1: \documentclass[bm]{epl_ARC}
2:
3: % special
4: \usepackage{ifthen}
5: \usepackage{ifpdf}
6:
7: % fonts
8: \usepackage{latexsym}
9: \usepackage{amsmath}
10: \usepackage{amssymb}
11: \usepackage{bm}
12:
13: % figures
14: \ifpdf
15: \usepackage{graphicx}
16: \usepackage{epstopdf}
17: \else
18: \usepackage{graphicx}
19: \usepackage{epsfig}
20: \fi
21:
22:
23:
24:
25: % math symbols I
26: \newcommand{\sinc}{\mbox{sinc}}
27: \newcommand{\const}{\mbox{const}}
28: \newcommand{\trc}{\mbox{trace}}
29: \newcommand{\intt}{\int\!\!\!\!\int }
30: \newcommand{\ointt}{\int\!\!\!\!\int\!\!\!\!\!\circ\ }
31: \newcommand{\ar}{\mathsf r}
32: \newcommand{\im}{\mbox{Im}}
33: \newcommand{\re}{\mbox{Re}}
34:
35:
36: % math symbols II
37: \newcommand{\eexp}{\mbox{e}^}
38: \newcommand{\bra}{\left\langle}
39: \newcommand{\ket}{\right\rangle}
40:
41:
42: % more math commands
43: \newcommand{\tbox}[1]{\mbox{\tiny #1}}
44: \newcommand{\bmsf}[1]{\bm{\mathsf{#1}}}
45: \newcommand{\amatrix}[1]{\begin{matrix} #1 \end{matrix}}
46: \newcommand{\pd}[2]{\frac{\partial #1}{\partial #2}}
47:
48:
49: % equations
50: \newcommand{\be}[1]{\begin{eqnarray}\ifthenelse{#1=-1}{\nonumber}{\ifthenelse{#1=0}{}{\label{e#1}}}}
51: \newcommand{\ee}{\end{eqnarray}}
52:
53:
54: % graphics
55: \newcommand{\drawline}{\begin{picture}(500,1)\line(1,0){500}\end{picture}}
56: \newcommand{\hide}[1]{}
57: \newcommand{\Cn}[1]{\begin{center} #1 \end{center}}
58: \newcommand{\mpg}[2][\hsize]{\begin{minipage}[b]{#1}{#2}\end{minipage}}
59: \newcommand{\putgraph}[2][width=\hsize]{\includegraphics[#1]{#2}}
60:
61: \begin{document}
62:
63:
64: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
66:
67:
\title{The conductance of a multi-mode ballistic ring: \\ beyond Landauer and Kubo}
68: \shorttitle{}
69:
70: \author{Swarnali Bandopadhyay, Yoav Etzioni and Doron Cohen}
71:
72: \institute{
73: {\small Department of Physics, Ben-Gurion University, Beer-Sheva 84105, Israel}
74: \vspace*{-0.3cm}
75: }
76:
77: \pacs{}{Europhysics Letters {\bf 76}, 739 (2006)\vspace*{-0.7cm}}
78:
79: %\pacs{03.65.-w} {Quantum mechanics}
80: %\pacs{05.45.Mt} {Quantum chaos}
81: %\pacs{73.23.-b} {Mesoscopic systems}
82:
83: \maketitle
84:
85:
86:
87: \begin{abstract}
88: The Landauer conductance of a two terminal
89: device equals to the number of open modes
90: in the weak scattering limit.
91: What is the corresponding result if we close
92: the system into a ring?
93: Is it still bounded by the number of open modes?
94: Or is it unbounded as in
95: the semi-classical (Drude) analysis?
96: It turns out that the calculation of
97: the mesoscopic conductance is similar
98: to solving a percolation problem.
99: The ``percolation" is in energy space
100: rather than in real space.
101: The non-universal structures and the sparsity
102: of the perturbation matrix cannot be ignored.
103: \end{abstract}
104:
105:
106: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
107:
108:
109: The theory for the conductance of {\em closed} mesoscopic rings
110: has attracted a lot of interest \cite{rings,debye,IS,loc,wilk,kamenev}.
111: In a typical experiment \cite{orsay}
112: a collection of mesoscopic rings are driven by
113: a time dependent magnetic flux $\Phi(t)$ which creates
114: an electro-motive-force (EMF) ${-\dot{\Phi}}$ in each
115: ring. Assuming that Ohm's law applies, the induced current
116: is ${I=-G\dot{\Phi} }$ and consequently Joule's law gives
117: %
118: \be{1}
119: \mbox{Rate of energy absorption}
120: \ \ = \ \ G\,\dot{\Phi}^2
121: \ee
122: %
123: where $G$ is called the conductance.
124: For diffusive rings the Kubo formula
125: leads to the Drude formula for $G$.
126: A major challenge in past studies
127: was to calculate the weak localization
128: corrections to the Drude result,
129: taking into account the level statistics
130: and the type of occupation \cite{kamenev}.
131: It should be clear that these corrections
132: do not challenge the leading order Kubo-Drude result.
133:
134:
135: It is just natural to ask what is the conductance
136: if the mean free path $\ell$ increases,
137: so that we have a ballistic ring as in Fig.~1,
138: where the total transmission is ${g_T \sim 1}$.
139: To be more precise, we assume that
140: the mean free path $\ell \approx L/(1-g_T)$
141: is much larger than the perimeter $L$ of the ring.
142: In such circumstances ``quantum chaos" considerations
143: become important.
144: %
145: %
146: Surprisingly this question has
147: not been addressed so far \cite{kbf},
148: and it turns out that the answer requires
149: considerations that go well beyond the
150: traditional framework.
151: Following \cite{kbr} we argue that the
152: calculation of the energy absorption in Eq.(\ref{e1})
153: is somewhat similar to solving a percolation problem.
154: The ``percolation" is in energy space
155: rather than in real space.
156: This idea was further elaborated in \cite{slr}
157: using a resistor network analogy (Fig.~2).
158: %
159: As in the standard derivation of the Kubo formula
160: it is assumed that the leading mechanism for absorption
161: is Fermi-golden-rule transitions. These are proportional
162: to the squared matrix elements $|\mathcal{I}_{nm}|^2$
163: of the current operator.
164: %
165: Still, the theory of \cite{kbr}
166: does not lead to the Kubo formula.
167: This is because the rate
168: of absorption depends crucially on the possibility
169: to make {\em connected} sequences of transitions,
170: and it is greatly reduced by the presence of bottlenecks.
171: It is implied that both the structure
172: of the $|\mathcal{I}_{nm}|^2$ band profile
173: and its sparsity play a major role in the calculation of $G$.
174:
175:
176: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
177:
178: The outline of this Letter is as follows:
179: (a) We define a model example for which the
180: analysis is carried out;
181: (b) We make a distinction between
182: the Landauer, the Drude and the
183: actual mesoscopic conductance.
184: (c) We calculate the matrix
185: elements of the current operator;
186: (d) We define an ``averaging" procedure
187: that allows the calculation of~$G$.
188: %
189: The result of the calculation is contrasted
190: with that of the conventional Kubo approach.
191:
192:
193:
194: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
195: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
196:
197: % \section{The model}
198:
199: We regard the ballistic ring (Fig.1) as a set
200: of $\mathcal{M}$ open modes,
201: and a small scattering region
202: that is characterized by its total transmission $g_T$.
203: To be specific we adopt a
204: convenient network model
205: where all the bonds (${a=1,2,...,\mathcal{M}}$)
206: have similar length ${L_a \sim L}$.
207: The scattering is described by
208: %
209: \be{2}
210: \bm{S} = \left( \amatrix{
211: \epsilon\exp\left(i\,2\pi\,\frac{a\, b}{\mathcal{M}}\right)
212: & \sqrt{1-\mathcal{M}\epsilon^2} \delta_{a,b} \\
213: \sqrt{1-\mathcal{M}\epsilon^2} \delta_{a,b}
214: & -\epsilon\exp\left(-i\,2\pi\,\frac{a\,b}{\mathcal{M}}\right)}
215: \right)
216: \ee
217: %
218: The transitions probability matrix $\bm{g}$
219: is obtained by squaring the absolute values
220: of the $\bm{S}$ matrix elements.
221: It is composed of a reflection matrix
222: $[\bm{g}^R]_{a,b} = \epsilon^2$
223: and a transmission matrix
224: ${[\bm{g}^T]_{a,b} = (1-\mathcal{M}\epsilon^2) \delta_{a,b}}$.
225: The total transmission is
226: $g_T = 1-\mathcal{M}\epsilon^2$.
227: If the system were open as in Fig.1c.
228: then its Landauer conductance would be
229: %
230: \be{3}
231: G_{\tbox{Landauer}}
232: = \frac{e^2}{2\pi\hbar}
233: \sum_{a,b} [\bm{g}^T]_{a,b}
234: = \frac{e^2}{2\pi\hbar} \mathcal{M}g_T
235: \ee
236: %
237: If we had a closed ring and we could assume
238: that there is no quantum interference within
239: the bonds, then we could use
240: the multimode conductance
241: formula of Ref.\cite{kbf}
242: %
243: \be{4}
244: G_{\tbox{Drude}} \ \ = \ \ \frac{e^2}{2\pi\hbar}
245: \sum_{a,b} \left[ 2\bm{g}^T / (1{-}\bm{g}^T{+}\bm{g}^R) \right]_{a,b}
246: \ \ = \ \ \frac{e^2}{2\pi\hbar} \mathcal{M}\frac{g_T}{1-g_T}
247: \ee
248: %
249: The first expression can be derived
250: in various ways: Boltzmann picture formalism;
251: semiclassical Kubo formalism; or quantum Kubo
252: calculation that employs a diagonal approximation.
253: In order to get the specific result for
254: our network model we had to invert the
255: matrix ${(1{-}\bm{g}^T{+}\bm{g}^R)}$.
256: %
257: % which involves some linear-algebra tricks.
258: %
259: We see that in the limit $g_T \rightarrow 1$
260: the semiclassical $G_{\tbox{Drude}}$
261: is unbounded, while $G_{\tbox{Landauer}}$ is
262: bounded by the number of open modes.
263:
264:
265:
266: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
267: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
268:
269: %\section{Objective}
270:
271: Our objective is to find the conductance
272: of the closed ring in circumstances such
273: that the motion inside the ring is essentially coherent
274: (quantum interference within the bonds is not ignored):
275: As in the traditional linear response theory (LRT)
276: it is assumed that the level broadening~$\Gamma$
277: is larger compared with the mean level spacing,
278: but otherwise very small semi-classically.
279: On the other hand, in contrast to LRT,
280: we assume ``mesoscopic circumstance",
281: meaning that the environmentally-induced relaxation
282: is very slow compared with the EMF-induced rate of transitions.
283: An extensive discussion of these conditions
284: can be found in \cite{kbr}.
285: The calculation of~$G$ is done using the formula
286: %
287: \be{5}
288: G \ = \ \pi\hbar \,\, \varrho_F^2 \times
289: \langle\langle |\mathcal{I}_{nm}|^2 \rangle\rangle
290: \ = \ \frac{e^2}{2\pi\hbar} \times
291: 2\mathcal{M}^2 \langle \langle |I_{nm}|^2 \rangle \rangle
292: \ \equiv \
293: \frac{e^2}{2\pi\hbar} \times
294: 2\mathcal{M}^2 \mathsf{g}
295: \ee
296: %
297: where $\varrho_F$ is the density of states
298: at the Fermi energy, and $\mathcal{I}_{nm}$
299: are the matrix elements of the current
300: operator. For our network system
301: ${\varrho_F = \mathcal{M} L/(\pi\hbar v_F)}$,
302: where $v_F$ is the Fermi velocity.
303: Furthermore it is convenient to
304: write ${\mathcal{I}_{nm}=-i(ev_F/L)I_{nm}}$
305: so as to deal with real dimensionless quantities,
306: leading to the second expression.
307: %
308: Eq.(\ref{e5}) would be the Kubo
309: formula if ${\langle\langle...\rangle\rangle}$
310: stood for a simple algebraic average.
311: But in view of the percolation-like nature
312: of the energy absorption process,
313: the definition of ${\langle\langle...\rangle\rangle}$
314: involves a more complicated ``averaging" procedure
315: that will be discussed and developed later.
316:
317:
318:
319:
320:
321: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
322: %\section{eigenstates}
323:
324: {\bf The eigenstates:}
325: Our model system, in the absence
326: of driving, is time reversal symmetric.
327: Consequently the unperturbed eigenfunctions
328: can be chosen as real
329: %
330: \be{7}
331: |\psi\rangle
332: =\sum_a
333: A_a \sin(kx+\varphi_a)
334: \,\otimes|a\rangle.
335: \ee
336: %
337: For a given $g_T$ we can find
338: numerically the eigenvalues
339: and the eigenstates, thus
340: obtaining a table ${(k_n, \varphi_a^{(n)}, A_a^{(n)})}$
341: with ${n=\mbox{level index}}$.
342: %
343: %\be{8}
344: %(k_n, \varphi_a^{(n)}, A_a^{(n)})
345: %\ \ \ \ \ \ \ \ n=\mbox{level index}
346: %\ee
347: %
348: In the limit of small $\epsilon$
349: it is not difficult to derive
350: the expressions
351: %
352: \be{9}
353: k_n &\approx&
354: \left(
355: 2\pi \times\mbox{\small integer}
356: \pm \frac{1}{\sqrt{\mathcal{M}}} \ \epsilon
357: \right)
358: \frac{1}{L_a}
359: \\
360: \varphi_a^{(n)} & \approx &
361: -\frac{\pi a^2}{\mathcal{M}}
362: -\frac{1}{2}kL_a
363: +\left\{ \begin{array}{cc}\pi/4\\3\pi/4\end{array} \right.
364: \ee
365: %
366: The numerical results over the
367: whole range of $g_T$ values are
368: presented in Fig.~3.
369: %
370: %
371: By normalization we have
372: ${\sum_a (L_a/2) A_a^2 \approx 1}$.
373: The degree of ergodicity of a
374: wavefunctions is characterized by
375: the participation ratio:
376: %
377: \be{11}
378: \mbox{PR} \equiv
379: \left[
380: \sum_a \left(\frac{L_a}{2} A_a^2 \right)^2
381: \right]^{-1}
382: \approx 1 + \frac{1}{3}(1-g_T) \mathcal{M}
383: \ee
384: %
385: The approximation in the last equality
386: is based on the following observations:
387: By definition we have ${\mbox{PR} \approx 1}$
388: for a wavefunction which is localized on one bond,
389: while ${\mbox{PR} \sim \mathcal{M}}$
390: for an ergodic wavefunction.
391: %
392: In the trivial regime $(1-g_T) \ll 1/\mathcal{M}$
393: the eigenstates are like those of
394: a reflectionless ring, with $\mbox{PR} \sim 1$.
395: Once $(1-g_T)$ becomes
396: larger compared with $1/\mathcal{M}$,
397: first order perturbation theory breaks down,
398: and the mixing of the levels is described
399: by a Wigner Lorentzian. The analysis is completely
400: analogous to that of the single mode case in Ref.\cite{kbr},
401: and leads to $\mbox{PR} \propto (1-g_T) \mathcal{M}$.
402: This is confirmed by the numerical analysis (Fig.~4).
403: In practice we have found that the proportionality
404: constant is roughly~$1/3$.
405: %
406: %
407: Our interest is focused in the {\em non-trivial}
408: ballistic regime
409: %
410: \be{0}
411: 1/\mathcal{M} \ \ \ll \ \ (1-g_T) \ \ \ll \ \ 1
412: \ee
413: %
414: where we have strong mixing of
415: levels ($\mbox{PR} \gg 1$), but still the
416: mean free path $\ell \approx L/(1-g_T)$
417: is very large compared with the
418: ring's perimeter ($\ell \gg L$).
419: It is important to realize that in this regime
420: we do not have ``quantum chaos" ergodicity.
421: Rather we have $\mbox{PR} \ll \mathcal{M}$
422: meaning that the wavefunctions occupy only a small
423: fraction of the classically accessible phase space.
424:
425:
426:
427: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
428: %\section{matrix elements}
429:
430:
431: {\bf The matrix elements:}
432: The current operator $\mathcal{I}$ is
433: the symmetrized version of $e\hat{v}\delta(\hat{x}-x_0)$,
434: where $\hat{v}$ and $\hat{x}$ are the velocity
435: and the position operators respectively.
436: The section through which the current is measured
437: is arbitrary and we simply take $x_0=+0$.
438: Given a set of eigenstates, it is
439: straightforward to calculate the matrix elements
440: of the current operator (Fig.~5), and to
441: get insight into their statistical properties
442: (e.g. Fig.~6). The scaled matrix elements are
443: %
444: \be{12}
445: I_{nm} \approx
446: \sum_a \frac{L_a}{2} A_a^{(n)} A_a^{(m)} \sin(\varphi_a^{(n)} - \varphi_a^{(m)})
447: \ee
448: %
449: % For $n=m$ we have $I_{nm}=0$ as expected from time
450: % reversal considerations.
451: % From now on we are interested in $n \neq m$.
452: %
453: Needless to say that small PR of wavefunctions
454: implies sparsity of $I_{nm}$. It is also worthwhile
455: to point out that there are several extreme cases
456: that allow simple estimates:
457: The case where $n$ and $m$ are
458: localized on different bonds leading to ${|I_{nm}|^2=0}$;
459: The case where $n$ and $m$ are nearly degenerate
460: states localized on the same wire
461: leading to ${|I_{nm}|^2=1}$;
462: The case where $n$ and $m$ are
463: ergodic and uncorrelated
464: leading to ${|I_{nm}|^2 \approx 1/(2\mathcal{M})}$;
465: Irrespective of this, it is clear that
466: by normalization the maximal value
467: that can be obtained is ${|I_{nm}|^2=1}$.
468:
469:
470:
471:
472: {\bf Landauer? Drude?} From Eq.(\ref{e5}) and
473: the above discussion we deduce that
474: %
475: \be{13}
476: && G\Big|_{\tbox{ergodic}} \ \ \ = \frac{e^2}{2\pi\hbar} \mathcal{M}
477: \\ \label{e14}
478: && G\Big|_{\tbox{maximal}} \ \ = \frac{e^2}{2\pi\hbar} 2\mathcal{M}^2
479: \ee
480: %
481: The first expression suggests agreement
482: with the Landauer result if we had
483: complete ``quantum chaos" ergodicity,
484: while Eq.(\ref{e14}) implies a necessary condition
485: for a correspondence with the semiclassical
486: result Eq.(\ref{e4}):
487: %
488: \be{15}
489: \frac{1}{1-g_T} \ll \mathcal{M}
490: \ee
491: %
492: This can be rephrased by saying that
493: the ballistic time
494: ${t_{cl} = (1-g_T)^{-1} \times (L/v_F)}$ should
495: be much smaller compared with the Heisenberg
496: time $t_H=\mathcal{M} \times (L/v_F)$.
497: %
498: In fact it has been argued \cite{kbr},
499: on the basis of a diagonal approximation,
500: that semiclassical correspondence is
501: indeed realized in the `Kubo calculation".
502: By ``Kubo calculation" we mean Eq.(\ref{e5})
503: with algebraic average over the near diagonal
504: matrix elements of ${|I_{nm}|^2}$.
505: The Kubo calculation might have a physical
506: validity in the presence of a strong
507: relaxation process that suppresses
508: the quantum nature of the dynamics.
509: See \cite{kbr} for a detailed discussion of this point.
510:
511:
512: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
513: %\section{The calculation of the conductance}
514:
515:
516: {\bf The FGR picture:}
517: The Hamiltonian in the adiabatic basis is
518: $\mathcal{H} \mapsto E_n\delta_{nm} + \dot{\Phi} W_{nm}$
519: where $W_{nm} = i\hbar\mathcal{I}_{nm}/(E_n{-}E_m)$,
520: and $-\dot{\Phi}$ is the EMF. The FGR transition rate
521: between level~$n$ and level~$m$ is proportional to $|W_{nm}|^2$
522: multiplied by a broadened $\delta(E_n{-}E_m)$
523: which we call $F()$. The effective broadening
524: of the levels reflects either the power spectrum
525: or the non-adiabaticity of the driving.
526: After trivial scaling the dimensionless
527: transition rates are
528: %
529: \be{16}
530: \mathsf{g}_{nm} \ \ = \ \ \frac{|I_{nm}|^2}{(n-m)^2} \
531: \frac{1}{\gamma} F\left(\frac{n-m}{\gamma}\right)
532: \ee
533: %
534: %
535: % For the purpose of numerical
536: % demonstration we assume $F(r)=\exp(-2|r|)$.
537: %
538: The dimensionless broadening parameter~$\gamma$
539: is identical with $\Gamma/\Delta$
540: of Ref.\cite{kbr,pmc} and with $\hbar\omega_0/\Delta$
541: of Ref.\cite{slr}, where $\Delta$ is the mean
542: level spacing. There is an implicit approximation
543: in Eq.(\ref{e16}), namely $(E_n-E_m)/\Delta \approx (n-m)$,
544: that underestimates the exceptionally large couplings
545: between pairs of almost degenerated levels.
546: But this is not going to be reflected in the
547: energy absorption rate, since the latter
548: is indifferent(!) to large sparse values.
549:
550:
551:
552: {\bf The calculation of the conductance:}
553: Given the transition rates $\mathsf{g}_{nm}$
554: we want to calculate the rate of energy
555: absorption and hence~$G$ as defined by Eq.(\ref{e1}).
556: It is most convenient to exploit
557: the ``resistor network" analogy of Ref.\cite{slr}(Fig.~2).
558: Within this framework~$\mathsf{g}$ of Eq.(\ref{e5})
559: is simply the {\em resistivity} of the network.
560: %
561: The practical numerical procedure is as follows:
562: {\bf \ (i)}~Cut an $N$~site segment out of the network.
563: {\bf \ (ii)}~Define a vector ${\bm{J}_n (n=1..N)}$
564: whose elements are all zero except the first
565: and the last that equal ${\bm{J}_1=+J}$ and ${\bm{J}_N=-J}$.
566: {\bf \ (iii)}~Solve the Kirchhoff equation
567: ${\bm{J}_n = \sum_{m} \mathsf{g}_{nm} (\bm{V}_n-\bm{V}_m)}$
568: for the vector $\bm{V}$.
569: {\bf \ (iv)}~Find the overall resistance of
570: the truncated network ${\mathsf{g}_N = J/(\bm{V}_N-\bm{V}_1)}$.
571: And finally: {\bf \ (v)}~Define the resistivity
572: as $\mathsf{g}^{-1}=\mathsf{g}_N^{-1}/N$.
573: %
574: %
575: For a locally homogeneous network it has been argued in \cite{kbr} that
576: ${\mathsf{g} \approx \langle\langle (1/2)\sum_m (m-n)^2 \mathsf{g}_{nm} \rangle\rangle}$,
577: where the sum over~$m$ reflects
578: the addition of resistors in parallel,
579: and the harmonic average ${\langle\langle...\rangle\rangle}$
580: reflects the addition of resistors in series.
581: This expression can be written simply
582: as $\mathsf{g} = \langle\langle |I_{nm}|^2 \rangle\rangle$,
583: with the implicit understanding that
584: the harmonic average is taken over the near diagonal
585: elements of the $\gamma$-smoothed $|I_{nm}|^2$ matrix.
586:
587:
588:
589: {\bf Numerical results:}
590: The results of the calculation are presented
591: in Figs.~7-8. The calculation has been done
592: numerically using the resistor network procedure
593: that has been explained in the previous paragraph.
594: %
595: We did not to rely on the harmonic average approximation
596: because there are prominent structures
597: (notably the strongly coupled nearly degenerate levels)
598: that make it a-priori questionable.
599: However from the numerics it turns out (not displayed)
600: that the harmonic average is doing quite well.
601: We mention this fact because it gives an insight for
602: the numerical results which are displayed in Fig.~7.
603:
604:
605: Our numerical results suggest that {\em typically}
606: ${G < G_{\tbox{Landauer}}}$.
607: For an optimal value of $\gamma$,
608: such that~$G$ is maximal, it seems that we still
609: have ${G \lesssim G_{\tbox{Landauer}}}$.
610: It is too difficult to figure out the numerical
611: prefactor which is involved in the latter inequality (Fig.~8).
612: Our conjecture is that this inequality is true {\em in general}
613: (disregarding the prefactor which is of order~1).
614: We did not find a mathematical argument
615: to establish this conjecture, except the
616: very simple case of a single mode ballistic
617: ring\cite{kbr} where the calculations
618: of~$G$ can be done analytically.
619:
620:
621: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
622: %\section{Conclusions}
623:
624: {\bf Conclusions:}
625: In this Letter we have studied the mesoscopic
626: conductance of a ballistic ring
627: with mean free path $\ell \gg L$.
628: The specific calculation has been done for
629: a network model, but all its main ingredients
630: are completely {\em generic}.
631: %
632: Ballistic rings with $\ell \gg L$ are not
633: typical ``quantum chaos" systems.
634: Their eigenfunctions are not ergodic over
635: the whole accessible phase space, and therefore
636: the perturbation matrix $\mathcal{I}_{nm}$
637: is highly structured and sparse. Consequently the
638: Kubo formula is no longer valid, and one has to
639: adopt an appropriate ``resistor network" procedure
640: in order to calculate the true mesoscopic conductance.
641: %
642: {\em However}, it should be emphasized that if
643: there is either a very effective
644: relaxation or decoherence process,
645: then the theory that we have discussed does not apply.
646: In the presence of strong environmental
647: influence one can justify, depending
648: on the {\em circumstances} \cite{kbf,kbr},
649: either the use of the traditional
650: Kubo-Drude result, or the use of the Landauer result.
651:
652:
653:
654: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
655: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
656:
657: {\bf Acknowledgment:}
658: Much of the motivation for this work came from
659: intriguing meetings of DC during 2004-2005
660: with Michael Wilkinson, who highlighted
661: the open question regarding the
662: feasibility to get $G>G_{\tbox{Landauer}}$
663: in the case of a multimode closed ring.
664: We also thank Bernhard~Mehlig, Tsampikos~Kottos
665: and Holger~Schanz for inspiring discussions.
666: The research was supported by the
667: Israel Science Foundation (grant No.11/02).
668: % and by a grant from the GIF, the German-Israeli Foundation
669: % for Scientific Research and Development.
670: % and by a grant from the DIP
671:
672:
673: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
674: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
675: \begin{thebibliography}{99}
676:
677:
678: \bibitem{rings}
679: M. Buttiker, Y. Imry and R. Landauer,
680: Phys. Lett. {\bf 96A}, 365 (1983).
681:
682:
683: \bibitem{debye}
684: The Debye relaxation mechanism is discussed by
685: R. Landauer and M. Buttiker,
686: Phys. Rev. Lett. {\bf 54}, 2049 (1985).
687: M. Buttiker, Phys. Rev. B {\bf 32}, 1846 (1985).
688:
689:
690: \bibitem{IS}
691: The Kubo formula is applied to diffusive rings
692: by Y. Imry and N.S. Shiren,
693: Phys. Rev. B {\bf 33}, 7992 (1986).
694:
695:
696: \bibitem{loc}
697: Y. Gefen and D. J. Thouless,
698: Phys. Rev. Lett. {\bf 59}, 1752 (1987).
699:
700:
701: \bibitem{wilk}
702: M. Wilkinson, J. Phys. A {\bf 21} (1988) 4021.
703: M. Wilkinson and E.J. Austin,
704: J. Phys. A {\bf 23}, L957 (1990).
705:
706:
707: \bibitem{kamenev} A review for the theory
708: of conductance of closed diffusive rings has
709: been written by A. Kamenev and Y. Gefen,
710: Int. J. Mod. Phys. {\bf B9}, 751 (1995).
711:
712:
713: \bibitem{orsay}
714: Measurements of conductance of closed
715: diffusive rings are described by
716: B. Reulet M. Ramin, H. Bouchiat and D. Mailly,
717: Phys. Rev. Lett. {\bf 75}, 124 (1995).
718:
719:
720: \bibitem{kbf}
721: D. Cohen and Y. Etzioni, J. Phys. A {\bf 38}, 9699 (2005).
722:
723:
724: \bibitem{kbr}
725: D. Cohen, T. Kottos and H. Schanz,
726: cond-mat/0505295, J. Phys. A {\bf 39}, 11755 (2006).
727:
728:
729: \bibitem{slr}
730: M. Wilkinson, B. Mehlig and D. Cohen,
731: Europhys. Lett. {\bf 75}, 709 (2006).
732:
733:
734:
735: \bibitem{pmc}
736: The $\Gamma$ issue is best discussed in Section~VIII of
737: D. Cohen, Phys. Rev. B {\bf 68}, 155303 (2003).
738:
739:
740:
741:
742: \end{thebibliography}
743:
744:
745:
746: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
747: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
748:
749:
750: % FIGURES:
751: % blsfig
752: % knumeric
753: % PRdistnu
754: % imageInm; imageInmzoom
755: % sigmavsgT; sigmavsgTnormal
756: % conductance
757:
758:
759: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
760: % illustration of the model
761:
762:
763: \newpage
764:
765: \mpg[0.4\hsize]{
766: \begin{center}
767: \putgraph[width=\hsize]{bls_models}
768: \end{center}
769:
770: {\footnotesize
771: {\bf Fig.1:}
772: {\bf (a)} A billiard example for a ballistic ring.
773: The annular region supports $\mathcal{M}$ open modes.
774: The electrons are scattered by a small bump.
775: {\bf (b)} A network model of a ballistic ring.
776: In the numerics the lengths of the $\mathcal{M}$
777: bonds (${0.9<L_a<1.1}$) are chosen in random.
778: The scattering is described by an $S$ matrix.
779: {\bf (c)} The associated open (leads) geometry which
780: is used in order to define the $S$ matrix
781: and the Landauer conductance. }
782:
783: }
784: %
785: %
786: \hspace{0.1\hsize}
787: %
788: %
789: \mpg[0.5\hsize]{
790: \begin{center}
791: \putgraph[width=0.9\hsize]{Resistors_Network}
792: \end{center}
793:
794: {\footnotesize
795: {\bf Fig.2:}
796: The EMF induces diffusion of probability
797: in energy space, and hence absorption of energy.
798: Within the framework of
799: the Fermi golden rule picture the flow
800: of the probability in the multi level system
801: is analogous to the flow of a current
802: via a resistor network. The resistance of
803: each ``resistors" corresponds to~$g_{nm}^{-1}$.
804: The inverse of the diffusion coefficient is
805: re-interpreted as the resistivity of the network.
806: On the right we display a truncated segment,
807: where $+J$ is the current injected from one end
808: of the network, while $-J$ is the same current
809: extracted from the other end.
810: The injected current to all the other nodes is zero.}
811:
812: }
813:
814:
815:
816:
817:
818:
819:
820:
821:
822:
823:
824: \mpg[0.45\hsize]{
825: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
826: % eigen-energies versus 1-g
827:
828:
829: \begin{center}
830: \includegraphics[clip,width=\hsize]{knumeric}
831: \end{center}
832:
833: {\footnotesize
834: {\bf Fig.3:}
835: The eigenvalues $k_n$ within a
836: small energy window around $k \sim 2000$
837: are shown as a function of $1-g_T$.
838: We consider here a network model
839: with $\mathcal{M}=50$ bonds.
840: The length of each bond
841: was chosen in random
842: within $0.9 < L_a <1.1$. }
843:
844: }
845: %
846: %
847: %
848: \hspace*{0.1\hsize}
849: %
850: %
851: \mpg[0.45\hsize]{
852: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
853: % PRs versus 1-g
854:
855: \vspace*{8mm}
856: \begin{center}
857: \includegraphics[clip,width=\hsize]{PR_dist}
858: \end{center}
859:
860: {\footnotesize
861: {\bf Fig.4:}
862: For each value of $g_T$ we calculate
863: the participation ratio (PR) for all the eigenstates.
864: We display (as symbols) the minimum value,
865: the maximum value, and a set of randomly chosen
866: representative values.
867: The solid line is the average PR,
868: while the dotted line is
869: the formula ${\mbox{PR} \approx 1+\frac{1}{3}(1-g_T)\mathcal{M}}$. }
870:
871: }
872:
873:
874:
875:
876: %%%%%%%%%%%%%%%%%%%%%
877: %%%%%%%%%%%%%%%%%%%%%
878: \newpage
879:
880:
881: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
882: % I_nm images
883:
884: \mpg{
885: \begin{center}
886: \includegraphics[clip,width=0.45\hsize]{imageInm}
887: \includegraphics[clip,width=0.45\hsize]{imageInmzoom}
888: \end{center}
889:
890: {\footnotesize
891: {\bf Fig.5:}
892: The image of the perturbation matrix
893: $|I_{nm}|^2$ for $g_T=0.9$.
894: The right panel is a zoomed image.
895: If we chose larger $1-g_T$ more elements
896: would become non-negligible,
897: and the matrix would become less
898: structured and less sparse. }
899:
900:
901: }
902:
903:
904:
905:
906: \vspace*{5mm}
907: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
908: % diagonals
909:
910:
911:
912: \mpg{
913: \begin{center}
914: % corrected figs
915: \includegraphics[clip,height=0.3\hsize]{sigmavsgTnormal}
916: \includegraphics[clip,height=0.3\hsize]{sigmavsgT}
917: \end{center}
918:
919: {\footnotesize
920: {\bf Fig.6:}
921: The $n$-averaged value
922: of $2\mathcal{M}^2 |I_{n,n+r}|^2$
923: as a function of $1-g_T$
924: for $r=1,2,3,4,5$.
925: The ergodic value $\mathcal{M}$
926: and half the maximal value $\mathcal{M}^2$
927: are indicated by horizontal dotted lines.
928: The left panel is normal scale,
929: while the right panel is log-log scale.
930: }
931:
932:
933: }
934:
935:
936:
937:
938:
939: \vspace*{5mm}
940: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
941: % conductance
942:
943:
944:
945: \mpg[0.45\hsize]{
946:
947: \begin{center}
948: % corrected fig
949: \includegraphics[clip,width=0.9\hsize]{conductance}
950: \end{center}
951:
952: {\footnotesize
953: {\bf Fig.7:}
954: The mesoscopic conductance $G$ in units of $e^2/(2\pi\hbar)$
955: as a function of $1-g_T$ for $\gamma=1,2,3,4,5$.
956: Note again that the total number of open modes
957: in our numerics is $\mathcal{M}=50$. The dotted
958: line is $G_{\tbox{Landauer}}$ while the dashed
959: line is $G_{\tbox{Drude}}$.}
960:
961:
962: }
963: %
964: %
965: \hspace{0.1\hsize}
966: %
967: %
968: \mpg[0.45\hsize]{
969:
970: \begin{center}
971: \putgraph[width=0.9\hsize]{Conductance_M}
972: \end{center}
973:
974: {\footnotesize {\bf Fig.8:}
975: The mesosocopic conductance
976: divided by the number of open modes
977: for $\mathcal{M}$ up to $300$.
978: Here $\gamma=3$ and $g_T=0.8$.
979: The different curves are calculated
980: with segments of length $N=30,50,70$,
981: so as to provide an estimate for the numerical error.
982: }
983:
984: }
985:
986:
987:
988:
989:
990:
991: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
992: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
993: \end{document}
994:
995:
996: