gamma)^4}
993: \end{eqnarray}
994: %
995: This constitutes a generalization of the well known
996: procedure used by Wigner in order to obtain
997: the local density of states \cite{vrn,dsp}.
998: However, in the present context we do not get a Lorentzian.
999: The width parameter $\Gamma$ is determined
1000: self consistently from normalization,
1001: leading to Eq.(\ref{e_42}) [disregarding numerical prefactor].
1002:
1003:
1004: We can summarize the above reasoning by saying
1005: that there is a {\em perturbative regime}
1006: that includes an adiabatic (FOPT) sub-regime.
1007: Outside of the adiabatic sub-regime we need {\em all orders}
1008: of perturbation theory leading to Fermi-golden-rule
1009: transitions, diffusion in energy space, and hence dissipation.
1010: %
1011: Thus the dissipative part of Kubo formula
1012: emerges only in the regime $\Gamma>\Delta$,
1013: which is just the opposite of the adiabaticity condition.
1014: %
1015: The next obvious step is to determine
1016: the boundary of the perturbative regime.
1017: Following \cite{crs,frc} we argue
1018: that the required condition is $\Gamma\ll\Delta_b$.
1019: The bandwidth $\Delta_b\propto\hbar$
1020: is defined as the energy width
1021: $|E_n-E_m|$ were the matrix
1022: elements $F_{nm}$ are not vanishingly small.
1023: If the condition $\Gamma\ll\Delta_b$
1024: is violated we find ourselves
1025: in the non-perturbative regime
1026: where the Kubo formula cannot be trusted \cite{crs,frc}.
1027:
1028:
1029:
1030: We still have to illuminate why we can
1031: get in the perturbative regime
1032: a dissipative {\em linear} response
1033: in spite of the breakdown of FOPT.
1034: The reason is having a separation of scales
1035: ($\Delta \ll \Gamma\ll\Delta_b$).
1036: The non-perturbative mixing on
1037: the small energy scale $\Gamma$ does not
1038: affect the rate of first-order
1039: transitions between distant levels
1040: \mbox{($\Gamma \ll |E_n-E_m| \ll \Delta_b$)}.
1041: Therefore Fermi golden rule picture
1042: applies to the description
1043: of the coarse grained energy spreading,
1044: and we get linear response.
1045:
1046:
1047:
1048: The existence of the adiabatic regime is obviously
1049: a quantum mechanical effect. If we take
1050: the formal limit $\hbar\rightarrow 0$
1051: the adiabaticity condition $\Gamma \ll \Delta$ breaks down.
1052: In fact the {\em proper} classical limit
1053: is {\em non-perturbative},
1054: because also the weaker condition
1055: $\Gamma \ll \Delta_b$ does not survive
1056: the $\hbar\rightarrow 0$ limit.
1057: For further details see \cite{vrn,dsp,crs,frc}.
1058: In the non-perturbative regime the
1059: quantum mechanical derivation of Kubo formula
1060: is not valid. Indeed we have demonstrated \cite{crs}
1061: the failure of Kubo formula in case of
1062: random-matrix models. But if the system
1063: has a classical limit, then Kubo formula
1064: still holds in the non-perturbative regime
1065: due to {\em semiclassical} (rather than quantum-mechanical) reasons.
1066:
1067:
1068:
1069: The discussion of {\em dissipation} assumes
1070: a generic situation such that
1071: the Schrodinger equation does not have
1072: a stationary solution. This means that
1073: driven one-dimensional systems are automatically
1074: excluded. Another non-generic possibility
1075: is to consider a special driving scheme,
1076: such as translation, rotation or dilation \cite{dil}.
1077: In such case the time dependent Hamiltonian
1078: ${\cal H}(\bm{x}(t))$ possesses a stationary solution
1079: (provided the ``velocity" $\dot{x}$
1080: is kept constant). Consequently we do
1081: not have a dissipation effect.
1082: In Section~11 we discuss the simplest
1083: example of pumping by translation, where
1084: the stationary adiabatic solution of Section~5
1085: is in fact exact, and no dissipation arises.
1086:
1087:
1088:
1089:
1090: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1091: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1092: \section{Application to pumping}
1093:
1094:
1095: So far we have discussed the response for driving in a very
1096: general way. From now on we focus on a system with a ring
1097: geometry as described in the Introduction, and illustrated
1098: in Fig.~1. The shape of the ring is controlled by some
1099: parameters $x_1$ and $x_2$, and $x_3$ is the magnetic flux.
1100: The generalized force $F^3$ which is conjugate
1101: to the flux is the current. The time integral over the
1102: current is the transported charge:
1103: %
1104: \begin{eqnarray}
1105: Q \ \ = \ \ \oint \langle F^{3} \rangle dt
1106: \end{eqnarray}
1107: %
1108: In fact a less misleading terminology is to talk
1109: about ``probability current" and
1110: "integrated probability current".
1111: From a purely mathematical point of view
1112: it is not important whether the transported
1113: particle has an electrical charge.
1114:
1115:
1116:
1117: Disregarding a possible persistent current contribution,
1118: the expression for the pumped charge is:
1119: %
1120: \begin{eqnarray} \label{e6}
1121: Q \ = \ -\left[ \oint \bmsf{G} \cdot d\bm{x} + \oint \bmsf{B} \wedge d\bm{x} \right]_{k=3}
1122: \end{eqnarray}
1123: %
1124: If we neglect the first term, which is associated
1125: with the dissipation effect, and average the second
1126: ("adiabatic") term over the flux, then we get
1127: %
1128: \begin{eqnarray} \label{e7}
1129: \overline{Q|_{\tbox{adiabatic}}} \ = \
1130: -\frac{1}{2\pi\hbar}\intt \bmsf{B} \cdot \vec{d\bm{x}} \wedge \vec{d\bm{x}}
1131: \end{eqnarray}
1132: %
1133: The integration should be taken over
1134: a cylinder of vertical height $2\pi\hbar$,
1135: and whose basis is determined by the projection of
1136: the pumping cycle onto the $(x_1,x_2)$ plane.
1137:
1138:
1139: We already pointed out that the Berry phase
1140: $(1/\hbar)\oint \bm{A}_n\cdot \vec{d\bm{x}}$
1141: is gauge invariant. Therefore from Stokes law it follows that
1142: $(1/\hbar)\intt \bmsf{B} \cdot \vec{d\bm{x}} \wedge \vec{d\bm{x}}$
1143: is independent of the surface, and therefore
1144: $(1/\hbar){\ointt} \bmsf{B} \cdot \vec{d\bm{x}} \wedge \vec{d\bm{x}} $
1145: with closed surface should be $2\pi\times$integer.
1146: Integrating over a cylinder, as in Eq.(\ref{e7}),
1147: is effectively like integrating over a closed surface
1148: (because of the $2\pi$~periodicity in the vertical
1149: direction). This means that the flux averaged $Q$
1150: of Eq.(\ref{e7}) has to be an integer.
1151:
1152:
1153:
1154:
1155: The common interest is in pumping cycles
1156: in the $\Phi=0$ plane. This means that the
1157: zero order conservative contribution to~$Q$,
1158: due to a persistent current, does not exist.
1159: Furthermore, from the reciprocity relations
1160: (see Appendix~B) it follows that
1161: $\fullG^{31}=- \fullG^{13}$,
1162: and $\fullG^{32}=- \fullG^{23}$,
1163: which should be contrasted with
1164: $\fullG^{12}=\fullG^{21}$.
1165: This means that a pumping cycle in the
1166: $\Phi=0$ plane is purely adiabatic:
1167: there is no dissipative contribution to~$Q$.
1168: Only the $\vec{\bm{B}}$ field (second term in Eq.(\ref{e6}))
1169: is relevant to the calculation of the pumped charge,
1170: and its vertical component ${B}^{12}$
1171: vanishes due to the time reversal symmetry.
1172:
1173:
1174:
1175: The absence of dissipative contribution for
1176: a cycle in the $\Phi=0$ plane, does not
1177: imply that dissipation is not an issue.
1178: The symmetric part of the conductance
1179: matrix $\symG^{ij}$ is in general non-zero,
1180: leading to an energy absorption
1181: rate which is proportional to $\dot{x}^2$.
1182: This implies that the energy absorption per cycle
1183: is proportional to $|\dot{x}|$. Therefore we
1184: are able to minimize the dissipation effect by making
1185: the pumping cycle very slow. Furthermore,
1186: if we get into the quantum-mechanical adiabatic regime,
1187: then $\symG^{ij}$ becomes extremely small, and then
1188: we can neglect the dissipation effect as long as
1189: quantum-mechanical adiabaticity can be trusted.
1190:
1191:
1192:
1193: Whenever the dissipation effect cannot be neglected,
1194: one should specify whether or how a {\em stationary operation} is achieved.
1195: In case of pumping in open system the stationary
1196: operation is implicitly guaranteed by having equilibrated reservoirs,
1197: where the extra energy is dissipated to infinity.
1198: In case of pumping in closed system the issue of stationary
1199: operation is more subtle:
1200: %
1201: In the adiabatic regime, to the extend that adiabaticity
1202: can be trusted, we have a stationary solution to the transport problem,
1203: as defined in Section~5.
1204: But outside of the adiabatic regime
1205: we have diffusion in energy space (Appendix~E)
1206: leading to a slow energy absorption (dissipation).
1207: %
1208: Thus a driven system is heated up gradually
1209: (though possibly very slowly). Strictly speaking
1210: a stationary operation is not achieved,
1211: unless the system is in (weak) thermal contact
1212: with some large bath.
1213: %
1214: Another way to reach a stationary operation,
1215: that does not involve an external bath,
1216: is by having an effectively bounded phase space.
1217: This is the case with the mixed phase space example
1218: which is discussed in Ref.\cite{ratchets}.
1219: There the stochastic-like motion takes place
1220: in a bounded chaotic region in phase space.
1221:
1222:
1223:
1224: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1225: \section{classical dissipative pumping}
1226:
1227: Before we discuss the quantum mechanical pumping,
1228: it is instructive to bring simple examples for
1229: {\em classical} pumping. In the following we consider
1230: one particle ($\bm{r}$)
1231: in a two dimensional ring as in Fig.1a.
1232:
1233: The first example is for classical {\em dissipative} pumping.
1234: The conductance $G=\fullG^{33}$
1235: can be calculated for this system \cite{wlf}
1236: leading to a mesoscopic variation of the Drude formula.
1237: The current is given by Ohm law $I=-G\times\dot{\Phi}$, where $-\dot{\Phi}$
1238: is the electro-motive-force.
1239:
1240: Consider now the following pumping cycle:
1241: Change the flux from $\Phi_1$ to $\Phi_2$,
1242: hence pumping charge \mbox{$Q = -G(1)\times (\Phi_2-\Phi_1)$}.
1243: Change the conductance from $G(1)$ to $G(2)$
1244: by modifying the shape of the ring.
1245: Change the flux from $\Phi_2$ back to $\Phi_1$,
1246: hence pumping charge \mbox{$Q(2) = -G(2) \times (\Phi_1-\Phi_2)$}.
1247: Consequently the net pumping is
1248: %
1249: \begin{eqnarray}
1250: Q \ = \ (G(2)-G(1)) \times (\Phi_2-\Phi_1)
1251: \end{eqnarray}
1252: %
1253: Thus we have used the dissipative part of the conductance
1254: matrix (first term in Eq.(\ref{e6})) in order to pump charge.
1255: %
1256: In the quantum mechanical version of this example
1257: extra care should be taken with respect to the
1258: zero order contribution of the persistent current.
1259:
1260: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1261: \section{classical adiabatic pumping}
1262:
1263:
1264: The second example is for classical {\em adiabatic} pumping.
1265: The idea is to trap the particle
1266: inside the ring by a potential well,
1267: and then to make a translation
1268: of the trap along a circle. The result of such a cycle
1269: is evidently $Q=1$. We would like to see
1270: how this trivial result emerges form the Kubo formula.
1271:
1272:
1273: Let $(\bm{r},\bm{p})$ be the canonical coordinate of the particle in the ring,
1274: while $(x_1,x_2)$ are the center coordinate of a trapping potential.
1275: The Hamiltonian is:
1276: %
1277: \begin{eqnarray}
1278: {\cal H}(\bm{r},\bm{p};\bm{x}(t)) &=&
1279: \frac{1}{2\mathsf{m}}
1280: \left[
1281: \mbf{p}_{\perp}^2 +
1282: \left(\mbf{p}_{\parallel}-\frac{1}{2\pi\sqrt{x_1^2+x_2^2}}\Phi(t)\right)
1283: \right] \nonumber \\ &+&
1284: U_{\tbox{trap}}(\mbf{r}_1{-}x_1(t),\mbf{r}_2{-}x_2(t))
1285: \end{eqnarray}
1286: %
1287: where $\mbf{p}_{\parallel}$ and $\mbf{p}_{\perp}$ are the components
1288: of the momentum along the ring and in the perpendicular (transverse) directions.
1289: The pumping is done simply by cycling the position of the trap.
1290: The translation of the trap is assumed to be along an inside circle
1291: of radius $R$,
1292: %
1293: \begin{eqnarray}
1294: \bm{x}(t) = (R\cos(\Omega t),R\sin(\Omega t),\Phi{=}\const)
1295: \end{eqnarray}
1296:
1297:
1298:
1299: In this problem the stationary solution of Section~5
1300: is an exact solution. Namely
1301: %
1302: \begin{eqnarray} \label{e_psit}
1303: |\psi(t)\rangle \ = \eexp{i\mathsf{m}\dot{\bm{x}}\cdot\bm{r}} \ |n(\bm{x}(t))\rangle
1304: \end{eqnarray}
1305: %
1306: where $|n(\bm{x})\rangle\mapsto \psi^{(n)}(\bm{r}-\bm{x})$ are
1307: the eigenfunctions of a particle in the trap.
1308: Eq.(\ref{e_psit}) is just Galilei transformation from
1309: the moving (trap) frame to the Laboratory frame.
1310:
1311:
1312: It is a-priori clear that in this
1313: problem the pumped charge per cycle is $Q=1$,
1314: irrespective of $\Phi$. Therefore the $\vec{\bm{B}}$ field
1315: must be
1316: %
1317: \begin{eqnarray} \label{e_52}
1318: \vec{\bm{B}} = -\frac{(x_1,x_2,0)}{2\pi (x_1^2+x_2^2)}
1319: \end{eqnarray}
1320: %
1321: This can be verified by calculation via Eq.(\ref{e29}).
1322: The singularity along the $x_3$ axis
1323: is not of quantum mechanical origin:
1324: It is not due to degeneracies, but rather
1325: due to the diverging current operator
1326: ($\partial {\cal H}/\partial x_3\propto 1/\sqrt{x_1^2+x_2^2}$).
1327:
1328:
1329:
1330:
1331:
1332:
1333:
1334:
1335: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1336: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1337: \section{quantum pumping}
1338:
1339:
1340: We turn now to the quantum mechanical case.
1341: Consider an adiabatic cycle that
1342: involves a particular energy level $n$.
1343: This level is assumed to have a degeneracy
1344: point at $(x_1^{(0)},x_2^{(0)},\Phi^{(0)})$.
1345: It follows that in fact there is
1346: a vertical chain of degeneracy points:
1347: %
1348: \begin{eqnarray} \label{e_chain}
1349: \mbox{chain} \ = \ (x_1^{(0)},x_2^{(0)},\Phi^{(0)}+2\pi\hbar\times\mbox{\small integer})
1350: \end{eqnarray}
1351: %
1352: These degeneracy points are important for the geometrical
1353: understanding of the $\vec{\bm{B}}$ field, as implied by Eq.(\ref{e29}).
1354: Every degeneracy point is like a monopole charge.
1355: The total flux that emerges from each monopole
1356: must be $2\pi\hbar\times$integer for a reason that was
1357: explained after Eq.(\ref{e7}).
1358: Thus the monopoles are quantized in units of~$\hbar/2$.
1359: %
1360: The $\vec{\bm{B}}$ field which is created (so to say)
1361: by a vertical chain of monopoles may have a different {\em near field}
1362: and {\em far field} behavior, which we discuss below.
1363:
1364:
1365:
1366: The far field region exists if the chains are
1367: well isolated. Later we explain that ``far"
1368: means $g_T \ll 1$, where $g_T$ is the Thouless conductance.
1369: The far field is obtained by regarding the chain as a smooth line.
1370: This leads {\em qualitatively} to the same field as in Eq.(\ref{e_52}).
1371: Consequently, for a ``large radius" pumping cycle
1372: in the $\Phi=0$ plane, we get $|Q|\approx1$.
1373: In the following we are interested in
1374: the deviation from exact quantization:
1375: If $\phi^{(0)}=0$ we expect to have
1376: $|Q| \ge 1$, while if $\phi^{(0)}=\pi$ we expect $|Q|\le 1$.
1377: Only for the $\phi$ averaged $Q$ of Eq.(\ref{e7})
1378: we get {\em exact quantization}.
1379:
1380:
1381:
1382: The deviation from $|Q|\approx 1$ is extremely large
1383: if we consider a tight pumping cycle around
1384: a $\phi^{(0)}=0$ degeneracy.
1385: After linear transformation of the shape parameters,
1386: the energy splitting $\Delta=E_n-E_m$ of the energy level~$n$
1387: from its neighboring (nearly degenerated) level~$m$
1388: can be written as
1389: %
1390: \begin{eqnarray}
1391: \Delta=
1392: ((x_1{-}x_1^{(0)})^2+(x_2{-}x_2^{(0)})^2+
1393: c^2(\phi{-}\phi^{(0)})^2)^{1/2}
1394: \end{eqnarray}
1395: %
1396: where $c$ is a constant. The monopole field is accordingly
1397: %
1398: \begin{eqnarray} \label{e9}
1399: \vec{\bm{B}} = \pm\frac{c}{2} \
1400: \frac{( x_1{-}x_1^{(0)}, x_2{-}x_2^{(0)}, x_3{-}x_3^{(0)}) }
1401: {((x_1{-}x_1^{(0)})^2+(x_2{-}x_2^{(0)})^2+
1402: (\frac{c}{\hbar})^2(x_3{-}x_3^{(0)})^2)^{3/2}}
1403: \end{eqnarray}
1404: %
1405: where the prefactor is determined by the requirement
1406: of having a single ($\hbar/2$) monopole charge.
1407: Assuming a pumping cycle of radius $R$ in the $\Phi=0$ plane
1408: we get from the second term of Eq.(\ref{e6})
1409: %
1410: \begin{eqnarray}
1411: Q \ = \ -\left[\oint \bmsf{B} \wedge d\bm{x} \right]_3
1412: \ = \ \mp \pi \sqrt{g_T}
1413: \end{eqnarray}
1414: %
1415: where
1416: %
1417: \begin{eqnarray}
1418: g_T \ = \
1419: \frac{1}{\Delta}
1420: \frac{\partial^2\Delta}{\partial\phi^2} \ = \
1421: \frac{c^2}{R^2}
1422: \end{eqnarray}
1423: %
1424: is a practical definition for the Thouless conductance
1425: in this context. It is used here simply as a measure
1426: for the sensitivity of an energy level to the magnetic flux $\Phi$.
1427:
1428: What we want to do in the next sections is to interpolate
1429: between the near field result, which is $Q={\cal O}(\sqrt{g_T})$,
1430: and the far field result, which is $Q={\cal O}(1)$.
1431: For this purpose it is convenient to consider
1432: a particular model that can be solved exactly.
1433:
1434:
1435: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1436: \section{The double barrier model}
1437:
1438:
1439: A simple example for quantum pumping is the double barrier model.
1440: An open geometry version of this model has been
1441: analyzed in \cite{barriers} using the $S$~matrix approach.
1442: The analogous closed geometry version is obtained
1443: by considering a one-dimensional ring with two delta barriers.
1444: As we are going to explain below,
1445: the pumping process in this model can be regarded
1446: as a particular example of an adiabatic transfer scheme:
1447: The electrons are adiabatically transfered from
1448: state to state one by one as in ``musical chair game".
1449:
1450:
1451: The two delta barriers version of the double barrier model
1452: is illustrated in Fig.2. The length of the ring is $L$,
1453: with periodic boundary conditions on $-(L/2)<r<(L/2)$.
1454: A dot region $|Q|<a/2$ is defined by the potential
1455: %
1456: \begin{eqnarray}
1457: U(r;c_1,c_2) =
1458: \frac{1}{c_1}\delta\left(r+\frac{a}{2}\right) +
1459: \frac{1}{c_2}\delta\left(r-\frac{a}{2}\right)
1460: \end{eqnarray}
1461: %
1462: It is assumed that $c_1$ and $c_2$ are small enough
1463: so one can classify the ring eigenstates into two
1464: categories: wire states, and dot states. The
1465: latter are those states that are localized in the
1466: dot region $|Q|<a/2$ in the limit of infinitely
1467: high barriers. We define the Fermi energy as
1468: the energy of the last occupied wire
1469: level in the limit of infinitely high barriers.
1470:
1471:
1472: The three parameters that we can control
1473: are the flux $x_3 = \Phi = \hbar\phi$,
1474: the bias $x_1=c_1-c_2$,
1475: and the dot potential $x_2=E_{\tbox{dot}}$
1476: which is related to $c_1+c_2$.
1477: The energy $E_{\tbox{dot}}$ correspond
1478: to the dot state which is closest to the Fermi
1479: energy $E_F$ from above.
1480: We assume that the other dot levels
1481: are much further away from the Fermi energy,
1482: and can be ignored. Note that another possible
1483: way to control the dot potential, is simply
1484: by changing a gate voltage: That means to
1485: assume that there is a control over the
1486: potential floor in the region $|Q|<a/2$.
1487:
1488:
1489:
1490: {\em The pumping cycle is assumed to be in the
1491: $\Phi=0$ plane, so there is no issue
1492: of conservative persistent current contribution}.
1493: The pumping cycle is defined as follows:
1494: We start with a positive bias ($x_1>0$)
1495: and lower the dot potential from a large $x_2>E_F$
1496: value to a small $x_2<E_F$ value.
1497: As a result, one electron is
1498: transfered via the {\em left} barrier into the
1499: dot region. Then we invert the bias
1500: ($x_1<0$) and raise back $x_2$. As a result
1501: the electron is transfered back into
1502: the wire via the {\em right} barrier.
1503:
1504:
1505:
1506: A closer look at the above scenario (Fig.2b)
1507: reveals the following:
1508: As we lower the dot potential across a wire
1509: level, an electron is adiabatically transfered
1510: once from left to right and then from right
1511: to left. As long as the bias is
1512: positive ($x_1>0$) the net charge being pumped
1513: is very small ($|Q| \ll 1$). Only the lowest wire
1514: level that participate in the pumping cycle
1515: carries $Q={\cal O}(1)$ net charge:
1516: It takes an electron from the left side,
1517: and after the bias reversal it emits it
1518: into the right side.
1519: %
1520: %
1521: %
1522: Thus the pumping process in this model can be regarded
1523: as a particular example \cite{avron} of an {\em adiabatic transfer scheme}:
1524: The electrons are adiabatically transfered from
1525: state to state, one by one, as in ``musical chair game".
1526:
1527:
1528:
1529: For a single occupied level the net $Q$ is the
1530: sum of charge transfer events that take place in
1531: four avoided crossings (two avoided crossings
1532: in case of the lowest level).
1533: For many particle occupation
1534: the total $Q$ is the sum over the net $Q$s
1535: which are carried by individual levels.
1536: For a dense zero temperature Fermi occupation
1537: the summation over all the net $Q$s is a telescopic sum,
1538: leaving non-canceling contributions only from the
1539: first and the last adiabatic crossings. The latter
1540: involve the last occupied level at the Fermi energy.
1541:
1542:
1543:
1544:
1545: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1546: \section{The three site lattice Hamiltonian}
1547:
1548:
1549: Rather than analyzing the two-delta-barriers version
1550: of the double barrier model, we consider below
1551: a simplified version that still contains the {\em same} essential
1552: ingredients. This is obtained by considering a three site lattice Hamiltonian.
1553: The advantage is obviously the possibility to make an exact analytical
1554: treatment that does not involve approximations.
1555:
1556:
1557: The middle site in the three site lattice Hamiltonian
1558: supports a single dot state,
1559: while the two other sites support
1560: two wire states. The Hamiltonian is
1561: %
1562: \begin{eqnarray}
1563: {\cal H} \mapsto \left(
1564: \matrix{
1565: 0 & c_1 & \eexp{i\phi} \cr
1566: c_1 & u & c_2 \cr
1567: \eexp{-i\phi} & c_2 & 0} \right)
1568: \end{eqnarray}
1569: %
1570: The three parameters are
1571: the bias $x_1=c_1-c_2$,
1572: the dot energy $x_2=u$,
1573: and the flux $x_3 = \Phi = \hbar\phi$.
1574: For presentation purpose we assume
1575: that \mbox{$0 < c_1,c_2 \ll 1$},
1576: and characterize the wire-dot coupling
1577: by the parameter $c=\sqrt{c_1c_2}$.
1578:
1579:
1580:
1581: The eigenstates are $E_n$.
1582: Disregarding the interaction
1583: with the dot ($c=0$) we have two
1584: wire states with \mbox{$E=\pm1$}.
1585: This implies degeneracies
1586: for $x_2=u=\mp1$.
1587: Once we switch on the coupling ($c>0$),
1588: the only possible degeneracies are between
1589: the even dot state and the odd wire state
1590: of the mirror symmetric Hamiltonian ($x_1=0$).
1591: The flux should be either integer
1592: (for degeneracy of the dot level with the lower wire level),
1593: or half integer
1594: (for degeneracy of the dot level with the upper wire level).
1595: Thus we have two vertical chains of degeneracies:
1596: %
1597: \begin{eqnarray} \nonumber
1598: \mbox{The negative chain}
1599: \ &=& \ (0,-1{+}c^2,2\pi\hbar\times\mbox{\small integer})
1600: \\ \nonumber
1601: \mbox{The positive chain}
1602: \ &=& \ (0,+1{-}c^2,\pi+2\pi\hbar\times\mbox{\small integer})
1603: \end{eqnarray}
1604:
1605:
1606:
1607:
1608:
1609: In order to calculate the $\vec{\bm{B}}$ field and pumped
1610: charge $Q$, we have to find the eigenvalues
1611: and the eigenvectors of the Hamiltonian matrix.
1612: The secular equation for the eigenvalues is
1613: %
1614: \begin{eqnarray} \nonumber
1615: E^3 - u E^2 - (1+c_1^2+c_2^2)E
1616: + u - 2 c_1 c_2 \cos(\phi) = 0
1617: \end{eqnarray}
1618: %
1619: Using the notations
1620: %
1621: \begin{eqnarray} \nonumber
1622: {\cal Q} &=& \frac{1}{9}u^2 + \frac{1}{3}(1+c_1^2+c_2^2)
1623: \\ \nonumber
1624: {\cal R} &=& \frac{1}{27}u^3
1625: + \frac{1}{6}(1+c_1^2+c_2^2)u - \frac{1}{2}u
1626: + c_1c_2 \cos(\phi)
1627: \\ \nonumber
1628: & & \cos(\theta)=\frac{{\cal R}}{\sqrt{{\cal Q}^3}}
1629: \end{eqnarray}
1630: %
1631: the roots of the above cubic equation are:
1632: %
1633: \begin{eqnarray}
1634: E_n = \frac{1}{3}u + 2\sqrt{{\cal Q}}\cos\left(\frac{1}{3} \theta + n\frac{2\pi}{3} \right)
1635: \end{eqnarray}
1636: %
1637: where $n=0,\pm1$.
1638: The corresponding eigenstates are:
1639: %
1640: \begin{eqnarray}
1641: |n(\bm{x})\rangle \mapsto
1642: \frac{1}{\sqrt{S}}
1643: \left(
1644: \matrix{
1645: c_2 \eexp{i\phi} + c_1 E_n \cr
1646: 1 - E_n^2 \cr
1647: c_1\eexp{-i\phi} + c_2 E_n} \right)
1648: \end{eqnarray}
1649: %
1650: where $S$ is the normalization, namely
1651: %
1652: \begin{eqnarray}
1653: S = (1 {-} E_n^2)^2
1654: + (c_1 {+} c_2 E_n)^2
1655: + (c_2 {+} c_1 E_n)^2
1656: \end{eqnarray}
1657: %
1658: For the calculation of the
1659: pumped charge in the next paragraph
1660: it is useful to notice that for $E=\pm 1$
1661: the normalization is $S=2(c_1 \pm c_2)^2$,
1662: while for $E=0$ the normalization is $S\approx 1$.
1663:
1664:
1665:
1666:
1667: After some algebra we find that
1668: the first component of the
1669: $\vec{\bm{B}}$ field in the
1670: $\Phi=0$ plane is
1671: %
1672: \begin{eqnarray}
1673: {\mbf{B}}^1 &=&
1674: -2\Im\left.\left\langle\frac{\partial}{\partial u} n(\bm{x}) \right|
1675: \frac{\partial}{\partial \phi} n(\bm{x}) \right\rangle
1676: \\ \ &=&
1677: -(c_1^2-c_2^2)
1678: \frac{1}{S^2}
1679: \frac{\partial S}{\partial u}
1680: \end{eqnarray}
1681: %
1682: Which is illustrated in Fig.3.
1683: From here it follow that if we
1684: keep constant bias, and change
1685: only $x_2=u$, then the pumped charge is:
1686: %
1687: \begin{eqnarray}
1688: Q = -\int {\mbf{B}}^1 dx_2 =
1689: \left.
1690: - (c_1^2-c_2^2)\frac{1}{S}
1691: \right|_{\tbox{initial}}^{\tbox{final}}
1692: \end{eqnarray}
1693: %
1694: For a planar ($\Phi=0$) pumping cycle around
1695: the negative vertical chain the main contribution
1696: to $Q$ comes from the two crossings
1697: of the $x_2 \approx -1$ line.
1698: Hence we get
1699: %
1700: \begin{eqnarray} \label{e_Qmodel}
1701: Q = \frac{c_1+c_2}{c_1-c_2} =
1702: \sqrt{1+2g_T}
1703: \end{eqnarray}
1704: %
1705: where the Thouless conductance
1706: in this context refers to the
1707: avoided crossing, and is defined as
1708: %
1709: \begin{eqnarray}
1710: g_T \equiv \left.\frac{1}{\Delta}
1711: \frac{\partial^2\Delta}
1712: {\partial\phi^2}\right|_{\phi=0}
1713: = \frac{2c_1c_2}{(c_1-c_2)^2}
1714: \end{eqnarray}
1715: %
1716: %
1717: %
1718: %
1719: %
1720: A similar calculation of the pumped charge for
1721: a planar cycle around the positive chain leads to
1722: %
1723: \begin{eqnarray}
1724: Q = -\frac{c_1-c_2}{c_1+c_2} =
1725: -\sqrt{1-2g_T}
1726: \end{eqnarray}
1727: %
1728: with $g_T = {2c_1c_2}/{(c_1 + c_2)^2}$.
1729: In both cases we have approximate quantization
1730: $Q=\pm 1 + {\cal O}(g_T)$ for $g_T \ll 1$,
1731: while for a tight cycle either $Q \rightarrow \infty$
1732: or $Q \rightarrow 0$ depending on which
1733: line of degeneracies is being encircled.
1734: If the pumping cycle encircles both chains then
1735: we get $Q = {4c_1c_2}/{(c_1^2-c_2^2)}$.
1736: In the latter case $Q={\cal O}(g_T)$ for $g_T \ll 1$,
1737: with no indication for quantization.
1738:
1739:
1740:
1741: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1742: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1743: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1744: \section{Summary and Discussion}
1745:
1746: We have shown how the Kubo formalism can
1747: be used in order to derive both classical and quantum
1748: mechanical results for the pumped charge $Q$ in
1749: a closed system. In this formulation the distinction
1750: between dissipative and non-dissipative contributions is manifest.
1751:
1752:
1753:
1754: Within the framework of the Kubo formalism (disregarding non-linear corrections)
1755: we have made a distinction between the following levels of treatment: \\
1756: %
1757: \begin{minipage}{\hsize}
1758: \vspace*{0.1cm}
1759: \begin{itemize}
1760: \setlength{\itemsep}{0cm}
1761: \item
1762: Strict adiabaticity \\ (outcome of zero order treatment)
1763: \item
1764: Adiabatic transport \\ (outcome of stationary first order treatment)
1765: \item
1766: Dissipation \\ (the result of first order transitions)
1767: \end{itemize}
1768: \vspace*{0.0cm}
1769: \end{minipage}
1770: %
1771: In the adiabatic regime one can assume
1772: a {\em stationary} solution to the adiabatic equation,
1773: which implies no dissipation effect.
1774: This leads to the picture of adiabatic transport,
1775: where the Berry phase is the outcome of a zero order treatment,
1776: while the ``geometric magnetism" of Eq.(\ref{e_30}) is the outcome
1777: of a first order treatment of the inter-level couplings.
1778:
1779:
1780:
1781: In some very special cases
1782: (translations, rotations and dilations)
1783: this assumption (of having a stationary solution)
1784: is in fact exact, but in generic circumstance this
1785: assumption is an approximation.
1786: Outside of the adiabatic regime the stationary solution
1787: cannot be trusted.
1788:
1789:
1790:
1791: Assuming quantized {\em chaotic dynamics} one
1792: argues that Fermi-golden-rule transitions
1793: between levels lead to (slow) diffusion
1794: in energy (Eq.(\ref{eD1})). This leads to the emergence
1795: of the dissipative part in the Kubo formula.
1796: %
1797: We have obtained an expression (Eq.(\ref{e_42}))
1798: for the energy scale $\Gamma \propto |\dot{x}|^{2/3}$
1799: that controls the dissipative effect.
1800: We have explained that the dissipative contribution
1801: to the Kubo formula is valid only in the regime
1802: $\Delta < \Gamma \ll \Delta_b$. Otherwise the dynamics
1803: is either of adiabatic nature ($\Gamma \ll \Delta$) or
1804: non-perturbative ($\Gamma>\Delta$).
1805:
1806:
1807:
1808: %%%%
1809:
1810:
1811: In order to calculate the pumped charge~$Q$ we have
1812: to perform a closed line integral over the conductance (Eq.(\ref{e6})).
1813: This may have in general both adiabatic and dissipative
1814: contributions. For the common pumping cycle in the
1815: $\Phi=0$ plane, only the adiabatic contribution exists.
1816: This follows from the reciprocity relations (Section~9).
1817: Still we have emphasized (without any contradiction)
1818: that in the same circumstances a dissipation effect
1819: typically accompanies the pumping process.
1820:
1821:
1822:
1823: The quantum adiabatic contribution
1824: to the pumping is determined
1825: by a line integral over a $\vec{\bm{B}}$ field
1826: which is created by {\em monopoles}.
1827: The monopoles, which are related
1828: to the degeneracies of the Hamiltonian,
1829: are located along vertical chains in $x$~space (Eq.(\ref{e_chain})).
1830: The 3~site model provides the simplest
1831: example for such vertical chains:
1832: By calculating the $\vec{\bm{B}}$ field which
1833: is created (so to say) by these chains,
1834: we were able to determine the charge
1835: which is pumped during a cycle (e.g. Eq.(\ref{e_Qmodel})).
1836:
1837:
1838:
1839: The (monopoles of the) vertical chains
1840: have {\em near field} regions (Eq.(\ref{e9})).
1841: If the chains are well isolated in $x$ space,
1842: then there are also {\em far field} regions.
1843: The far field regions are defined as those where
1844: the Thouless conductance is very small ($g_T \ll 1$).
1845: Pumping cycles that are contained in the
1846: far field region of a given chain lead to
1847: an approximately quantized pumping
1848: \mbox{$Q = \mbox{\small integer}+{\cal O}(g_T)$}.
1849: %
1850: %
1851: It is important to realize that the existence
1852: of far field regions in $x$~space is associated with having
1853: a low dimensional system far away from the classical limit.
1854: In a quantized chaotic system it is unlikely
1855: to have $g_T \ll 1$ along a pumping cycle.
1856: As we take the $\hbar\rightarrow0$ limit the
1857: vertical chains become very dense,
1858: and the far field regions disappear.
1859:
1860:
1861:
1862:
1863:
1864: In the subtle limiting case of open geometry
1865: we expect to get agreement with the
1866: $S$-matrix formula of B\"{u}ttiker Pr\'{e}tre
1867: and Thomas (BPT) \cite{BPT}.
1868: Using the notations of the present Paper
1869: the BPT formula for the current
1870: that comes out of (say) the right lead
1871: can be written as:
1872: %
1873: \begin{eqnarray}
1874: \fullG^{3j} = \frac{e}{2\pi i}
1875: \trc\left(P\frac{\partial S}{\partial x_j}
1876: S^{\dag}\right)
1877: \end{eqnarray}
1878: %
1879: where $P$ is the projector on the right lead channels.
1880: For $G^{33}$ the above reduces to the Landauer formula.
1881: The details regarding the relation between
1882: the Kubo formula and the BPT formula will be
1883: published in a separate paper \cite{pmo}. Here we just note that
1884: the derivation is based on a generalization
1885: of the Fisher-Lee approach \cite{datta,fisher,stone}.
1886:
1887:
1888:
1889: Finally it is important to remember that the theory of
1890: driven systems is the corner stone for the analysis
1891: of interaction between ``slow" and ``fast" degrees of freedom.
1892: Assume that that the $x_j$ are in fact dynamical variable,
1893: and that the conjugate momenta are $p_j$.
1894: The standard textbook example is the study
1895: of diatomic molecules. In such case $x_j$ are the locations of the nuclei.
1896: The total Hamiltonian is assumed to be of the general form
1897: %
1898: \begin{eqnarray}
1899: {\cal H}_{\tbox{total}}=\frac{1}{2M}\sum_j p_j^2 \ + \ {\cal H}(\bm{x})
1900: \end{eqnarray}
1901: %
1902: where ${\cal H}$ is the Hamiltonian of the
1903: "fast" degrees of freedom (in the context
1904: of molecular physics these are the electrons).
1905: Rather than using the standard basis,
1906: one can use the Born-Oppenheimer basis
1907: $|x,n(\bm{x})\rangle = |x\rangle \otimes |n(\bm{x})\rangle$.
1908: Then the Hamiltonian can be written as
1909: %
1910: \begin{eqnarray} \nonumber
1911: {\cal H}_{\tbox{total}}=
1912: \frac{1}{2M}\sum_j(p_j-\mbf{A}^j_{nm}(\bm{x}))^2
1913: \ + \ \delta_{nm}E_n(\bm{x})
1914: \end{eqnarray}
1915: %
1916: where the interaction term is consistent with Eq.(\ref{e_23}).
1917: Thus it is evident that the theory of driven systems
1918: is a special limit of this problem,
1919: which is obtained if we treat the $x_j$ as classical variables.
1920:
1921:
1922: %\newpage
1923: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1924: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1925:
1926: \ \\ \ \\
1927: {\bf Acknowledgments:}
1928: It is my pleasure to thank Yshai Avishai (Ben-Gurion University), Yosi Avron (Technion),
1929: Thomas Dittrich (Colombia), Shmuel Fishman (Technion), Tsampikos Kottos (Gottingen),
1930: and Holger Schantz (Gottingen) for useful discussions.
1931: This research was supported by the Israel Science Foundation (grant No.11/02),
1932: and by a grant from the GIF, the German-Israeli Foundation for Scientific
1933: Research and Development.
1934:
1935:
1936:
1937: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1938: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1939: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1940: \appendix
1941:
1942:
1943: \clearpage
1944: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1945: \section{The Kubo formula: Standard Derivation}
1946:
1947: In this Appendix we present an elementary textbook-style
1948: derivation of the Kubo formula. For notational simplicity
1949: we write the Hamiltonian as \mbox{${\cal H}=\bmsf{H}_0-f(t)\bmsf{V}$}.
1950: %
1951: It is assumed that the system, in the absence
1952: of driving, is prepared in a stationary state $\rho_0$.
1953: In the presence of driving we look for a first order solution
1954: $\rho(t) = \rho_0 +\tilde{\rho}(t)$.
1955: The equation for $\tilde{\rho}(t)$ is:
1956: %
1957: \begin{eqnarray}
1958: \frac{\partial \tilde{\rho}(t)}{\partial t} \approx
1959: -i[\bmsf{H}_0, \tilde{\rho}(t)] + if(t)[\bmsf{V},\rho_0]
1960: \end{eqnarray}
1961: %
1962: This equation can be re-written as
1963: %
1964: \begin{eqnarray} \nonumber
1965: \frac{\partial}{\partial t}
1966: (\bmsf{U}_0(t)^{-1} \tilde{\rho}(t) \bmsf{U}_0(t))
1967: \approx
1968: if(t)[ \bmsf{U}_0(t)^{-1}\bmsf{V}\bmsf{U}_0(t),\rho_0]
1969: \end{eqnarray}
1970: %
1971: where $\bmsf{U}_0(t)$ is the evolution operator
1972: which is generated by $\bmsf{H}_0$.
1973: The solution of the latter equation is
1974: %
1975: \begin{eqnarray}
1976: \tilde{\rho}(t)
1977: \ \approx \
1978: \int^{t} i [\bmsf{V}(-(t{-}t')), \rho_0] \ f(t')dt'
1979: \end{eqnarray}
1980: %
1981: where we use the usual definition
1982: of the ``interaction picture" operator
1983: $\bmsf{V}(\tau)=\bmsf{U}_0(\tau)^{-1}\bmsf{V}\bmsf{U}_0(\tau)$.
1984:
1985:
1986:
1987:
1988: Consider now the time dependence of the expectation value
1989: \mbox{$\langle \bmsf{F} \rangle_t = \trc(\bmsf{F}\rho(t))$}
1990: of an observable. Disregarding the zero order contribution,
1991: the first order expression is
1992: %
1993: \begin{eqnarray} \nonumber
1994: \langle \bmsf{F} \rangle_t
1995: \ &\approx& \
1996: \int^{t} i \trc\left(\bmsf{F} [\bmsf{V}(-(t{-}t')), \rho_0]\right) \ f(t')dt'
1997: \\ \nonumber
1998: &=& \
1999: \int^{t} \alpha(t-t') \ f(t')dt'
2000: \end{eqnarray}
2001: %
2002: where the response kernel $\alpha(\tau)$ is
2003: defined for $\tau>0$ as
2004: %
2005: \begin{eqnarray}
2006: \alpha(\tau) &=& i\ \trc\left(\bmsf{F} [\bmsf{V}(-\tau), \rho_0]\right)
2007: \nonumber \\
2008: \ &=& i\ \trc\left([\bmsf{F}, \bmsf{V}(-\tau)] \rho_0\right)
2009: \nonumber \\
2010: \ &=& i \langle [\bmsf{F}, \bmsf{V}(-\tau)] \rangle
2011: \nonumber \\
2012: \ &=& i \langle [\bmsf{F}(\tau), \bmsf{V}] \rangle
2013: \end{eqnarray}
2014: %
2015: We have used above the cyclic property of the trace operation;
2016: the stationarity $\bmsf{U}_0\rho_0\bmsf{U}_0^{-1}=\rho_0$ of
2017: the unperturbed state; and the definition
2018: $\bmsf{F}(\tau)=\bmsf{U}_0(\tau)^{-1}\bmsf{F}\bmsf{U}_0(\tau)$.
2019:
2020:
2021:
2022:
2023:
2024: \newpage
2025: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2026: \section{Remarks regarding the generalized susceptibility}
2027:
2028: In this appendix we would like to further illuminate the relation
2029: between the generalized susceptibility and the conductance matrix.
2030: %
2031: The generalized susceptibility $\chi^{kj}(\omega)$ is
2032: the Fourier transform of the causal response kernel $\alpha^{kj}(\tau)$.
2033: Therefore it is an analytic function in the upper
2034: half of the complex $\omega$ plan, whose real and imaginary
2035: parts are related by Hilbert transforms (Kramers-Kronig relations):
2036: %
2037: \begin{eqnarray}
2038: \chi_0^{kj}(\omega) \equiv \re [\chi^{kj}(\omega)] =
2039: \int_{-\infty}^{\infty} \frac{\im[\chi^{kj}(\omega')]}{\omega'-\omega} \ \frac{d\omega'}{\pi}
2040: \end{eqnarray}
2041: %
2042: The imaginary part of $\chi^{kj}(\omega)$
2043: is the sine transforms of $\alpha^{kj}(\tau)$,
2044: and therefore it is proportional to $\omega$
2045: for small frequencies.
2046: Consequently it is convenient to write the
2047: Fourier transformed version of Eq.(\ref{e_3}) as
2048: %
2049: \begin{eqnarray} \label{e3}
2050: [\langle F^k \rangle]_{\omega} \ = \ \sum_j
2051: \chi_0^{kj}(\omega) [x_j]_{\omega}
2052: -\mu^{kj}(\omega) [\dot{x}_j]_{\omega}
2053: \end{eqnarray}
2054: %
2055: where the dissipation coefficient is defined as
2056: %
2057: \begin{eqnarray}
2058: \mu^{kj}(\omega) =
2059: \frac{\im[\chi^{kj}(\omega)]}{\omega} =
2060: \int_0^{\infty} \alpha^{kj}(\tau) \frac{\sin(\omega\tau)}{\omega}d\tau
2061: \end{eqnarray}
2062: %
2063: In this paper we ignore the first term in Eq.(\ref{e3})
2064: which signify the non-dissipative in-phase response.
2065: Rather we put the emphasis on the ``DC limit" ($\omega\rightarrow 0$)
2066: of the second term. Thus the conductance
2067: matrix \mbox{$\fullG^{kj}=\mu^{kj}(\omega \rightarrow 0)$}
2068: is just a synonym for the term ``dissipation coefficient".
2069: However, ``conductance" is a better (less misleading) terminology:
2070: it does not have the (wrong) connotation
2071: of being specifically associated with dissipation,
2072: and consequently it is less confusing to say that
2073: it contains a (non-dissipative) adiabatic component.
2074:
2075:
2076:
2077: For systems where time reversal symmetry is broken
2078: due to the presence of a magnetic field ${\cal B}$,
2079: the response kernel, and consequently the generalized susceptibility
2080: and the conductance matrix satisfies the Onsager reciprocity relations
2081: %
2082: \begin{eqnarray}
2083: \alpha^{ij}(\tau, -{\cal B}) \ &=& \ [\pm] \ \alpha^{ji}(\tau, {\cal B}) \\
2084: \chi^{ij}(\omega, -{\cal B}) \ &=& \ [\pm] \ \chi^{ji}(\omega, {\cal B}) \\
2085: \fullG^{ij} (-{\cal B}) \ &=& \ [\pm] \ \fullG^{ji}({\cal B})
2086: \end{eqnarray}
2087: %
2088: where the plus (minus) applies if the signs of $F^i$ and $F^j$
2089: transform (not) in the same way under time reversal.
2090: These reciprocity relations follow from
2091: the Kubo formula (Eq.(\ref{e_kubo}),
2092: using
2093: $K^{ij}(-\tau,-{\cal B}) = - [\pm] K^{ij}(\tau,{\cal B}) $,
2094: together with the trivial identity
2095: $K^{ij}(-\tau,{\cal B}) = - K^{ji}(\tau,{\cal B}) $.
2096: In Section.~9 we discuss the implications of the reciprocity relations
2097: in the context of pumping.
2098:
2099:
2100:
2101: \clearpage
2102: \onecolumngrid
2103: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2104: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2105: \section{Expressions for $\bmsf{B}$ and $\fullG$}
2106:
2107:
2108: The functions $C^{ij}(\tau)$ and $K^{ij}(\tau)$ are
2109: the expectation values of hermitian operators.
2110: Therefore they are real functions. It follows that the
2111: real part of their Fourier transform
2112: is a symmetric function with respect to $\omega$,
2113: while the imaginary part of their Fourier transform
2114: is anti symmetric with respect to $\omega$.
2115: By definition they satisfy $C^{ij}(\tau) = C^{ji}(-\tau)$
2116: and $K^{ij}(\tau) = -K^{ji}(-\tau) $.
2117: %
2118: It is convenient to regard them as the real and imaginary parts
2119: of one complex function $\Phi^{ij}(\tau)$. Namely,
2120: %
2121: \begin{eqnarray}
2122: \Phi^{ij}(\tau) &=& \langle F^i(\tau)F^j(0)\rangle \ \ = \ \ C^{ij}(\tau)-i\frac{\hbar}{2}K^{ij}(\tau) \\
2123: C^{ij}(\tau) &=& \frac{1}{2}\left(\Phi^{ij}(\tau) + \Phi^{ji}(-\tau)\right) \\
2124: K^{ij}(\tau) &=& \frac{i}{\hbar} \left(\Phi^{ij}(\tau) - \Phi^{ji}(-\tau)\right)
2125: \end{eqnarray}
2126:
2127:
2128:
2129:
2130:
2131: It is possible to express the decomposition $\fullG^{ij}=\symG^{ij}+\mbf{B}^{ij} $
2132: in terms of $\tilde{K}^{ij}(\omega)$. Using the definition Eq.(\ref{e_33}) we get:
2133: %
2134: \begin{eqnarray} \label{eB4}
2135: \fullG^{ij}
2136: \ \ = \ \
2137: \int_0^{\infty}K^{ij}(\tau)\tau d\tau
2138: \ \ = \ \
2139: {-}\int_{-\infty}^{\infty}\frac{\re[\tilde{K}^{ij}(\omega)]}{\omega^2}
2140: \frac{d\omega}{2\pi} +
2141: \left[\frac{1}{2}\frac{\im[\tilde{K}^{ij}(\omega)]}{\omega}\right]_{\omega{=}0}
2142: \end{eqnarray}
2143: %
2144: %
2145: %
2146: %
2147: The first term is antisymmetric with respect to its indexes,
2148: and is identified as $\mbf{B}^{ij}$.
2149: The second term is symmetric with respect to its indexes,
2150: and is identified as $\symG^{ij}$.
2151: The last step in the above derivation involves
2152: the following identity that hold for any real function $f(\tau)$
2153: %
2154: \begin{eqnarray}
2155: \int_0^{\infty}f(\tau)\tau d\tau
2156: =
2157: \int_{-\infty}^{\infty}\frac{d\omega}{2\pi}\tilde{f}(\omega)
2158: \int_0^{\infty}\eexp{-i\omega\tau}\tau d\tau
2159: =
2160: \int_{-\infty}^{\infty}\frac{d\omega}{2\pi}
2161: \tilde{f}(\omega)
2162: \left(-\frac{1}{\omega^2}+i\pi\delta'(\omega)\right)
2163: =
2164: \nonumber \\
2165: =
2166: \int_{-\infty}^{\infty}\frac{d\omega}{2\pi}
2167: \left(-\frac{\re[\tilde{f}(\omega)]}{\omega^2}
2168: -\pi\im[\tilde{f}(\omega)]\delta'(\omega)\right)
2169: =
2170: {-}\int_{-\infty}^{\infty}\frac{\re[\tilde{f}(\omega)]}{\omega^2}
2171: \frac{d\omega}{2\pi} +
2172: \left[\frac{1}{2}\frac{\im[\tilde{f}(\omega)]}{\omega}\right]_{\omega{=}0}
2173: \end{eqnarray}
2174: %
2175: Note that $\im[\tilde{f}(\omega)]$ is the sine transform of $f(\tau)$,
2176: and therefore it is proportional to $\omega$ is the limit of small frequencies.
2177:
2178:
2179:
2180: It is of practical value to re-derive Eq.(\ref{eB4})
2181: by writing $\Phi^{ij}(\tau)$ using the energies $E_n$
2182: and the matrix elements~$F^i_{nm}$.
2183: Then we can get from it straightforwardly (using the definitions)
2184: all the other expressions. Namely:
2185: %
2186: \begin{eqnarray}
2187: \Phi^{ij}(\tau) \ &=& \
2188: \sum_n f(E_n) \sum_{m} F^i_{nm}F^j_{mn}
2189: \exp\left(-i\frac{E_m{-}E_n}{\hbar}t \right)
2190: \\
2191: \tilde{\Phi}^{ij}(\omega) \ &=& \
2192: \sum_n f(E_n) \sum_{m} F^i_{nm}F^j_{mn}
2193: \ 2\pi\delta\left(\omega-\frac{E_m{-}E_n}{\hbar} \right)
2194: \\
2195: \chi^{ij}(\omega) \ &=& \ \sum_{n,m} f(E_n)\left(
2196: \frac{-F^i_{nm}F^j_{mn}}{\hbar\omega{-}(E_m{-}E_n){+}i0}
2197: +\frac{F^j_{nm}F^i_{mn}}{\hbar\omega{+}(E_m{-}E_n){+}i0}\right)
2198: \\
2199: \symG^{ij} \ &=& \ -2\pi\hbar \sum_n f(E_n) \sum_{m(\ne n)}
2200: \re\left[F^i_{nm}F^j_{mn}\right]
2201: \ \delta'(E_m-E_n)
2202: \\ \label{eB10}
2203: \mbf{B}^{ij} \ &=& \ 2\hbar \sum_n f(E_n) \sum_{m(\ne n)}
2204: \frac{\im\left[
2205: F^i_{nm}F^j_{mn}\right]}
2206: {(E_m-E_n)^2}
2207: \end{eqnarray}
2208: %
2209: %
2210: One observes that the expression for $\mbf{B}^{ij}$ coincides
2211: with the adiabatic transport result Eq.(\ref{e_30}).
2212: Alternatively this identification can be obtained by expressing
2213: the sum in Eq.(\ref{eB10}) as an integral, getting form it
2214: the first term in Eq.(\ref{eB4}):
2215: %
2216: \begin{eqnarray}
2217: \mbf{B}^{ij} \ \ = \ \
2218: \frac{2}{\hbar}\int_{-\infty}^{\infty}
2219: \frac{\im[\tilde{\Phi}^{ij}(\omega)]}{\omega^2}\frac{d\omega}{2\pi}
2220: \ \ = \ \
2221: \frac{2}{\hbar}\int_{-\infty}^{\infty}
2222: \frac{\im[\tilde{C}^{ij}(\omega)]-\frac{\hbar}{2}\re[ \tilde{K}^{ij}(\omega) ]}
2223: {\omega^2}\frac{d\omega}{2\pi}
2224: \ \ = \ \
2225: -\int_{-\infty}^{\infty}
2226: \frac{\re[\tilde{K}^{ij}(\omega)]}{\omega^2}\frac{d\omega}{2\pi}
2227: \end{eqnarray}
2228:
2229:
2230:
2231: \clearpage
2232: \onecolumngrid
2233: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2234: \section{Expressing $\tilde{K}(\omega)$ using $\tilde{C}(\omega)$}
2235:
2236: We can use the following manipulation in order
2237: to relate $\tilde{K}^{ij}(\omega)$ to $\tilde{C}^{ij}(\omega)$,
2238: %
2239: \begin{eqnarray} \label{eC1}
2240: \tilde{K}^{ij}(\omega)&=& \sum_n f(E_n) \ \tilde{K}^{ij}_n(\omega)
2241: \\ \nonumber &=&
2242: \frac{i}{\hbar} 2\pi \sum_{nm}f(E_n)
2243: (F^i_{nm}F^j_{mn}\delta(\omega+\omega_{nm})-F^j_{nm}F^i_{mn}\delta(\omega-\omega_{nm}))
2244: \\ \nonumber &=&
2245: \frac{i}{\hbar} 2\pi \sum_{nm}f(E_m)
2246: (-F^i_{nm}F^j_{mn}\delta(\omega+\omega_{nm})+F^j_{nm}F^i_{mn}\delta(\omega-\omega_{nm}))
2247: \\ \nonumber &=&
2248: \frac{i}{\hbar} 2\pi \sum_{nm} \frac{f(E_n)-f(E_m)}{2}
2249: (F^i_{nm}F^j_{mn}\delta(\omega+\omega_{nm})-F^j_{nm}F^i_{mn}\delta(\omega-\omega_{nm}))
2250: \\ \nonumber &=&
2251: -i \omega \pi \sum_{nm}\frac{f(E_n)-f(E_m)}{E_n-E_m}
2252: (F^i_{nm}F^j_{mn}\delta(\omega+\omega_{nm})+F^j_{nm}F^i_{mn}\delta(\omega-\omega_{nm}))
2253: \\ \nonumber &=&
2254: -i\omega\sum_{n} f'(E_n) \ C^{ij}_n(\omega)
2255: \end{eqnarray}
2256: %
2257: where we use the notation $\omega_{nm}=(E_n-E_m)/\hbar$.
2258: The third line differs from the second line by permutation of the dummy
2259: summation indexes, while the fourth line is the sum of the second
2260: and the third lines divided by 2. In the last equality we assume small $\omega$.
2261: If the levels are very dense, then we can replace the summation by integration,
2262: leading to the relation:
2263: %
2264: \begin{eqnarray}
2265: \int g(E)dE \ f(E) \ \tilde{K}^{ij}_{E}(\omega)
2266: \ \ = \ \
2267: -i \omega \int g(E)dE \ f'(E) \ \tilde{C}^{ij}_{E}(\omega)
2268: \end{eqnarray}
2269: %
2270: where $\tilde{K}^{ij}_{E}(\omega)$ and $\tilde{C}^{ij}_{E}(\omega)$
2271: are microcanonically smoothed functions.
2272: Since this equality hold for any smoothed $f(E)$, it follows
2273: that the following relation holds (in the limit $\omega\rightarrow0$):
2274: %
2275: \begin{eqnarray}
2276: \tilde{K}^{ij}_E(\omega)
2277: \ = \
2278: i\omega \frac{1}{g(E)}\frac{d}{dE}\left[g(E)C^{ij}_E(\omega)\right]
2279: \end{eqnarray}
2280: %
2281: If we do not assume small $\omega$, but instead assume canonical state,
2282: then a variation on the last steps in Eq.(\ref{eC1}),
2283: using the fact that $(f(E_n){-}f(E_m))/ (f(E_n){+}f(E_m))= \tanh((E_n{-}E_m)/(2T))$
2284: is an odd function, leads to the relation
2285: %
2286: \begin{eqnarray}
2287: \tilde{K}^{ij}_T(\omega) \ = \ i\omega \times
2288: \frac{1}{\hbar\omega}\tanh\left(\frac{\hbar\omega}{2T}\right) \ C^{ij}_T(\omega)
2289: \end{eqnarray}
2290: %
2291: Upon substitution of the above expressions in the Kubo formula for $\symG^{ij}$,
2292: one obtains the Fluctuation-Dissipation relation.
2293:
2294:
2295:
2296: %\clearpage
2297:
2298: \ \\ \ \\
2299: \line(1,0){490}
2300: \ \\ \ \\
2301:
2302: \twocolumngrid
2303: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2304: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2305: \section{The Kubo formula and the diffusion in energy space}
2306:
2307: The illuminating derivation of Eq.(\ref{e43}) is based on
2308: the observation that energy absorption is related to
2309: having diffusion in energy space. Let us assume that
2310: the probability distribution $\rho(E)=g(E)f(E)$ of the energy
2311: satisfies the following diffusion equation:
2312: %
2313: \begin{eqnarray} \label{eD1}
2314: \frac{\partial \rho}{\partial t} \ = \
2315: \frac{\partial}{\partial E}
2316: \left(g(E)D_E \frac{\partial}{\partial E}
2317: \left(\frac{1}{g(E)}\rho\right)\right)
2318: \end{eqnarray}
2319: %
2320: The energy of the system is
2321: $\langle {\cal H} \rangle=\int E \rho(E)dE$.
2322: It follows that the rate of energy absorption is
2323: %
2324: \begin{eqnarray}
2325: \frac{d}{dt}\langle {\cal H} \rangle
2326: = - \int_0^{\infty} dE \ g(E) \ D_E
2327: \ \frac{\partial}{\partial E}
2328: \left(\frac{\rho(E)}{g(E)}\right)
2329: \end{eqnarray}
2330: %
2331: For a microcanonical preparation we get
2332: %
2333: \begin{eqnarray} \label{e48}
2334: \frac{d}{dt}\langle {\cal H} \rangle
2335: \ = \
2336: \frac{1}{g(E)}
2337: \frac{d}{dE}\left[g(E) \ D_E \right]
2338: \end{eqnarray}
2339: %
2340: This diffusion-dissipation relation reduces
2341: immediately to the fluctuation-dissipation relation
2342: if we assume that the diffusion in energy
2343: space due to the driving is given by
2344: %
2345: \begin{eqnarray} \label{e49}
2346: D_E \ = \
2347: \frac{1}{2}\sum_{ij}
2348: \tilde{C}^{ij}_E(\omega{\rightarrow}0)
2349: \ \dot{x}_i \dot{x}_j
2350: \end{eqnarray}
2351: %
2352: Thus it is clear that a theory for linear response
2353: should establish that there is a diffusion process in energy
2354: space due to the driving, and that the diffusion
2355: coefficient is given by Eq.(\ref{e49}).
2356: More importantly, this approach also allows
2357: treating cases where the expression
2358: for $D_E$ is non-perturbative,
2359: while the diffusion-dissipation relation Eq.(\ref{e48}) still holds!
2360:
2361:
2362: A full exposition (and further reference) for this route
2363: of derivation can be found in \cite{vrn,dsp,crs,frc}.
2364: Here we shall give just the classical derivation of Eq.(\ref{e49}),
2365: which is extremely simple. We start with the identity
2366: %
2367: \begin{eqnarray}
2368: \frac{d}{dt}\langle {\cal H} \rangle
2369: \ = \ \left\langle \frac{\partial {\cal H}}{\partial t} \right\rangle \ =
2370: \ -\sum_k \dot{x}_k F^k(t)
2371: \end{eqnarray}
2372: %
2373: Assuming (for presentation purpose) that
2374: the rates $\dot{x}_k$ are constant numbers,
2375: it follows that energy changes are related
2376: to the fluctuating $F^k(t)$ as follows:
2377: %
2378: \begin{eqnarray}
2379: \delta E \ = \ \langle {\cal H} \rangle_t - \langle {\cal H} \rangle_0 \ =
2380: \ -\sum_k \dot{x}_k \int_0^t F^k(t')dt'
2381: \end{eqnarray}
2382: %
2383: Squaring this expression, and performing
2384: microcanonical averaging over initial conditions we obtain:
2385: %
2386: \begin{eqnarray}
2387: \delta E^2(t) \ = \ \sum_{ij} \dot{x}_i\dot{x}_j \int_0^t\int_0^t
2388: C^{ij}_E(t''-t') dt' dt''
2389: \end{eqnarray}
2390: %
2391: where $C^{ij}_E(t''-t')= \langle F^i(t') F^j(t'') \rangle$ is the
2392: correlation function. For very short times this equation
2393: implies ``ballistic'' spreading ($\delta E^2 \propto t^2$)
2394: while on {\em intermediate time scales} it leads
2395: to diffusive spreading \mbox{$\delta E^2(t) =2D_E t$}, where
2396: %
2397: \begin{eqnarray}
2398: D_E \ = \ \frac{1}{2} \sum_{ij} \dot{x}_i\dot{x}_j \int_{-\infty}^{\infty}
2399: C^{ij}_E(\tau) d\tau
2400: \end{eqnarray}
2401: %
2402: The latter result assumes a short correlation time.
2403: This is also the reason that the integration
2404: over $\tau$ can be extended form
2405: $-\infty$ to $+\infty$. Hence we get Eq.(\ref{e49}).
2406: We note that for long times the systems deviates
2407: significantly from the initial microcanonical preparation.
2408: Hence, for long times, one should justify the
2409: use of the diffusion equation (\ref{eD1}).
2410: This leads to the classical slowness condition
2411: which is discussed in Ref.\cite{frc}.
2412:
2413:
2414: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2415: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2416: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2417: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2418: \begin{thebibliography}{99}
2419:
2420:
2421: % general
2422:
2423: \bibitem{landau}
2424: L.D. Landau and E.M. Lifshitz,
2425: {\em Statistical physics}, (Buttrworth Heinemann 2000).
2426:
2427: \bibitem{imry}
2428: Y. Imry, {\em Introduction to Mesoscopic Physics}
2429: (Oxford Univ. Press 1997), and references therein.
2430:
2431: \bibitem{datta}
2432: S. Datta, {\em Electronic Transport in Mesoscopic Systems}
2433: (Cambridge University Press 1995).
2434:
2435: \bibitem{marcus_rev}
2436: L.P. Kouwenhoven et al,
2437: Proc. of Advanced Study Inst. on Mesoscopic
2438: Electron Transport, edited by L.L. Sohn,
2439: L.P. Kouwenhoven and G. Schon (Kluwer 1997).
2440:
2441:
2442: % pumping
2443:
2444: \bibitem{thouless}
2445: D. J. Thouless, Phys. Rev. {\bf B27}, 6083 (1983).
2446:
2447: \bibitem{BPT}
2448: M. B\"{u}ttiker et al, Z. Phys. {\bf B94}, 133 (1994). \
2449: P. W. Brouwer, Phys. Rev. {\bf B58}, R10135 (1998). \
2450: J. E. Avron et al, Phys. Rev. B {\bf 62}, R10 618 (2000).
2451:
2452: \bibitem{aleiner}
2453: T. A. Shutenko, I. L. Aleiner and B. L. Altshuler, Phys. Rev. {\bf B61}, 10366 (2000).
2454:
2455: \bibitem{barriers}
2456: Y. Levinson, O. Entin-Wohlman, and P. Wolfle, cond-mat/0010494. \
2457: M. Blaauboer and E.J. Heller, Phys. Rev. B 64, 241301(R) (2001).
2458:
2459: \bibitem{ora}
2460: O. Entin-Wohlman, A. Aharony, and Y. Levinson Phys. Rev. B 65, 195411 (2002).
2461:
2462: \bibitem{pmo}
2463: D. Cohen, cond-mat/0304678.
2464:
2465: \bibitem{ratch} P.Reimann, Phys. Rep. {\bf 361}, 57 (2002). \
2466: Special issue, Appl. Phys. A {\bf 75} (2002). \
2467: P.Reimann, M. Grifoni, and P.Hanggi, Phys. Rev. Lett. {\bf 79}, 10 (1997).
2468:
2469: \bibitem{ratchets}
2470: H. Schanz, M.F. Otto, R. Ketzmerick, and T. Dittrich, Phys. Rev. Lett. {\bf 87}, 070601 (2001).
2471:
2472:
2473:
2474: % adiabatic equation
2475:
2476: \bibitem{berry}
2477: M.V. Berry, Proc. R. Soc. Lond. A {\bf 392}, 45 (1984).
2478:
2479:
2480: \bibitem{avron}
2481: J. E. Avron and L. Sadun, Phys. Rev. Lett. {\bf 62}, 3082 (1989). \
2482: J. E. Avron and L. Sadun, Ann. Phys. {\bf 206}, 440 (1991). \
2483: J.E. Avron, A. Raveh, and B. Zur Rev. Mod. Phys. {\bf 60}, 873 (1988).
2484:
2485: \bibitem{robbins}
2486: J.M. Robbins and M.V. Berry, J. Phys. A {\bf 25}, L961 (1992). \
2487: M.V. Berry and J.M. Robbins, Proc. R. Soc. Lond. A {\bf 442}, 659 (1993). \
2488: M.V. Berry and E.C. Sinclair, J. Phys. A {\bf 30}, 2853 (1997).
2489:
2490:
2491:
2492: % regimes, dissipation
2493:
2494: \bibitem{vrn}
2495: D. Cohen in {\em New directions in quantum chaos},
2496: Proceedings of the International School of Physics ``Enrico Fermi", Course CXLIII,
2497: Edited by G. Casati, I. Guarneri and U. Smilansky,
2498: (IOS Press, Amsterdam 2000).
2499:
2500: \bibitem{dsp}
2501: D. Cohen in {\em Dynamics of Dissipation},
2502: Proceedings of the 38th Karpacz Winter School of Theoretical Physics,
2503: Edited by P. Garbaczewski and R. Olkiewicz,
2504: (Springer, 2002)
2505:
2506: \bibitem{crs}
2507: D. Cohen, Phys. Rev. Lett. {\bf 82}, 4951 (1999). \
2508: D. Cohen and T. Kottos, Phys. Rev. Lett. 85, 4839 (2000).
2509:
2510: \bibitem{frc}
2511: D. Cohen, Annals of Physics 283, 175 (2000).
2512:
2513: \bibitem{wilk}
2514: M. Wilkinson, J. Phys. A {\bf 21}, 4021 (1988). \
2515: M. Wilkinson and E.J. Austin, J. Phys. A {\bf 28}, 2277 (1995).
2516:
2517:
2518:
2519: % relevant
2520:
2521: \bibitem{ophir}
2522: O.M. Auslaender and S. Fishman, Phys. Rev. Lett. {\bf 84}, 1886 (2000). \
2523: O.M. Auslaender and S. Fishman, J. Phys. A {\bf 33}, 1957 (2000).
2524:
2525: \bibitem{qkr}
2526: S. Fishman in {em Quantum Chaos},
2527: Proceedings of the International School
2528: of Physics "Enrico Fermi", Course CXIX,
2529: Ed. G. Casati, I. Guarneri and U. Smilansky
2530: (North Holland 1991).
2531:
2532: \bibitem{dil}
2533: A. Barnett, D. Cohen and E.J. Heller, Phys. Rev. Lett. {\bf 85}, 1412 (2000).
2534:
2535: \bibitem{wlf}
2536: A. Barnett, D. Cohen and E.J. Heller, J. Phys. A {\bf 34}, 413 (2001).
2537:
2538: \bibitem{fisher}
2539: D.S. Fisher and P.A. Lee, Phys. Rev. B {\bf 23}, 6851 (1981).
2540:
2541: \bibitem{stone}
2542: H.U. Baranger and A.D. Stone, Phys. Rev. B {\bf 40}, 8169 (1989).
2543:
2544: \end{thebibliography}
2545:
2546: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2547: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2548:
2549:
2550:
2551:
2552: \clearpage
2553: \onecolumngrid
2554:
2555: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2556: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2557: %begin{figure}
2558: \centerline{\epsfig{figure=pmp_fig,width=0.55\hsize}}
2559: %caption{
2560: {\footnotesize FIG1.
2561: Illustration of a ring system (a).
2562: The shape of the ring is controlled
2563: by some parameters $x_1$ and $x_2$.
2564: The flux through the ring is $x_3=\Phi$.
2565: A system with equivalent topology,
2566: and abstraction of the model are
2567: presented in (b) and (c).
2568: The ``dot" can be represented by an $S$ matrix
2569: that depends on $x_1$ and $x_2$. In (d) also the
2570: flux $x_3$ is regarded as a parameter of the dot.
2571: If we cut the wire in (d) we get the open
2572: two lead geometry of (e). If we put many such
2573: units in series we get the period system in (f).
2574: }
2575: %end{figure}
2576: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2577: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2578:
2579: \ \\
2580:
2581: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2582: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2583: %begin{figure}[h]
2584: \centerline{
2585: \epsfig{figure=pmp_delta,height=0.2\hsize}
2586: \epsfig{figure=pmp_levels,height=0.2\hsize}
2587: }
2588: %caption{
2589: {\footnotesize FIG2.
2590: Schematic illustration of quantum pumping
2591: in a closed wire-dot system. The net charge via the third
2592: level (thick solid line on the right) is vanishingly
2593: small: As the dot potential is lowered an electron
2594: is taken from the left side (first avoided crossing),
2595: and then emitted back to the left side
2596: (second avoided crossing).
2597: Assuming that the bias is inverted before the
2598: dot potential is raised back, only the second level
2599: carry a net charge $Q={\cal O}(1)$.}
2600: %end{figure}
2601: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2602: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2603:
2604:
2605: \ \\
2606:
2607: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2608: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2609: %begin{figure}[h]
2610: \centerline{\epsfig{figure=pmp_bfield_en,width=0.5\hsize}}
2611: \
2612: %caption{
2613: {\footnotesize FIG3.
2614: The first component of the $\vec{\bm{B}}$ field
2615: for a particle in the middle level of
2616: the 3~site lattice model. It is plotted
2617: as a function of the dot potential $x_2=u$.
2618: The other parameters are $\phi=0$, and $c_1=0.1$,
2619: while $c_2=0.04$ for the thick line and
2620: $c_2=0.02$ for the thin line.
2621: In the limit $c_2 \rightarrow 0$,
2622: all the charge that is transfered from the
2623: left side into the dot during the first avoided crossing,
2624: is emitted back into the left side during the second
2625: avoided crossing.
2626: Inset: The eigenenergies $E_n(\bm{x})$ for
2627: the $c_2=0.04$ calculation.}
2628: %end{figure}
2629: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2630: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2631:
2632:
2633:
2634: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2635: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2636: \end{document}
2637: