1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: \documentclass[aps,prb,twocolumn,floats]{revtex4}
4:
5: % special
6: \usepackage{ifthen}
7: \usepackage{ifpdf}
8: %\usepackage{color}
9:
10:
11: % fonts
12: \usepackage{latexsym}
13: \usepackage{amsmath}
14: \usepackage{amssymb}
15: \usepackage{bm}
16: %\usepackage{wasysym}
17:
18:
19: \ifpdf
20: \usepackage{graphicx}
21: \usepackage{epstopdf}
22: \else
23: \usepackage{graphicx}
24: \usepackage{epsfig}
25: \fi
26:
27: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
28:
29: % NEW
30: %\definecolor{red}{rgb}{1.,0.,0.}
31: %\newcommand{\revision}[1]{\textcolor{red}{#1}}
32: \newcommand{\abs}[1]{\left|#1\right|}
33: \newcommand{\signC}{^{\!\!c}}
34: \newcommand{\signA}{^{\!\!a}}
35: \newcommand{\citeref}[1]{[\onlinecite{#1}]}
36:
37:
38: % math symbols I
39: \newcommand{\sinc}{\mbox{sinc}}
40: \newcommand{\const}{\mbox{const}}
41: \newcommand{\trc}{\mbox{trace}}
42: \newcommand{\intt}{\int\!\!\!\!\int }
43: \newcommand{\ointt}{\int\!\!\!\!\int\!\!\!\!\!\circ\ }
44: \newcommand{\ar}{\mathsf r}
45: \newcommand{\im}{\mbox{Im}}
46: \newcommand{\re}{\mbox{Re}}
47:
48: % math symbols II
49: \newcommand{\eexp}{\mbox{e}^}
50: \newcommand{\bra}{\left\langle}
51: \newcommand{\ket}{\right\rangle}
52:
53: % Mass symbol
54: \newcommand{\mass}{\mathsf{m}}
55: \newcommand{\Mass}{\mathsf{M}}
56:
57: % more math commands
58: \newcommand{\tbox}[1]{\mbox{\tiny #1}}
59: \newcommand{\bmsf}[1]{\bm{\mathsf{#1}}}
60: %\newcommand{\amatrix}[1]{\matrix{#1}}
61: \newcommand{\amatrix}[1]{\begin{matrix} #1 \end{matrix}}
62: \newcommand{\pd}[2]{\frac{\partial #1}{\partial #2}}
63:
64: % equations
65: \newcommand{\be}[1]{\begin{eqnarray}\ifthenelse{#1=-1}{\nonumber}{\ifthenelse{#1=0}{}{\label{e#1}}}}
66: \newcommand{\ee}{\end{eqnarray}}
67:
68: % arrangement
69: \newcommand{\hide}[1]{}
70: \newcommand{\drawline}{\begin{picture}(500,1)\line(1,0){500}\end{picture}}
71: \newcommand{\bitem}{$\bullet$ \ \ \ }
72: \newcommand{\Cn}[1]{\begin{center} #1 \end{center}}
73: \newcommand{\mpg}[2][1.0\hsize]{\begin{minipage}[b]{#1}{#2}\end{minipage}}
74: \newcommand{\mpgt}[2][1.0\hsize]{\begin{minipage}[t]{#1}{#2}\end{minipage}}
75:
76:
77: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
78: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
79:
80: \begin{document}
81:
82: \title[Transport in a double well system]
83: {Quantum dynamics and transport in a double well system}
84:
85: \author{Itamar Sela and Doron Cohen}
86:
87: \affiliation{
88: Department of Physics, Ben-Gurion University, Beer-Sheva 84005, Israel}
89:
90: \begin{abstract}
91: The simplest one-dimensional model for the studying
92: of non-trivial geometrical effects is a ring shaped device
93: which is formed by joining two arms.
94: We explore the possibility to model such a system
95: as a two level system (TLS).
96: Of particular interest is the analysis of {\em quantum~stirring},
97: where it is not evident that the topology
98: is properly reflected within the framework of the TLS modeling.
99: On the technical side we provide
100: a practical ``neighboring level" approximation
101: for the analysis of such quantum devices,
102: which remains valid even if the TLS modeling does not apply.
103: \end{abstract}
104:
105: \maketitle
106:
107:
108: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
109: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
110: \section{Introduction}
111: \label{S1}
112:
113: In this paper we explore the possibility to model
114: a ring shaped device (Fig.1(a)) as a two level system (TLS) (Fig.1(b)).
115: We shall see that both technical and conceptual
116: difficulties are involved.
117: The model Hamiltonian is
118: %
119: \be{1}
120: {\cal H} = \frac{1}{2\mass}\hat{p}^2+
121: V_A(\hat{x}-x_A)+V_B(\hat{x}-x_B)
122: \ee
123: %
124: with periodic boundary conditions
125: over $x\in[-L/2,L/2]$ so as to have a ring geometry,
126: as illustrated in Fig.1(a).
127: $V_A$ and $V_B$ represent high barriers,
128: such that the ring is composed of two weakly coupled arms.
129: We assume that we have control over some geometrical parameters
130: of the model and in particular over the heights $X_A$ and $X_B$
131: of both barriers.
132: %
133: Our main interest is in the current that
134: flows in the system. The current through an arbitrary point $x_0$ is
135: obtained as the expectation value of the operator
136: %
137: \be{2}
138: {\cal I} = \frac{1}{2\mass}
139: \left(
140: \hat{p} \, \delta(\hat{x}-x_0)+
141: \delta(\hat{x}-x_0) \, \hat{p}
142: \right)
143: \ee
144:
145:
146: Having defined the system and its observables
147: we can consider various dynamical scenarios
148: such as coherent {\em Bloch~oscillations} between the two arms.
149: Then we can ask whether a TLS modeling
150: is meaningful. Of particular interest for us is
151: the analysis of {\em quantum~stirring} \citeref{pmx}:
152: this means to induce a circulating current by periodic modulation
153: of the potential.
154:
155:
156: We note that transport due to periodic modulations
157: of the potential~\citeref{Thouless} has been
158: studied mainly in the context of quantum pumping~\citeref{bpt,BPT2,Avron},
159: where the current is induced between reservoirs.
160: The notion of quantum stirring relates
161: to closed geometry, where the emerging physical picture
162: is significantly different~\citeref{pMB,pms}.
163:
164:
165: The quantum stirring problem highlights
166: an obvious topological subtlety: one wonders whether
167: the non trivial topology of the ring is properly
168: reflected in the effective TLS model.
169:
170:
171: On the technical side we define the unperturbed
172: Hamiltonian ${\cal H}_0$ as the $X_A=X_B=\infty$ limit.
173: In this limit the
174: two arms are disconnected from each other,
175: and the diagonalization
176: gives a set of eigen-energies $E_i$ such that each eigenstate belongs
177: to only one of the two arms. Then we make either $X_A$ or $X_B$ or both finite,
178: and we ask what is the perturbation matrix $W_{ij}$ in the reduced Hamiltonian
179: %
180: \be{3}
181: {\cal H}_{ij}
182: = \left(\amatrix{E_1 & 0 \cr 0 & E_2}\right)
183: + \left(\amatrix{W_{11} & W_{12} \cr W_{21} & W_{22}}\right)
184: \ee
185: %
186: Obviously we would like to express the perturbation
187: using the transmission coefficients of the barriers.
188:
189:
190:
191: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
192: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
193: \section{Outline}
194: \label{S2}
195:
196: In the first part of the paper (Sections~\ref{S3}-\ref{S5})
197: we establish the building blocks.
198: We derive expressions for the perturbation matrix $W_{ij}$
199: and for the reduced current operator ${\cal I}_{ij}$,
200: %and for some other useful operators ${\cal F}_{ij}$,
201: and figure out how the topology is reflected
202: in the reduced description.
203:
204:
205: In the second part of the paper (Sections~\ref{S6}-\ref{S7})
206: we turn to the applications.
207: We begin with the simplest problems:
208: The coherent Bloch oscillations of a particle
209: in a mirror symmetric device,
210: and the Wigner decay of a particle
211: from a short arm to a long arm.
212: Then we continue with the quantum stirring problem,
213: and show how one can derive expressions
214: for the geometric conductance.
215:
216:
217: In the third part of the paper (Sections~\ref{S8}-\ref{S11})
218: we address some non-trivial technical points
219: that are associated with the analysis,
220: thus exploring the limitations of the TLS modeling.
221: We demonstrate that even if the TLS modeling
222: does not apply, we still can use
223: a {\em neighboring level approximation}
224: in order to extract results for the geometric conductance.
225:
226:
227: In the Summary (Section~\ref{S12}) we highlight the practical value
228: of our findings for the purpose of design and analysis
229: of quantum stirring devices, and we briefly
230: relate to the experimental measurement issue.
231:
232:
233:
234: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
235: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
236: \section{The TLS modeling scheme}
237: \label{S3}
238:
239:
240: The unperturbed eigenstates $\psi^{i}(x)$ are labeled
241: as ${i=1,2}$, corresponding to the two arms of the ring.
242: The associated eigen-energies are ${E_i = k_i^2/(2\mass)}$.
243: We have
244: %
245: \be{4}
246: \psi^{(1)}(x) =
247: \left\{\amatrix{
248: \sqrt{\frac{2}{L_1}}\sin(k_1x+\varphi_1) & \mbox{if $x \in 1$st arm} \cr
249: 0 & \mbox{if $x \in 2$nd arm}
250: } \right.
251: \ee
252: %
253: and a similar expression for $\psi^{(2)}(x)$,
254: where $L_i$ is the length of the $i$th arm,
255: and $|\varphi|<\pi/2$.
256: Two representative eigenstates are illustrated in Fig.2.
257: Note that the wavenumber of the particle in the $i$th arm
258: is $k_i=(\pi/L_i)\times\mbox{\small integer}$.
259: %
260: Our interest is in a very small
261: energy range ${E_1 \sim E_2 \sim E}$,
262: where the wavenumbers are ${k_1 \sim k_2 \sim k_E}$,
263: corresponding to the velocity ${v_E=(2E/\mass)^{1/2}}$.
264: We would like to ignore all the other levels.
265: Later we discuss the validity conditions
266: for this TLS modeling scheme.
267:
268:
269: Once we lower from infinity one barrier, say barrier A,
270: the two states become coupled. In section \ref{S4}
271: we consider a delta barrier and obtain the
272: following expression for the perturbation matrix:
273: %
274: \be{333}
275: W^A_{ij} = -\frac{v_E}{2\sqrt{L_iL_j}} \ \sqrt{g_A}
276: \ee
277: %
278: where $g_A$ is the transmission of the barrier.
279: For $i\neq j$ the minus sign is a convention
280: that fixes the gauge (see Appendix~A).
281: %
282: In section \ref{S5} we show that essentially the
283: same result applies to any other type of barrier,
284: but the ${i=j}$ expression for the energy shift
285: should be somewhat generalized.
286:
287:
288: If both barriers are finite the two associated
289: perturbation terms should be added together,
290: and one obtains for the energy difference
291: %
292: \be{6}
293: \varepsilon = (E_1 + W_{11}^A + W_{11}^B )-(E_2 + W_{22}^A+W_{22}^B)
294: \ee
295: %
296: and for the coupling
297: %
298: \be{7}
299: \frac{\kappa}{2} = W_{12}^A+W_{12}^B = -\frac{v_E}{2\sqrt{L_1L_2}} \left(\sqrt{g_A} \pm\signC \sqrt{g_B}\right)
300: \ee
301: %
302: The latter expression involves a relative sign $\pm\signC$
303: that cannot be gauged away (see Appendix~A).
304: If we had magnetic flux penetrating
305: through the ring we could have, instead
306: of the $\pm\signC$, an arbitrary phase factor.
307:
308: Using the Pauli matrices we can write the TLS Hamiltonian as
309: %
310: \be{0}
311: \mathcal{H}_{ij} \ \ = \ \
312: \frac{\varepsilon}{2} \bm{\sigma}_z + \frac{\kappa}{2} \bm{\sigma}_x
313: \ \ \equiv \ \
314: \frac{\bm{\Omega}}{2} \cdot \bf{\sigma}
315: \ee
316: %
317: Defining $\theta$ as the angle between
318: $\Omega$ and the "z" axis, with the convention $0<\theta<\pi$,
319: the eigenstates ${n_0}$ and ${m_0}$ of this Hamiltonian are:
320: %
321: \be{9}
322: |n_0\rangle
323: =
324: \left( \begin{array}{c}
325: \mp\sin{(\theta/2)} \\ \cos{(\theta/2)}
326: \end{array} \right),
327: \ \ \ \
328: |m_0\rangle
329: =
330: \left( \begin{array}{c}
331: \cos{(\theta/2)} \\ \pm\sin{(\theta/2)}
332: \end{array} \right)
333: \ee
334: %
335: where the ${\pm}$ indicates the sign of $\kappa$.
336: The energy difference between these eigenstates is
337: %
338: \be{0}
339: \Omega \ \ = \ \ \sqrt{\varepsilon^2 + \kappa^2}
340: \ee
341: %
342: If we have a symmetric well then the effective coupling
343: between odd and even levels vanishes (${\kappa=0}$),
344: and then we can get a degeneracy provided we tune appropriately
345: the energy level difference~$\varepsilon$.
346:
347: The TLS description is valid if $W_{12}$
348: is much smaller compared with the level spacing, namely
349: %
350: \be{11}
351: \mbox{max}\{g_A, \ g_B\} \ \ \ll \ \ L_2/L_1
352: \ee
353: %
354: where without loss of generality we assume $L_1>L_2$.
355:
356: In section \ref{S9} we are going to derive expressions
357: for the current $\mathcal{I}^A$ through barrier~A,
358: as defined by Eq.(\ref{e2}) with ${x_0=x_A}$.
359: One observes that the matrix elements of this
360: operator in the ``standard basis" of Eq.(\ref{e4}) vanish,
361: because the unperturbed wavefunctions are zero
362: at the barriers. The more careful treatment
363: reveals that the reduced operator that gives
364: the net current from the first arm to the second arm is
365: %
366: \be{10}
367: \mathcal{I}_{ij} \ \ = \ \ \frac{\kappa}{2} \, \bm{\sigma}_y
368: \ee
369: %
370: and it turns out that $\mathcal{I}_{ij}^A=\lambda_A\mathcal{I}_{ij}$,
371: and $\mathcal{I}_{ij}^B=\lambda_B\mathcal{I}_{ij}$,
372: where the splitting ratio is defined as
373: %
374: \be{12}
375: \lambda_A=\frac{W_{12}^A}{W_{12}^A+ W_{12}^B}=
376: \frac{\sqrt{g_A}}{\sqrt{g_A}\pm\signC\sqrt{g_B}}
377: \ee
378: %
379: with a similar definition for $\lambda_B$.
380: We have ${\lambda_A+\lambda_B=1}$, but contrary
381: to the naive point of view ${0<\lambda_A<1}$ is not implied.
382: Rather, if the two states have opposite parity,
383: then one~$\lambda$ is larger than $100\%$,
384: while the other~$\lambda$ is negative.
385: We shall see later in Section \ref{S12} that the physical
386: interpretation of the ``splitting ratio"
387: requires recognition in the existence of induced
388: circulating current in the system.
389: Thus the multiple path topology of the system
390: is reflected in the TLS modeling via~$\lambda$.
391:
392:
393:
394: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
395: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
396: \section{The expression for $W_{ij}$ for a delta barrier}
397: \label{S4}
398:
399: Let us assume that barrier~B is infinitely high,
400: while barrier~A is modeled as a delta function.
401: In other words: we consider the simplest possibility
402: of having an infinite well [${(-L/2)<x<(L/2)}$]
403: which is divided at ${x=x_A}$ by a delta function:
404: %
405: \be{0}
406: V_A(x-x_A) \ \ = \ \ X_A\delta(x-x_A)
407: \ee
408: %
409: The total perturbation is obtained from
410: a sequence of infinitesimal variations
411: of the barrier height
412: %
413: \be{13}
414: {\cal H}(X_A)
415: \ \ &=& \ \
416: {\cal H}(\infty)-\int^\infty_{X_A}
417: \frac{\partial{\cal H}}{\partial X} \ \mbox{d}X
418: \\
419: \ \ &\equiv& \ \ {\cal H}(\infty)+W^A
420: \ee
421: %
422: For any value of $X$ the Hilbert space of the system
423: is spanned by a set of (real) eigenfunction labeled by~$n$.
424: The matrix elements for an infinitesimal variation
425: of the barrier height is
426: %
427: \be{0}
428: \left(\frac{\partial{\cal H}}{\partial X}\right)_{nm}
429: \ \ = \ \ \psi^{(n)}(x_A) \,\, \psi^{(m)}(x_A)
430: \ee
431: %
432: Using the matching conditions for a delta potential
433: at ${x=x_A}$ we can express the wave function by its
434: derivative:
435: %
436: \be{0}
437: \psi^{(n)}(x_A)
438: = \frac{1}{2\mass X_A} \left[\partial\psi^{(n)}(x_A{+}0)-\partial\psi^{(n)}(x_A{-}0)\right]
439: \ee
440: %
441: A more elegant way of writing this relation is
442: %
443: \be{16}
444: \psi^{(n)}(x_A)
445: = \frac{1}{2\mass X_A} \sum_{a=1,2} \partial_a\psi^{(n)}(x_A)
446: \ee
447: %
448: where $\partial_a$ is defined as the {\em radial} derivative
449: in the direction of the $a$th arms that stretch out
450: of the junction at ${x=x_A}$. Defining the total
451: radial derivative as ${\partial=\partial_1+\partial_2}$
452: we get
453: %
454: \be{0}
455: \left(\frac{\partial{\cal H}}{\partial X}\right)_{nm}
456: =
457: \frac{1}{(2\mass X_A)^2}
458: \,\,
459: \partial\psi^{(n)}(x_A)
460: \,\,
461: \partial\psi^{(m)}(x_A)
462: \ee
463: %
464: %
465: %
466: For a large barrier with small transmission
467: %
468: \be{18}
469: g_A \ \ \approx \ \ \left(\frac{v_E}{X_A}\right)^2 \ \ \ll \ \ 1
470: \ee
471: %
472: the $n$th and $m$th states remain similar
473: to some unperturbed $i$th and $j$th states.
474: Accordingly, upon integration we get
475: from Eq.(\ref{e13}) the result
476: %
477: \be{20}
478: W^A_{ij} = -\frac{1}{4\mass^2 X_A}
479: \Big[\partial\psi^{(i)}(x_A)\Big]
480: \,\,
481: \Big[\partial\psi^{(j)}(x_A)\Big]
482: \ee
483: %
484: Note that in the last equation the contribution to the
485: total derivative $\partial$ comes from one term only,
486: because each unperturbed wavefunction $\psi^{(i)}(x)$ is non-zero
487: only in one box. Using Eq.(\ref{e18}) we get Eq.(\ref{e333}).
488:
489:
490:
491: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
492: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
493: \section{The expression for $W_{ij}$ for a general barrier}
494: \label{S5}
495:
496: It is possible to deduce an expression for $W_{ij}$
497: without assuming a specific form of potential barrier.
498: For the purpose of this calculation we describe the barrier
499: at ${x=x_A}$ by a general scattering matrix
500: %
501: \be{0}
502: S &=& \eexp{i\gamma}
503: \left(
504: \begin{array}{cc}
505: i\sqrt{1-g} \ \eexp{i\alpha} & -\sqrt{g} \\
506: -\sqrt{g} & i\sqrt{1-g} \ \eexp{-i\alpha}
507: \end{array}
508: \right)
509: \ee
510: %
511: Regarding the barrier as a {\em junction} it can be
512: embedded either in a {\em closed} ring geometry with the two {\em arms} attached,
513: or in an {\em open} one-dimensional geometry with two infinite {\em leads} attached.
514: In both cases the differential representation
515: of $W$ should be the {\em same}, because $W$ is local in space.
516: In other words $W^A_{ij}$ should come out the same
517: for the wavefunctions $\psi^{(i)}(x)$ and $\psi^{(j)}(x)$
518: of the ring, if in the vicinity of ${x=x_A}$
519: they are identical with $\Psi^{(i)}(x)$ and $\Psi^{(j)}(x)$
520: of the scattering geometry.
521:
522:
523: In the scattering geometry it is conventional
524: to label the two leads by ${a=1,2}$
525: and to define a radial coordinate $r=|x-x_A|$.
526: The flux normalized scattering states
527: of the junction (assuming outgoing waves)
528: are $\Psi^{(i+)}$. By definition we have
529: %
530: \be{0}
531: \Psi^{(1+)} =
532: \left\{\amatrix{
533: \frac{1}{\sqrt{v_E}} [\eexp{-ik_E r} - S_{11}\eexp{ik_E r}] & \mbox{if $r \in 1$st lead} \cr
534: \frac{1}{\sqrt{v_E}} [-S_{21}\eexp{ik_E r}] & \mbox{if $r \in 2$nd lead}
535: } \right.
536: \ee
537: %
538: A similar expression holds for $\Psi^{(2+)}$.
539: If the leads are not coupled, the scattering
540: matrix becomes $S_0$ with ${g=0}$.
541: In the vicinity of ${x=x_A}$ the unperturbed scattering states
542: coincide with those of Eq.(\ref{e4}) up to normalization.
543: Namely, in the vicinity of ${x=x_A}$ we have the relation
544: %
545: \be{4444}
546: \Psi^{(i)}(x) = -i\left(\frac{2L_i}{v_E}\right)^{1/2}
547: \eexp{i\varphi_i}
548: \,\, \psi^{(i)}(x)
549: \ee
550: %
551: where
552: %
553: \be{0}
554: \varphi_i = \frac{1}{2} \left( \gamma_0 + \frac{\pi}{2} \pm \alpha_0 \right)
555: \ee
556: %
557: with $\pm$ sign for $i=1,2$ respectively.
558:
559:
560: The relation between the scattering matrix
561: and the perturbation matrix~$W$ can be
562: deduced via the $T$~matrix formalism.
563: The $S$ matrix is related to the $T$ matrix
564: through ${S = (1-iT)S_0}$, or more explicitly
565: %
566: \be{0}
567: [SS_0^{-1}]_{ij} = \delta_{ij} - i\langle \Psi^{(i)} | T | \Psi^{(j)} \rangle
568: \ee
569: %
570: In leading order $T$ equals $W$ so we have
571: %
572: \be{0}
573: \langle \Psi^{(i)} | W | \Psi^{(j)} \rangle \ \ \approx \ \ i (S-S_0) \ S_0^{-1}
574: \ee
575: %
576: where
577: %
578: \be{0}
579: S-S_0 =
580: \eexp{i\gamma_0}
581: \left(
582: \begin{array}{cc}
583: \eexp{i\alpha_0}(\delta\gamma+\delta\alpha) & \sqrt{g} \\
584: \sqrt{g} & \eexp{-i\alpha_0}(\delta\gamma-\delta\alpha)
585: \end{array}
586: \right)
587: \ee
588: %
589: Thus
590: %
591: \be{0}
592: \langle \Psi^{(i)} | W | \Psi^{(j)} \rangle =
593: -\left(
594: \begin{array}{cc}
595: \delta\gamma+\delta\alpha & \sqrt{g} \ \eexp{i\alpha_0} \\
596: \sqrt{g} \ \eexp{-i\alpha_0} & \delta\gamma-\delta\alpha
597: \end{array}
598: \right)
599: \ee
600: %
601: Using Eq.(\ref{e4444}) we deduce that each
602: element of ${\langle \psi^{(i)} | W | \psi^{(j)} \rangle}$
603: involves multiplication by ${v_E/(4 L_i L_j)^{1/2}}$,
604: while the $\alpha_0$ is canceled out.
605: This leads to Eq.(\ref{e333}) for the ${i \ne j}$ coupling,
606: and a generalized expression for the energy level shifts.
607:
608:
609:
610: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
611: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
612: \section{Wigner decay and Bloch oscillations}
613: \label{S6}
614:
615:
616: If the two arms of the ring have exactly the same length ${L_1=L_2=L/2}$,
617: then the coherent Bloch oscillations of a wavepacket
618: in such a symmetric double well are characterized by the frequency
619: %
620: \be{0}
621: \Omega_{\tbox{Bloch}} = 2|W_{12}| = \frac{v_E}{L_1}\left|\sqrt{g_A}+\sqrt{g_B}\right|
622: \ee
623:
624:
625: If one arm of the ring ($L_1$) is very long,
626: and the other arm ($L_2$) is short,
627: then a particle placed initially at the short arm
628: will decay into the quasi continuum of the long arm.
629: The decay rate is given by the Fermi golden rule
630: %
631: \be{27}
632: \Gamma = \frac{2\pi}{\Delta}|W_{12}|^2
633: \ee
634: %
635: where $\Delta=(\pi/L)v_E$ is the mean level spacing.
636: If the arms are coupled through barrier~A,
637: while barrier~B is infinitely high,
638: then the decay rate is
639: %
640: \be{0}
641: \Gamma &=& \frac{v_E}{2L_2} \ g_A
642: \ee
643: %
644: This result agrees with the well known Gamow formula:
645: the decay rate is given by the attempt
646: frequency multiplied by the probability
647: to cross the barrier.
648:
649: If both barriers are finite, it is important to notice that
650: the quasi-continuum of the long arm is composed of odd and even states.
651: The state of the short arm, which is either even or odd, is coupled to states of
652: the same parity with a plus sign in the expression of Eq.(\ref{e7})
653: and to states of the opposite parity with a minus sign.
654: Accordingly, the decay rate is the sum of the decay rate to states of
655: the same parity and the decay rate to states of the opposite parity
656: %
657: \be{0}
658: \Gamma
659: \ \ &=& \ \
660: \sum_{\pm}
661: \frac{2\pi}{\Delta_{\pm}}
662: \left|
663: \frac{v_E}{2\sqrt{L_1L_2}}
664: \left(\sqrt{g_A} \pm\signC \sqrt{g_B}\right)
665: \right|^2
666: \nonumber\\
667: \ \ &=& \ \
668: \frac{v_E}{2L_2}\left(g_A+g_B\right)
669: \ee
670: %
671: where $\Delta_{\pm}=2\Delta$ is the mean level spacing
672: for states with the same parity. So inspite of the parity
673: considerations we still get the naive result that agree
674: with Gamow formula.
675:
676:
677: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
678: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
679: \section{Quantum stirring}
680: \label{S7}
681:
682: We assume that we have control over geometrical parameters
683: of the device, such as the potential floor in each arm,
684: the barriers heights, their location, or any other gate controlled
685: feature of the potential landscape. With a control parameter~$X$
686: we associate a generalized force operator
687: %
688: \be{21}
689: {\cal F} = -\frac{\partial{\cal H}}{\partial X}
690: \ee
691: %
692: Quantum stirring means to induce
693: a circulating current by changing
694: the parameter~$X$. We assume that the parametric variation is
695: adiabatic so we have a linear relation $\langle I \rangle = - G \dot{X}$,
696: where $G$ is know as the geometric conductance \citeref{pmc}.
697: The Kubo formalism implies that $G$ equals to the Berry curvature \citeref{berry1,avron2,berry2}:
698: %
699: \be{22}
700: G = \sum_{m(\neq n)}\frac{2 \ \im[{\cal I}_{nm}]{\cal F}_{mn}}{(E_m-E_n)^2}
701: \ee
702: %
703: where $n$ is the level in which the particle is prepared,
704: and $m$ are the other levels.
705:
706:
707:
708:
709: Within the framework of the TLS modeling the sum in Eq.(\ref{e22})
710: contains only one term which involves the states $n_0$ and $m_0$
711: of Eq.(\ref{e9}). For the matrix element of the current operator we get
712: %
713: \be{34}
714: {\cal I}_{n_0m_0}
715: = \left[\lambda\frac{\kappa}{2}\bm{\sigma}_y\right]_{n_0m_0}
716: = i\lambda \, \frac{\kappa}{2}
717: \ee
718: %
719: where $\lambda$ is the appropriate splitting ratio.
720: The matrix element of the generalized force
721: operator is calculated using Eqs.(\ref{e9}) and (\ref{e21})
722: %
723: \be{35}
724: {\cal F}_{m_0n_0}
725: &=&
726: -\frac{1}{2}
727: \left[
728: \frac{\partial \varepsilon}{\partial X}\bm{\sigma}_z
729: + \frac{\partial \kappa}{\partial X} \bm{\sigma}_x
730: \right]_{m_0n_0}
731: \\
732: &=&
733: \pm \ \frac{1}{2}\sin(\theta) \ \frac{\partial \varepsilon}{\partial X}
734: -\frac{1}{2}\cos(\theta) \ \frac{\partial \kappa}{\partial X}
735: \\
736: &=&
737: \frac{1}{2\Omega}
738: \left(
739: \kappa \frac{\partial \varepsilon}{\partial X}
740: - \varepsilon \frac{\partial \kappa}{\partial X}
741: \right)
742: \ee
743: %
744: where we used ${\pm\sin(\theta)=\kappa/\Omega}$
745: and ${\cos(\theta)=\varepsilon/\Omega}$.
746: This leads for the following result for the
747: geometric conductance:
748: %
749: \be{36}
750: G =
751: \frac{\lambda\kappa}{2\Omega^3}
752: \left[
753: \kappa \frac{\partial \varepsilon}{\partial X}
754: - \varepsilon \frac{\partial \kappa}{\partial X}
755: \right]
756: \ee
757: %
758: In the analysis of the operation of a stirring device
759: we typically have a well defined region
760: where the potential is being varied. We may call
761: this segment ``the pump". It is convenient to measure
762: the current elsewhere, where the potential is fixed.
763: If barrier~A is not part of the ``pump"
764: then we can measure the current at $x_0=x_A$.
765: Then it follows from the definitions of $\lambda$
766: and $\kappa$ that the product $\lambda\kappa$
767: does not change with time, even if barrier~B is modulated.
768: Then we can rewrite the above formula as
769: %
770: \be{0}
771: G =
772: \frac{\lambda_0\kappa_0}{2\Omega^3}
773: \left[
774: \kappa \frac{\partial \varepsilon}{\partial X}
775: - \varepsilon \frac{\partial \kappa}{\partial X}
776: \right]
777: \ee
778: %
779: where $\lambda_0$ and $\kappa_0$ are that values at
780: some arbitrary moment of time.
781: Typically the variation of $X$ leads to a sequence
782: of level crossings if $\kappa$ is disregarded.
783: These become avoided crossings if $\kappa$
784: is taken into account.
785: At the vicinity of a crossing we typically
786: can use a linear approximation:
787: %
788: \be{0}
789: \varepsilon
790: \ \ &=& \ \ \dot{\varepsilon} \times (X-X_0)
791: \\
792: \kappa
793: \ \ &=& \ \ \kappa_0 + \dot{\kappa} \times (X-X_0)
794: \ee
795: %
796: The amount of probability $dQ=Idt$ which is being
797: transported equals ${-GdX}$. For an individual crossing
798: the $dX$ integration over $G$ can be performed using
799: %
800: \be{0}
801: &&\int_{-\infty}^{+\infty}\frac{a(b+cx) \ \mbox{d}x}
802: {\left(a^2x^2+(b+cx)^2\right)^{3/2}}=\frac{2a}{b\sqrt{a^2+c^2}}\\
803: &&\int_{-\infty}^{+\infty}\frac{c \, ax \ \mbox{d}x}
804: {\left(a^2x^2+(b+cx)^2\right)^{3/2}}=-\frac{2c^2}{ab\sqrt{a^2+c^2}}
805: \ee
806: %
807: Then we get the result
808: %
809: \be{44}
810: Q \ \ = \ \ \pm \lambda_0 \sqrt{1+\left({\dot{\kappa}}/{\dot{\varepsilon}}\right)^2}
811: \ee
812: %
813: where the $\pm$ is determined according to the sign
814: of~$\dot{\varepsilon}$. We observed that in order to get
815: the ``quantized" value ${Q=1}$ there should be neither topological
816: splitting (${\lambda=1}$) nor barrier modulation (${\dot{\kappa}=0}$)
817: during the transition.
818:
819:
820: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
821: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
822: \section{The neighboring level approximation scheme}
823: \label{S8}
824:
825: A major interest is in systems with zero temperature Fermi occupation.
826: In such a case Eq.(\ref{e22}) has to be summed over $n$ up to
827: the Fermi energy. It turns out \citeref{pmx} that the result is dominated
828: by the contribution that come from the coupling between the last
829: occupied level and its neighboring empty level. This suggests
830: to adopt a neighboring level approximation scheme that holds
831: irrespective of the validity of the TLS modeling, and coincides
832: with it if the condition of Eq.(\ref{e11}) is satisfied.
833: The key idea is to characterize each eigenstate by a mixing parameter
834: %
835: \be{0}
836: \Theta \ \ \equiv \ \ 2\arctan\left(\sqrt{\frac{\mbox{Prob}(x\in2)}{\mbox{Prob}(x\in1)}}\right)
837: \ee
838: %
839: such that $\Theta=0$ for states that belong to the first arm
840: and $\Theta=\pi$ for states that belong to the second arm.
841: Numerical examples are presented in Figs.~3-4.
842: If we are given $\Theta$ then we can construct the eigenstate
843: using a procedure that we describe below. If the TLS modeling
844: applies then $\Theta$ becomes essentially the same as $\theta$.
845:
846:
847: Let us see how we construct the wavefunction given the
848: energy ${E=E_n}$, the mixing parameter ${\Theta=\Theta_n}$,
849: and the parity $\pm\signA$ with respect to (say) barrier~A,
850: as defined in Appendix~A. Consequently it is convenient
851: to set the origin such that ${x_A=0}$,
852: and write the $n$th eigenstate of the ring as
853: %
854: \be{37}
855: \psi^{(n)}(x) =
856: \left\{\amatrix{
857: \pm\signA C_1\sin(k_1x+\varphi_1) & \mbox{if $x \in 1$st arm} \cr
858: C_2\sin(k_2x+\varphi_2) & \mbox{if $x \in 2$nd arm}
859: } \right.
860: \ee
861: %
862: where $C_i>0$, and $|\varphi|<\pi/2$.
863: Assuming $k_EL\gg1$, the amplitudes satisfy the normalization condition
864: %
865: \be{0}
866: \frac{1}{2}L_1{C_1}^2+\frac{1}{2}L_2{C_2}^2 \approx 1
867: \ee
868: %
869: It follows that
870: %
871: \be{40}
872: C_1 &\approx& \sqrt{\frac{2}{L_1}} \ \cos\left(\frac{\Theta}{2}\right)\\
873: \label{e41}
874: C_2 &\approx& \sqrt{\frac{2}{L_2}} \ \sin\left(\frac{\Theta}{2}\right)
875: \ee
876: %
877: We still have to say what are
878: the wavenumbers $k_1$ and $k_2$,
879: and the phase shifts $\varphi$.
880: Let us see first how they are determined
881: within the framework of the TLS modeling,
882: and then how they can be found irrespective
883: of the TLS modeling.
884:
885:
886: Naively the $|n_0\rangle$ and $|m_0\rangle$ eigenstates,
887: within the framework of the TLS modeling,
888: are the superposition of the basis states
889: of Eq.(\ref{e4}) and accordingly
890: %
891: \be{42}
892: \Theta &=& \theta, \, \pi{-}\theta \\
893: \label{e443}
894: k_1 &=& \mbox{corresponds to the unperturbed $E_{1}$} \\
895: \label{e444}
896: k_2 &=& \mbox{corresponds to the unperturbed $E_{2}$} \\
897: \varphi_1 &=& \mbox{same as the unperturbed} \\
898: \varphi_2 &=& \mbox{same as the unperturbed}
899: \ee
900: %
901: while the true eigenstates are with (see Fig.2)
902: %
903: \be{47}
904: \Theta &\approx& \theta, \, \pi{-}\theta \\
905: k_1 &=& \mbox{corresponds to $E_n$} \\
906: k_2 &=& \mbox{corresponds to $E_n$} \\
907: \varphi_1 &=& \mbox{shifted} \\
908: \varphi_2 &=& \mbox{shifted}
909: \ee
910: %
911: To be more specific, we have $\Theta^{(m_0)} \approx \theta$
912: and $\Theta^{(n_0)} \approx \pi-\theta$
913: and hence ${\Theta^{(m_0)}{+}\Theta^{(n_0)}\approx \pi}$
914: if the TLS modeling is valid (see Fig.~4).
915: We note that from Eq.(\ref{e40}-\ref{e41}) it follows
916: that within the framework of the TLS approximation we have
917: %
918: \be{52}
919: C_i^{(m_0)}C_i^{(n_0)} \ \ \approx \ \ \frac{1}{L_i} \sin(\theta)
920: \ee
921: %
922: This will be used later on in order to obtain
923: simplified expressions for the matrix elements
924: of various operators.
925:
926:
927:
928: Whether $k_1$ and $k_2$ in Eqs.(\ref{e443}-\ref{e444}) correspond to the
929: same energy or not, is not a big difference
930: for us because we assume ${k_1 \sim k_2 \sim k_E}$
931: in any case. The main problem with the naive version
932: is related to the phase shifts, as demonstrated in Fig.~2.
933: The variation of the phase shift as
934: the barriers are lowered reflect that there
935: is a non-zero probability to find the particle
936: in the region of the barriers.
937: In particular if the phases $\varphi_i$ remained
938: the same it would imply that all the matrix
939: elements of $I^A$ and $I^B$ would be zero.
940: It is essential to take the variation of $\varphi$
941: into account in order to get a non-zero result
942: for the geometrical conductance.
943: We shall discuss the calculation of the matrix
944: elements $\mathcal{I}_{nm}$ and $\mathcal{F}_{nm}$
945: in the next sections. First we would like to discuss
946: how the required information on the variation
947: of the phases~$\varphi$ can be extracted.
948:
949:
950:
951: In order to express $\varphi$ by $\Theta$,
952: we write the wave function of Eq.(\ref{e37})
953: as ingoing and outgoing waves and set the origin $x=0$ at
954: either one of the barriers, for example barrier A.
955: We match the wave functions of the two bonds by
956: the barrier scattering matrix
957: %
958: \be{58}
959: %
960: \left(
961: \begin{array}{c}
962: {\pm\signA} C_1 \ \eexp{+i\varphi_1} \\
963: C_2 \ \eexp{+i\varphi_2}
964: \end{array}
965: \right)
966: %
967: &=& {\bm S}_A
968: %
969: \left(
970: \begin{array}{c}
971: {\pm\signA} C_1 \ \eexp{-i\varphi_1} \\
972: C_2 \ \eexp{-i\varphi_2}
973: \end{array}
974: \right)
975: %
976: \ee
977: %
978: and get closed equations for the phase shifts
979: %
980: \be{48}
981: \sqrt{1-g} \sin(2\varphi_1-\alpha-\gamma) {=}
982: 1-\frac{g}{2}\left(1+\left(\frac{C_2}{C_1}\right)^2\right)\\
983: \sqrt{1-g} \sin(2\varphi_2+\alpha-\gamma) {=}
984: 1-\frac{g}{2}\left(1+\left(\frac{C_1}{C_2}\right)^2\right)
985: \ee
986:
987:
988:
989:
990:
991: So far everything is exact. So once we have $\Theta$
992: we can find the phases and construct the wavefunction.
993: We would like to focus in the rest of this section
994: in the regime where the TLS modeling applies.
995: Assuming that $\Theta$ is determined by $\theta$
996: we want to find what are $\varphi_1$ and $\varphi_2$,
997: so as to construct a proper wavefunction.
998: Neglecting terms of order $g$ and expanding $\arcsin(1{-}x)$
999: as $\pi/2\pm\sqrt{2x}$ we obtain
1000: %
1001: \be{49}
1002: \varphi_1 &\approx& \frac{\gamma+\alpha}{2}+\frac{\pi}{4}
1003: \pm\signA\frac{\sqrt{g}}{2} \ \sqrt{\frac{L_1}{L_2}} \
1004: \tan\left(\frac{\Theta}{2}\right)
1005: \ee
1006: %
1007: where the $\pm\signA$ sign should be the same as in Eq.(\ref{e58}),
1008: which can be established by direct substitution.
1009: A similar expression can be obtained for $\varphi_2$.
1010: We note that within the framework of this TLS approximation we have
1011: %
1012: \be{53}
1013: \varphi_1^{(m_0)}-\varphi_1^{(n_0)} =
1014: \pm\sqrt{g_A} \left(\frac{L_1}{L_2}\right)^{1/2} \frac{1}{\sin(\theta)}
1015: \ee
1016: %
1017: where the sign is the same as that of $\kappa$.
1018: We now have all the building blocks needed
1019: for the calculation of the matrix elements.
1020:
1021:
1022: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1023: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1024: \section{The expression for $\mathcal{I}_{nm}$}
1025: \label{S9}
1026:
1027: If we adopt the TLS point of view,
1028: we can postulate a self-consistent definition
1029: of the current operator based on the continuity equation.
1030: For this purpose we define the occupation operator ${\cal N}$
1031: for one of the arms as
1032: %
1033: \be{0}
1034: {\cal N}
1035: \ \ = \ \
1036: \left(
1037: \begin{array}{cc}
1038: 1 & 0 \\
1039: 0 & 0
1040: \end{array}
1041: \right)
1042: \ee
1043: %
1044: and deduce the definition of the current operator from
1045: %
1046: \be{0}
1047: \frac{d}{dt}{\cal N}
1048: \ \ = \ \
1049: i[\mathcal{H},\mathcal{N}]
1050: \ \ \equiv \ \ {\cal I}
1051: \ee
1052: %
1053: where ${\cal I}$ is given by Eq.(\ref{e10}).
1054: If we turn off the coupling at barrier~A
1055: we get the same expression multiplied by $\lambda_B$,
1056: while if we turn off the coupling at barrier~B
1057: we get the same expression multiplied by $\lambda_A$.
1058:
1059: The above reasoning bypass the confrontation which is
1060: involved in carrying out a direct calculation,
1061: and hence contains an uncontrolled error
1062: which is associated with the assumption that a TLS
1063: description of Hilbert space is valid.
1064: %
1065: If we revert to the original definition of Eq.(\ref{e2}),
1066: then the matrix elements are given by
1067: %
1068: \be{0}
1069: {\cal I}_{nm} = i\frac{1}{2\mass}\left(
1070: \partial\psi^{(n)} \ \psi^{(m)}-\psi^{(n)} \ \partial\psi^{(m)}\right)
1071: \Big|_{x=x_0}
1072: \ee
1073: %
1074: For the calculation of ${\cal I}^A_{nm}$ we set $x_0=x_A=0$.
1075: As was already pointed out, in order to get a non-trivial
1076: result, we have to take into account the phase shifts $\varphi$
1077: which was calculated in the previous section.
1078: Substituting the wave function of Eq.(\ref{e37}) we get
1079: %
1080: \be{61}
1081: {\cal I}^A_{nm} &=&
1082: -i\frac{1}{2\mass}C_1^{(m)}C_1^{(n)}\left[
1083: \frac{k_{m}+k_{n}}{2} \ \sin(\varphi_1^{(m)}-\varphi_1^{(n)})
1084: \right.\nonumber\\
1085: &&+ \left.\frac{k_{n}-k_{m}}{2} \ \sin(\varphi_1^{(m)}+\varphi_1^{(n)})
1086: \right]
1087: \ee
1088: %
1089: Whenever the TLS modeling applies we can substitute
1090: Eqs.(\ref{e52}) and (\ref{e53}) into Eq.(\ref{e61}).
1091: Neglecting the second term we get
1092: %
1093: \be{70}
1094: {\cal I}^A_{n_0m_0} \approx {\mp}i\frac{v_{\tbox{E}}}{2\sqrt{L_1L_2}}\sqrt{g_A}
1095: \ee
1096: %
1097: where the sign is the same as that of $-\kappa$.
1098: One notices that the expression for ${\cal I}^A_{n_0m_0}$
1099: can be written as Eq.(\ref{e34}) where $\kappa$
1100: and $\lambda$ are given by Eq.(\ref{e7}) and Eq.(\ref{e12}).
1101:
1102:
1103:
1104:
1105:
1106: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1107: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1108: \section{Stirring by barrier modulation}
1109: \label{S11}
1110:
1111:
1112: In this section we calculate the geometric conductance
1113: as determined by the matrix elements of the generalized force
1114: that is associated with modulation
1115: of a delta barrier.
1116: The motivation is to verify the results of the reduced
1117: description against the direct full Hilbert space calculation.
1118: The potential barrier is given by
1119: %
1120: \be{0}
1121: V_B(\hat{x}) = X_B\delta(\hat{x}-x_B)
1122: \ee
1123: %
1124: The stirring is induced by variation of the barrier
1125: height $X_B$. The associated generalized force is
1126: %
1127: \be{0}
1128: {\cal F} = -\frac{\partial{\cal H}}{\partial X_B}
1129: \ \ = \ \ -\delta(\hat{x}-x_B)
1130: \ee
1131: %
1132: with the matrix elements
1133: %
1134: \be{0}
1135: {\cal F}_{mn} \ \ = \ \ -\psi^{(n)}\psi^{(m)}
1136: \ee
1137: %
1138: For the wavefunctions amplitudes we use Eq.(\ref{e52})
1139: and for the phase shifts Eq.(\ref{e49}).
1140: We also substitute the scattering matrix parameters
1141: that describe a delta barrier
1142: %
1143: \be{0}
1144: \gamma_B &\approx& -\pi/2+\sqrt{g_B}\\
1145: \alpha_B &=& 0
1146: \ee
1147: %
1148: where the approximation is valid for $g_B\ll1$ and
1149: the relation of $g_B$ and $X_B$ is given in Eq.(\ref{e18}).
1150: With the above approximations we get
1151: %
1152: \be{0}
1153: {\cal F}_{m_0n_0} \approx g_B\frac{L_2-L_1}{4L_1L_2}\sin(\theta) \ {\mp} \
1154: \frac{g_B}{2\sqrt{L_1L_2}}\cos(\theta)
1155: \ee
1156: %
1157: where the sign should be the same as that of $\mp\signC\kappa$.
1158: In order to verify the consistency with the TLS expression,
1159: we differentiate Eq.(\ref{e6}) and Eq.(\ref{e7}):
1160: %
1161: \be{0}
1162: \frac{\partial \varepsilon}{\partial X_B} &=&
1163: g_B\frac{L_2-L_1}{2L_1L_2}\\
1164: \frac{\partial \kappa}{\partial X_B} &=&
1165: \pm\signC\frac{g_B}{\sqrt{L_1L_2}}
1166: \ee
1167: %
1168: and substitute into Eq.(\ref{e35}).
1169: Indeed we obtain the same result
1170: for ${\cal F}_{m_0n_0}$ as above.
1171:
1172:
1173: The geometric conductance of Eq.(\ref{e22}) involves
1174: the multiplication of ${\cal F}_{m_0n_0}$ with ${\cal I}_{n_0m_0}$, leading to
1175: %
1176: \be{0}
1177: G &=& \frac{1}{4} \ v_{E}^2 \frac{L_2-L_1}
1178: {\left(L_1L_2\right)^2} \ \frac{g_A \ g_B\pm\signC{g_A}^{1/2}{g_B}^{3/2}}{\Omega^3}
1179: \nonumber\\
1180: &\mp\signC& \frac{1}{4} \ v_{E}^2 \frac{L_2+L_1}
1181: {\left(L_1L_2\right)^2} \ \frac{g_A \ g_B+{g_A}^{1/2}{g_B}^{3/2}}{\Omega^3}
1182: \ee
1183: %
1184: The calculation of the transport proceeds as in Section~\ref{S7}.
1185:
1186:
1187:
1188: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1189: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1190: \section{Stirring by barrier translation}
1191: \label{S10}
1192:
1193:
1194: In complete analogy with the previous section we
1195: would like to calculate the geometric conductance
1196: as determined by the matrix elements
1197: of the generalized force which is associated
1198: with the translation of the barrier:
1199: %
1200: \be{0}
1201: {\cal F} = -\frac{\partial{\cal H}}{\partial x_B} \ = \
1202: X_B\delta'(\hat{x}-x_B)
1203: \ee
1204: %
1205: One obtains
1206: %
1207: \be{0}
1208: {\cal F}_{mn} =
1209: -X_B\left(\overline{\partial\psi^{(n)}} \ \psi^{(m)}+
1210: \overline{\partial\psi^{(m)}} \ \psi^{(n)}\right)
1211: \ee
1212: %
1213: where $\overline{\partial\psi}$ is the
1214: average derivative on both sides of the barrier.
1215: We simplify this expression by using Eq.(\ref{e16}):
1216: %
1217: \be{0}
1218: {\cal F}_{mn}=\frac{1}{2\mass}
1219: \left[\partial\psi_1^{(n)}\partial\psi_1^{(m)}-
1220: \partial\psi_2^{(n)}\partial\psi_2^{(m)}
1221: \right]_{x=x_B}
1222: \ee
1223: %
1224: Assuming high barriers we get in leading order
1225: %
1226: \be{0}
1227: {\cal F}_{mn} \approx -\frac{1}{2}\mass v_{E}^2\left( C_1^{(m)}C_1^{(n)} + C_2^{(m)}C_2^{(n)}\right)
1228: \ee
1229: %
1230: which together with Eq.(\ref{e52}) leads to
1231: %
1232: \be{0}
1233: {\cal F}_{m_0n_0} \approx -\frac{1}{2}\mass v_{E}^2 \ \frac{L_1+L_2}{L_1L_2} \ \sin(\theta)
1234: \ee
1235: %
1236: In order to compare the above result for ${\cal F}_{m_0n_0}$
1237: with the TLS result of Eq.\ref{e35}
1238: we calculate the variation of the potential floor
1239: by taking in Eq.(\ref{e6}) the energies of infinite wells
1240: with $L_1=x_B$ and $L_2=L-x_B$. We get
1241: %
1242: \be{0}
1243: \frac{\partial \varepsilon}{\partial x_B} &=& -\mass v_E^2
1244: \ \frac{L_1+L_2}{L_1L_2} +\mathcal{O}(\sqrt{g})\\
1245: \frac{\partial \kappa}{\partial x_B} &=& v_E
1246: \ \frac{L_2-L_1}{\left(L_1L_2\right)^{3/2}}
1247: \left(\sqrt{g_A}\pm\signC\sqrt{g_B}\right)
1248: \ee
1249: %
1250: Substitute into Eq.(\ref{e35}) indeed leads
1251: to the same result for ${\cal F}_{m_0n_0}$ as above.
1252: Note that in this case (unlike the previous section)
1253: the second term in Eq.(\ref{e35}) which involves
1254: the variation of $\kappa$ is of higher order in $g_B$
1255: and therefore should be excluded.
1256:
1257:
1258: The geometric conductance of Eq.(\ref{e22}) involves
1259: the multiplication of ${\cal F}_{m_0n_0}$ with ${\cal I}_{n_0m_0}$, leading to
1260: %
1261: \be{0}
1262: G \ \ = \ \ -\frac{1}{2}\sqrt{g_{A}} \ \mass v_{E}^4 \
1263: \frac{L_1+L_2}{\left(L_1L_2\right)^2} \ \frac{\sqrt{g_A}\pm\signC\sqrt{g_B}}{\Omega^3}
1264: \ee
1265: %
1266: The calculation of the transport proceeds as in Section~\ref{S7}.
1267: One realizes that a translation of the barrier
1268: is effectively equivalent to the variation of the potential floor difference,
1269: as long as it does not involve modulation of its transmission
1270: (which is assumed to be small).
1271:
1272:
1273: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1274: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1275: \section{Error estimates and limitations}
1276: \label{S13}
1277:
1278: If we vary a parameter $X$ then the energy levels $E_n(X)$ form
1279: a ``spaghetti" which is characterized by a mean level spacing~$\Delta$
1280: and possibly by narrow avoided crossings with splitting~$\Delta_0$.
1281: For the ring system that we are considering
1282: it follows from the estimate of $\kappa$ that
1283: %
1284: \be{0}
1285: \frac{\Delta_0}{\Delta} \ \ \sim \ \ \mbox{min}\{1, \sqrt{bg}\}
1286: \ee
1287: %
1288: where $b=L_1/L_2$ and $g=\mbox{max}\{g_A,g_B\}$.
1289: The condition Eq.(\ref{e11}) for the applicability
1290: of the TLS modeling ensures $\Delta_0\ll\Delta$.
1291: In such circumstances Eq.(\ref{e22}) for the geometric
1292: conductance, which in essence is a sum
1293: of the type ${\sum_{n=0}^{\infty} (\Delta_0+ n\Delta)^{-2}}$,
1294: implies that the error that is involved
1295: in the neighboring level approximation is
1296: %
1297: \be{0}
1298: \frac{\mbox{error}(G)}{G}
1299: \ \ \sim \ \ \left( \frac{\Delta_0}{\Delta} \right)^2
1300: \ \ \sim \ \ bg \ \ \ll 1
1301: \ee
1302: %
1303: Once the TLS modeling fails the error becomes of order unity.
1304: This sounds bad, but in fact it is not so bad.
1305: The good news is that the far levels contribute to~$G$
1306: a correction which is of the same order as the leading term.
1307: Therefore with the neighboring level approximation
1308: we can still get a realistic estimate disregarding numerical
1309: prefactors of order unity.
1310:
1311:
1312: Having $b\gg1$ is very interesting, because then we have
1313: a non-trivial intermediate regime ${1/b \ll g \ll 1}$
1314: where neither 1st order perturbation theory with respect
1315: to ``zero" height barriers, nor 1st order perturbation
1316: theory with respect to ``infinite" barriers applies.
1317: This is the regime where each level of the small arm
1318: forms a distinct Wigner resonance with the
1319: quasi-continuum states of the long arm.
1320: Obviously the TLS modeling is not applicable in this regime,
1321: but the neighboring level approximation still provides
1322: a decent starting point for a calculation. We shall
1323: explore this Wigner regime in a future work.
1324:
1325:
1326: One may also wonder whether the specific results that we have
1327: obtained for stirring using a {\em delta} barrier applies also
1328: for a {\em thick} barrier. On physical grounds it is quite
1329: obvious that the induced current is determined by
1330: the scattering matrix of the modulated barrier.
1331: Consequently if the $S(E)$ of the modulated barrier is $E$~independent
1332: within the energy range of interest, it can be regarded
1333: as representing a delta function, and the results should come out the same.
1334:
1335:
1336: Finally one may wonder about the implications of finite
1337: temperature or non-adiabatic driving. These aspects are
1338: complementary to the theme of the present paper.
1339: Namely, as discussed in \citeref{pmx}, at finite temperatures
1340: the statistics of the occupation should be taken into account.
1341: So we have to average (so to say) over the level that we have
1342: labeled as~$n_0$ with an appropriate weight as implied by
1343: the Fermi function. On the other hand the non-adiabatic effects
1344: require to introduce in the denominator of the
1345: Kubo formula Eq.(\ref{e22}) a term that represents
1346: the ``width" of the Fermi-golden-rule transitions.
1347: Then the weight of the neighboring level in the sum becomes smaller
1348: compared with the total weight of the far levels.
1349:
1350:
1351:
1352: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1353: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1354: \section{Summary}
1355: \label{S12}
1356:
1357: We have developed a practical procedure for the
1358: analysis of a one dimensional double well system,
1359: which is both powerful and illuminating.
1360: The procedure assumes that we have a way to find
1361: the eign-energies $E_n$ of the device,
1362: and the mixing ratio $\Theta_n$ of each of them.
1363: Given the transmissions of the barriers
1364: we further characterize the device by
1365: the splitting ratio $\lambda$. With these ingredients
1366: in hand we can analyze any stirring process
1367: and obtain explicit expressions for the geometric
1368: conductance~$G$. The calculation simplifies
1369: if the TLS modeling applies, because then the
1370: mixing ratio can be determined form
1371: the diagonalization of a $2\times 2$ matrix.
1372:
1373:
1374: In particular we obtain explicit expressions
1375: for~$G$ due to either barrier translation
1376: (generalizing a result that has been obtained in~\citeref{pmx}),
1377: or barrier modulation
1378: (generalizing a result that has been obtained in~\citeref{pms}),
1379: and verify that they agree
1380: with the naive self-consistent TLS calculation.
1381: We see that whenever the TLS modeling applies
1382: the proper calculation in the full Hilbert space
1383: gives the same result as the naive calculation
1384: in the TLS Hilbert space.
1385:
1386:
1387: As a by product of the TLS analysis we find that
1388: the pumped ``charge" during an avoided crossing
1389: is not quantized (see Eq.(\ref{e44})),
1390: not only because of the topological splitting effect,
1391: but also due to a dynamical effect
1392: that arises if the barrier is modulated.
1393:
1394:
1395:
1396: The practical importance of the TLS modeling
1397: in condense matter physics is obvious.
1398: On the other hand the specific application to
1399: the study of quantum stirring deserves
1400: a few words regarding the measurement procedure
1401: and the experimental relevance. As explained
1402: in \citeref{pmx} it should be clear that the measurement
1403: of current in a closed circuit requires special
1404: techniques \citeref{orsay,expr1,expr2}.
1405: %
1406: These techniques are typically used in order
1407: to probe persistent currents,
1408: which are zero order (conservative) effect,
1409: while in the present paper we were discussing driven
1410: currents, which are a first-order (geometric) effect.
1411: It is of course also possible to measure the dissipative
1412: conductance (as in~\citeref{orsay}).
1413: %
1414: During the measurement the coupling to the system
1415: should be small. These are so called {\em weak measurement}
1416: conditions. More ambitious would be to measure
1417: the counting statistics, i.e. also the second moment
1418: of $Q$ as discussed in \citeref{cnb,cnz} which is
1419: completely analogous to the discussion of noise measurements
1420: in open systems \citeref{levitov,nazarov}.
1421: %
1422: Finally it should be pointed out that the formalism
1423: above, and hence the results, might apply to
1424: experiments with superconducting circuits (see \citeref{JJ}).
1425:
1426:
1427:
1428:
1429: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1430: \appendix
1431: \section{Conventions and notations}
1432:
1433: Consider two segments that are connected
1434: at points that are labeled as $x_A$ and $x_B$.
1435: In the absence of coupling each segment is
1436: regarded as a one dimensional box.
1437: The unperturbed eigenstates are labeled by~$i$
1438: (or optionally by~$j$). In the TLS scheme ${i=1,2}$.
1439: If the coupling is non-zero the exact eigenstates
1440: are labeled by~$n$ (or optionally by $m$).
1441: Within the framework of the neighboring level
1442: approximation scheme we focus on two levels
1443: that we label as ${n=n_0}$ and ${m=m_0}$.
1444: If the TLS modeling applies then the states $n_0$ and $m_0$
1445: are regarded as linear combinations of ${i=1,2}$.
1446:
1447: The unperturbed states ${i=1,2}$ are characterized
1448: by their parities $\pm^{\!\!1}$ and $\pm^{\!\!2}$ respectively.
1449: The relative sign $\pm\signC$ in Eq.(\ref{e7})
1450: equals the product of $\pm^{\!\!1}$ and $\pm^{\!\!2}$.
1451: %
1452: Inverting the arbitrary gauge sign
1453: of either $\psi^{(1)}(x)$ or $\psi^{(2)}(x)$
1454: would multiply the expression in Eq.(\ref{e333})
1455: by a global minus sign, while the relative
1456: sign $\pm\signC$ remains unchanged.
1457: The gauge invariant relative sign is due to the fact that
1458: the unperturbed states are either odd or even:
1459: we have plus sign if both states have the
1460: same parity and minus sign if they have opposite parity.
1461:
1462:
1463: Each exact eigenfunction~$n$, as written as in Eq.(\ref{e37}),
1464: is characterized by what we call the parity $\pm\signA$
1465: with respect to barrier~A. Positive parity means that
1466: the radial derivatives as defined in Eq.(\ref{e16})
1467: have both the same sign. Optionally we can define $\pm^{\!\!b}$
1468: as the parity with respect to barrier~B.
1469: %
1470: This parity $\pm\signA$ is not a symmetry related quantum number,
1471: but it is merely required in order to define the wavefunction of Eq.(\ref{e37})
1472: in a unique way given the energy and the mixing ratio.
1473: If the TLS modeling applies then for positive (negative) $\kappa$
1474: the state $n_0$ of Eq.(\ref{e9}) has negative (positive) parity,
1475: while the $m_0$ state has positive (negative) parity.
1476: Within this framework the parity $\pm^{\!\!b}$ with respect
1477: to barrier~B is $\pm\signA$ multiplied by $\pm\signC$.
1478:
1479: We have verified that the various $\pm$ signs through the
1480: paper are consistent, which is not always evident in a superficial look.
1481:
1482:
1483:
1484:
1485:
1486: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1487: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1488:
1489: \acknowledgments
1490:
1491: This research was supported by grants from
1492: the USA-Israel Binational Science Foundation (BSF),
1493: and from the Deutsch-Israelische Projektkooperation (DIP).
1494:
1495:
1496:
1497:
1498: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1499: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1500: \newpage
1501: \begin{thebibliography}{99}
1502:
1503:
1504: %%Quantum stirring
1505:
1506: \bibitem{pmx}
1507: The most recent publication in this line of study is:
1508: I. Sela and D. Cohen, Phys. Rev. B {\bf 77}, 245440 (2008).
1509: For older references see there.
1510:
1511:
1512: %%Quantum pumping
1513:
1514: \bibitem{Thouless}
1515: D. J. Thouless, Phys. Rev. B {\bf 27}, 6083 (1983).
1516:
1517: \bibitem{bpt}
1518: M. Buttiker, H. Thomas and A. Pretre, Z. Phys. B {\bf 94}, 133 (1994).
1519:
1520: \bibitem{BPT2}
1521: P. W. Brouwer, Phys. Rev. B {\bf 58}, 10135 (1998).
1522:
1523: \bibitem{Avron}
1524: J. E. Avron, A. Elgart, G. M. Graf and L. Sadun, Phys. Rev. B {\bf 62}, 10618 (2000).
1525:
1526: \bibitem{pMB}
1527: M.~Moskalets and M.~B{\"u}ttiker,
1528: Phys. Rev. B {\bf 68}, 161311 (2003).
1529:
1530: \bibitem{pms}
1531: I. Sela and D. Cohen, J. Phys. A {\bf 39}, 3575 (2006).
1532:
1533:
1534:
1535: %%Kubo
1536:
1537: \bibitem{pmc}
1538: D. Cohen, Phys. Rev. B {\bf 68}, 155303 (2003).
1539:
1540: \bibitem{berry1}
1541: M.V. Berry, Proc. R. Soc. Lond. A {\bf 392}, 45 (1984).
1542:
1543: \bibitem{avron2}
1544: J. E. Avron, A. Raveh and B. Zur, Rev. Mod. Phys. {\bf 60}, 873 (1988).
1545:
1546: \bibitem{berry2}
1547: M.V. Berry and J.M. Robbins, Proc. R. Soc. Lond. A {\bf 442}, 659 (1993).
1548:
1549:
1550:
1551:
1552: % measurement
1553:
1554: \bibitem{orsay}
1555: Measurements of currents in arrays
1556: of closed rings are described by: \
1557: B. Reulet M. Ramin, H. Bouchiat and D. Mailly,
1558: Phys. Rev. Lett. {\bf 75}, 124 (1995).
1559:
1560: \bibitem{expr1}
1561: Measurements of currents in individual closed rings
1562: using SQUID is described in: \
1563: N.C. Koshnick, H. Bluhm, M.E. Huber, K.A. Moler, Science 318, 1440 (2007).
1564:
1565: \bibitem{expr2}
1566: A new micromechanical cantilevers technique for measuring currents
1567: in closed rings is described in:
1568: A.C. Bleszynski-Jayich, W.E. Shanks, R. Ilic, J.G.E. Harris, arXiv:0710.5259,
1569: Journal of Vacuum Science \& Technology B {\bf 26}, 1412 (2008).
1570:
1571:
1572: %measure
1573:
1574: \bibitem{cnb}
1575: M. Chuchem and D. Cohen, J. Phys. A {\bf 41}, 075302 (2008).
1576:
1577: \bibitem{cnz}
1578: M. Chuchem and D. Cohen, Phys. Rev. A {\bf 77}, 012109 (2008).
1579:
1580: \bibitem{levitov}
1581: L.S. Levitov and G.B. Lesovik, JETP Letters {\bf 58}, 230 (1993).
1582:
1583: \bibitem{nazarov}
1584: Y.V. Nazarov and M. Kindermann, European Physical Journal B {\bf 35}, 413 (2003).
1585:
1586: \bibitem{JJ}
1587: M. Mottonen, J. P. Pekola, J. J. Vartiainen,
1588: V. Brosco and F. W. J. Hekking, Phys. Rev. B {\bf 73}, 214523 (2006).
1589:
1590:
1591:
1592: \end{thebibliography}
1593:
1594: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1595: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1596:
1597: \clearpage
1598:
1599:
1600: \begin{figure}[h]
1601:
1602: \includegraphics[width=\hsize]{pmdring}
1603:
1604: \caption{
1605: {\em Panel~(a)}: Illustration of a ring shaped device
1606: that is divided by the barriers~$V_A$ and~$V_B$
1607: into two arms of length $L_1$ and $L_2$.
1608: The current is measured through the dashed section
1609: near barrier~A. In the quantum stirring scenario
1610: it is assumed that there is a gate control
1611: over the potential floor of each arm,
1612: or over the height or the location of barrier~B.
1613: {\em Panel~(b)}:
1614: Within the framework of the TLS modeling, the
1615: reduced Hilbert space contains two levels.
1616: The perturbation $W_{ij}$ is due to having finite
1617: rather than infinite barriers, so it corresponds
1618: to the difference ${\mathcal{H}-\mathcal{H}(\infty)}$
1619: and not to ${V = \mathcal{H}-\mathcal{H}(0)}$.
1620: See the text for further details.}
1621: \end{figure}
1622:
1623:
1624:
1625:
1626: \begin{figure}[h]
1627:
1628: \includegraphics[width=0.6\hsize]{pmd_Fig3b}
1629: \includegraphics[width=0.6\hsize]{pmd_Fig3a}
1630:
1631: \caption{
1632: {\em Upper panel:} Two nearly degenerate eigenfunctions $\psi(x)$ of a
1633: particle in a ring with arms of length ${L_1=1}$ and ${L_2=2.23}$.
1634: These are the two unperturbed states of Eq.(\ref{e4}).
1635: {\em Lower panel:} The exact eigenfunctions
1636: assuming that the barriers are finite (${g_A\approx0.28}$ and ${g_B\approx0.06}$).
1637: These do not vanish at the barriers,
1638: and therefore cannot be written as a superposition of
1639: the unperturbed states. Still we explain in the text
1640: how a decent approximation for the former can be obtained
1641: using the neighboring levels approximation scheme.}
1642: \end{figure}
1643:
1644:
1645:
1646:
1647:
1648: \begin{figure}[h]
1649:
1650: \includegraphics[width=0.7\hsize]{pmd_Fig6_a}
1651: \includegraphics[width=0.7\hsize]{pmd_Fig6_b}
1652:
1653: \caption{
1654: We consider a particle of mass ${\mass=1}$ in a ring
1655: of length $L=151.43$. The position of barrier~B is $X$,
1656: so we have ${L_1=X}$ and ${L_2=L-X}$.
1657: We calculate numerically $k_n$ and $\Theta^{(n)}$ for two
1658: neighboring levels (solid and dashed lines).
1659: The sum $\Theta^{(m)}{+}\Theta^{(n)}$ is plotted as a dash-dotted line.
1660: We have high barriers with ${g_A\sim10^{-2}}$ and ${g_B\sim10^{-5}}$.
1661: Accordingly we expect TLS modeling to be valid:
1662: The dotted lines indicate the values ${\Theta=\pi/2}$ (expected crossing point)
1663: and ${\Theta=\pi}$ (expected sum). For sake of comparison
1664: there is a third dotted line that indicates
1665: the value of $\Theta$ that corresponds to equal amplitudes ${C_1=C_2}$.
1666: }
1667: \end{figure}
1668:
1669:
1670: \begin{figure}[h]
1671:
1672: \includegraphics[width=0.49\hsize]{pmd_Fig6_c}
1673: \includegraphics[width=0.49\hsize]{pmd_Fig6_e}
1674: \\
1675: \includegraphics[width=0.49\hsize]{pmd_Fig6_d}
1676: \includegraphics[width=0.49\hsize]{pmd_Fig6_f}
1677:
1678: \caption{
1679: The same as the previous figure, but here
1680: the TLS modeling does not apply.
1681: In the left panels
1682: one barrier is high (${g_A\sim10^{-2}}$)
1683: and one barrier is low (${g_B\sim0.9}$),
1684: while in the right panels
1685: both barrier are low (${g_A\sim g_B\sim0.9}$).
1686: }
1687: \end{figure}
1688:
1689:
1690:
1691: \clearpage
1692:
1693:
1694: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1695: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1696: \end{document}
1697: