1: \documentstyle[multicol,aps,epsf,here]{revtex}
2: \tighten
3: \newcommand{\PostScript}[7]{
4: \begin{figure}[H]
5: \begin{center}
6: \leavevmode
7: \epsfysize=#1cm
8: \vspace{#2cm}
9: \epsfbox{#3}
10: \par
11: \parbox{#5cm}{
12: \vspace{#4cm}
13: \caption[figure]{\renewcommand{\baselinestretch}{1} \small \normalsize #6}
14: \label{#7}}
15: \end{center}
16: \end{figure}
17: }
18:
19: \begin{document}
20:
21: \newcommand{\half}{\frac {1}{2} }
22: \newcommand{\eg}{{\em e.g.} }
23: \newcommand{\ie}{{\em i.e.} }
24: \newcommand{\etc}{{\em etc.}}
25: \newcommand{\si}{\simeq}
26: \newcommand{\etal}{{\em et al.\ }}
27: \newcommand{\cf}{{\em cf. }}
28:
29: \newcommand{\dd}[2]{{\rmd{#1}\over\rmd{#2}}}
30: \newcommand{\pdd}[2]{{\partial{#1}\over\partial{#2}}}
31: \newcommand{\pa}[1]{\partial_{#1}}
32: \newcommand{\pref}[1]{(\ref{#1})}
33:
34: \newcommand {\be}[1]{
35: %% For draft %%%%%%%%
36: %{\marginpar{{\scriptsize\ \\ \ #1}}}
37: \begin{eqnarray} \mbox{$\label{#1}$}}
38:
39: \newcommand{\ee}{\end{eqnarray}}
40:
41: \def\rdown{\rho_{\downarrow}}
42: \def\pa{\partial}
43: \def\th{\theta}
44: \def\bea{\begin{eqnarray}}
45: \def\eea{\end{eqnarray}}
46: \def\be{\begin{equation}}
47: \def\ee{\end{equation}}
48: \def\pa{\partial}
49: \def\eps{\epsilon}
50: \def\th{\theta}
51: \def\na{\nabla}
52: \def\nn{\nonumber}
53: \def\lan{\langle}
54: \def\ran{\rangle}
55: \def\pr{\prime}
56: \def\rarrow{\rightarrow}
57: \def\larrow{\leftarrow}
58:
59: %%%%%%%%%%%%%%%%%%%%%
60: %\vspace* {-21 mm}
61: %\begin{flushright}
62: % NORDITA-2000/XX CM \\
63: % April-2000 \\
64: %\end{flushright}
65: %%%%%%%%%%%%%%%%%%%%
66:
67: \title{Scattering phase shifts in quasi-one-dimension}
68: \vskip 20 mm
69: \author{P. Singha Deo and Swarnali Bandopadhyay}
70: \address{S.N. Bose National Centre for Basic Sciences,
71: J.D.Block, Sector III, Salt Lake City, Calcutta 700098, India}
72: \author{Sourin Das}
73: \address{Harish Chandra Research Institute, Chhatnag Road, Jhusi,
74: Allahabad 211019, India }
75: \date{\today}
76:
77: \maketitle
78:
79: \begin{abstract}
80:
81: Scattering of an electron in quasi-one dimensional quantum wires have many
82: unusual features, not found in one, two or three dimensions. In this work
83: we analyze the scattering phase shifts due to
84: an impurity in a multi-channel quantum wire
85: with special emphasis on negative slopes in the scattering phase shift
86: versus incident energy curves and the Wigner delay time.
87: Although at first sight, the large number
88: of scattering matrix elements show phase shifts of different character and
89: nature, it is possible to see some pattern and understand these features.
90: The behavior of scattering phase shifts in one-dimension can be seen as a
91: special case of these features observed in quasi-one-dimensions.
92: The negative slopes can occur at any arbitrary energy and
93: Friedel sum rule is completely violated in
94: quasi-one-dimension at any arbitrary energy and any arbitrary regime.
95: This is in contrast to one, two or
96: three dimensions where such negative slopes and violation of
97: Friedel sum rule happen only at low energy where
98: the incident electron feels the potential very strongly (i.e.,
99: there is a very well defined regime, the WKB regime, where
100: FSR works very well).
101: There are some novel
102: behavior of scattering phase shifts at the critical energies where
103: $S$-matrix changes dimension.\\
104:
105: \hfill\\
106: PACS: 73.23.-b, 72.10.-d, 72.10.Bg
107: \end{abstract}
108: \begin{multicols}{2}
109:
110: \narrowtext
111: %%%%%%%%%%%%%%%%%%%%%%
112: \section{Introduction}
113: \label{s1}
114: %%%%%%%%%%%%%%%%%%%%%%
115:
116: Elastic scattering in one, two and three dimensions is well understood
117: \cite{new82}. In a scattering process there are some important physical
118: quantities like scattering amplitude and scattering phase shift. While the
119: scattering intensity is directly related to the scattering amplitude, the
120: scattering phase shifts are also very important physical quantities and
121: the Friedel-sum-rule (FSR) relates them to the density of states (DOS).
122:
123: %%%
124: \PostScript{6}{0}{fig1.ps}{0.5}{14}{
125: \hspace{-.5cm}
126: A rectangular quantum biliard or quantum dot, weakly coupled to leads.
127: The dotted line is along the x-axis and the dashed line is along
128: the y-axis.
129: }{f2}
130: %%%
131:
132: At low temperatures, inelastic collisions are greatly suppressed. As a
133: result the phase coherence length of an electron can become a few microns.
134: Mesoscopic systems are defined as systems in which the phase coherence
135: length exceeds the sample size. In such a system elastic scattering is the
136: dominant feature and such mesoscopic samples can be understood as phase
137: coherent elastic scatterers. With the experimental realizations of
138: mesoscopic systems and their possibility of being applied in
139: nano-technology and quantum computing, understanding scattering effects in
140: quasi-one dimensions (Q1D) has also become important at present. This is
141: also essential because the Landauer conductance formula relates the
142: conductance to partial scattering intensities and also one can now probe
143: scattering phase shift directly in an experiment \cite{yac95,sch97,CER97}.
144: Mesoscopic samples are normally made up of metals or semiconductors, and
145: the defects in them are generally point defects. Hence we will restrict
146: our analysis to delta function potential impurities.
147:
148: Recently a new kind of scattering phase shift was discussed in Q1D, in
149: connection with the violation of the parity effect \cite{deo96}. To
150: explain this phase shift and the violation of parity effect, here we
151: elaborate some portions of Ref. \cite{deo96} and explain what are symmetry
152: dictated nodes (SDN) and non-symmetry dictated nodes (NSDN) that can arise
153: in a Q1D system. For example let us consider a rectangular quantum
154: biliard or dot connected to leads by quantum mechanical tunneling as
155: shown in Fig.1. The system has reflection symmetry across the x-axis as
156: well as the y-axis. Also the x-y components separate and spanning nodes
157: (nodes that span across the direction of propagation and shown by dashed
158: line in Fig.1) as well as non-spanning nodes (shown by dotted line in
159: Fig.1) develop in the geometry, dictated by the reflection symmetries.
160: There are various symmetries that give rise to nodes in the wave-function
161: and we call them SDN (for example the antisymmetric property of the many
162: body wave function result in nodes). But if quantum coherence extends to
163: some distance inside the leads then one can model the phase coherent
164: quantum dot as shown in Fig.2.
165:
166: %%%
167: \PostScript{6}{0}{fig2.ps}{0.5}{14}{
168: \hspace{-.5cm}
169: A more realistic model of the quantum dot of Fig.1.
170: }{f2}
171: %%%
172:
173: %%%
174: \PostScript{6}{0}{fig3.ps}{0.5}{14}{
175: \hspace{-.5cm}
176: Two one dimensional quantum wires of equal lengths, AB and CD, shown by solid lines,
177: placed along x and y directions, respectively. The origin (x=0, y=0) is
178: at the mid-point of AB. (a) CD is not connected to AB. (b) CD is connected
179: to AB.
180: }{f2}
181: %%%
182:
183: Note that in this case also reflection-symmetry holds in the x-direction
184: as well as in the y-direction, but x-y components do not separate. So by
185: tuning the boundary condition in y-direction by a gate voltage one can
186: develop nodes that try to develop across y-direction but also act across
187: x-direction and change the phase of the wave function in x-direction by
188: $\pi$. We call them NSDN because they do not originate due to the
189: symmetry of the Hamiltonian. There are many configurations of such NSDN
190: \cite{comment}, the simplest one was discussed for the stub geometry
191: (shown in Fig.3) in ref. \cite{deo96}.
192:
193: Consider for example two finite one dimensional quantum wires of equal
194: length, AB and CD placed perpendicular to each other as shown in Fig.3 by
195: the solid lines. When CD is completely detached from AB as shown in
196: Fig.3(a), then the quantum mechanical wave function in AB and CD in the
197: ground state is shown by the dotted lines. They are basically the ground
198: state wave function in an infinite potential well in one dimension (1D).
199: As is known to us, the ground state wave functions are by symmetry, even
200: parity states without any nodes, except at the boundary. But when CD is
201: attached to AB to give a T-shaped stub structure as shown in Fig.3(b),
202: then CD forms a node at C, which is also the midpoint of AB. The wave
203: function in this case is again shown by dotted lines and the wave function
204: between A and B is no longer an even parity state but an odd parity state.
205: The node at C between A and B does not originate from the symmetry of the
206: Hamiltonian and is not a symmetry dictated node (SDN). It is rather forced
207: by the boundary condition in the y-direction and is a NSDN. An
208: infinitesimal change in the length CD makes this node disappear and then
209: we have no node between A and B. The node at C induces a phase change by
210: $\pi$ and when we join A and B together to form a ring-stub system, we
211: also get persistent currents without parity effect, as parity of the
212: persistent currents is sensitive to the number of nodes in the wave
213: function \cite{leg91}.
214:
215: Fano resonances \cite{fan} are a very general feature of Q1D
216: \cite{bag90,tek93,deo98} systems in the presence of defects and the Fano
217: resonances are characterized by a zero-pole pair of the transmission
218: amplitude in the complex energy plane. When semi infinite leads are
219: attached to A and B in Fig.3(b) then the transmission amplitude of the
220: stub structure also has zero-pole pair and one gets Fano resonances
221: \cite{tek93}. At the energy corresponding to the pole a charge gets
222: trapped by the scatterer and there is also an energy when there is a zero
223: transmission across the scatterer. At the energy corresponding to the
224: zero, scattering phase shift discontinuously changes by $\pi$ due to the
225: NSDN \cite{deo96}.
226:
227: It seems at present that this phase due to NSDN is necessary to understand
228: the experimental results of Refs. \cite{yac95,sch97,CER97}. Initial
229: analysis of the experiments in terms of Friedel-sum-rule\cite{yey95}
230: revealed the shortcomings of applying Friedel-sum-rule to the quantum dot.
231: The phase change due to NSDN can explain the experimental data was first
232: proposed in Ref.\cite{jay96} and later discussed in
233: Refs.\cite{deo98,cho98,xu98,lee99,tan99,yeycm}.
234:
235: The Friedel-sum-rule (FSR) can be stated as \cite{fri52}
236: \begin{equation}
237: \th (E_{2})-\th (E_{1}) \approx \pi N(E_{2},E_{1}) .
238: \end{equation}
239: In 1D, 2D and 3D, the equality is known to be approximate,
240: and is almost exact in the WKB regime where generally transport
241: occurs.
242: Here \hspace{.1cm} $N(E_{2},E_{1})$ is the variation
243: in the number of states
244: in the energy interval $[E_{1},E_{2}]$ due to the
245: scatterer and \cite{lan61}
246: \be
247: \th = \frac{1}{2} \sum \xi_{i} = \frac{1}{2i}
248: ln (det [S] ) ,
249: \ee
250: $S$ being an n$\times$n scattering matrix and $e^{i\xi_{i}}$,
251: i=1,2,....n are
252: the n eigenvalues of the unitary matrix $S$.
253: In differential form the FSR can also be stated as
254: \begin{equation}
255: \frac{\pa \th }{\pa E} = \frac{1}{2i} \frac{\pa}{\pa E} ln (det [S] )
256: \approx \pi (\rho(E) - \rho_0(E)) ,
257: \end{equation}
258:
259: \noindent where $(\rho(E) - \rho_0(E))$ is the variation of the DOS or the
260: difference in the DOS due to the presence of the potential. $\rho(E)$ and
261: $\rho_0(E)$ can be found by integrating the local DOS $\rho(x, y, z, E)$
262: and $\rho_0(x, y, z, E))$ which are related to the electron probability.
263: However, if FSR is to be useful in mesoscopic systems where scattering
264: phase shifts can be accurately measured, and where we are always
265: interested to study a finite region of space, then we must see to what
266: extent it can give the local DOS. In fact in 1D, 2D or 3D it does so
267: extremely well and so we ask the question that how good it is in Q1D. We
268: shall show that in Q1D, FSR will fail to give the local DOS as well as the
269: global DOS at all energy regimes.
270:
271: For a symmetric scatterer in a strictly 1D system when transmission and
272: reflection amplitudes are denoted by $t$ and $r$, respectively,
273: \[ S = \left(\begin{array}{cc}
274: \displaystyle r & \displaystyle t \\
275: \displaystyle t & \displaystyle r
276: \end{array}\right) .\]
277: In this case, one can show \cite{tan99}
278: \be
279: \frac{\pa \th }{\pa E} = \frac{\pa arg(t)}{\pa E} ,
280: \ee
281: which means sum of the phases of the eigen values of S
282: is equal to the phase of some particular matrix element of S.
283:
284: But now we know that in systems that are not strictly 1D, one can have
285: zero-pole pairs and then Eq.(4) is not valid because of the $\pi$ phase
286: shifts induced by NSDN. Note that for the system in Fig.3(b), although the
287: scattering matrix is $2\times 2$, the point C is connected to 3 directions
288: and is hence a Q1D system. A quantum wire with a finite width and only one
289: propagating channel \cite{deo98} also is a Q1D system with a $2 \times 2$
290: scattering matrix. For such systems, that are not strictly one dimensional
291: but has a $2\times 2$ scattering matrix
292: \bea
293: \frac{\pa \th }{\pa E} \ne \frac{\pa arg(t)}{\pa E} \\
294: \mbox{but} \hspace{1cm} \frac{\pa \th }{\pa E} \approx \pi (\rho (E)
295: - \rho_0 (E)) ,
296: \eea
297: i.e., when we go from 1D (with 2x2 S matrix) to Q1D (also with 2x2 S
298: matrix) then Eq. (3) holds but Eq. (4) does not hold. This analysis was
299: presented by Lee \cite{lee99} and by B{\"u}ttiker and
300: Taniguchi\cite{tan99}. Their analysis is restricted to the system in
301: Fig.3(b) and $S$-matrices that are 2$\times$2. Eq.(3) is not violated in
302: the presence of NSDN and $\pi$ phase slips because $det[S]=r^2-t^2$ and if
303: $arg(t)$ changes by $\pi$ then $[arg(t^2)]$ changes by $2\pi$ or $0$ and
304: hence $det[S]$ is unaffected by the $\pi$ phase slips. Thus for such a
305: single propagating channel this phase shift is well understood by now. It
306: has been emphasized that the multi-channel case also needs to be studied
307: \cite{yeycm}, specially since scattering phase shifts can now be probed
308: experimentally in the single channel case \cite{yac95,sch97} as well as in
309: the multi-channel case \cite{CER97}, but no such study has been reported
310: so far.
311:
312: In particular, the FSR is very important in condensed matter Physics,
313: because from the scattering phase shift (that can be determined
314: experimentally) one can know the local DOS inside a disordered sample
315: without knowing its internal details. Even theoretically, except in very
316: simple situations, the wave function inside the scatterer has infinite
317: degrees of freedom and exact calculation of local DOS from the exact wave
318: function inside the scatterer may be non-trivial. While a good estimate of
319: local DOS from the S-matrix, which is on its own a very useful quantity,
320: can greatly reduce the complexities. In this work we will study the $n$
321: channel scattering problem in a Q1D quantum wire, with special emphasis on
322: FSR and Wigner delay time \cite{smi59,kumar}. Refs. \cite{lee99} and
323: \cite{tan99} parameterize the $S$-matrix in a particular way (there are in
324: fact many different ways of parameterizing the $S$-matrix ) in which the
325: scattering matrix elements become independent of energy. We will show that
326: this energy dependence, that are not important in 1D play a very crucial
327: role in Q1D multi-channel scattering. Hence in section II we will
328: generalize the work of Refs.\cite{lee99} and \cite{tan99} for real energy
329: dependent 2$\times$2 scattering matrices. The $n$ channel case will be
330: analyzed in section III and IV. In section V we will show some novel phase
331: shifts at critical energies where $S$ matrix changes dimensions. Section
332: VI is devoted to conclusions.
333:
334:
335: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
336: \section{Scattering in one-dimension and negative values of $d \th /dE$}
337: \label{s2}
338: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
339:
340: In Fig.4 we consider a potential that is described in details in the
341: figure caption. The quantum mechanical wave function or the solution to
342: the Schr{\"o}dinger equation in different regions is also shown and
343: explained in the figure and its caption. We will always normalize the
344: incoming wave-function such that its amplitude is 1. Griffiths boundary
345: conditions for this system gives the following equations \cite{gri,deo93}
346: (we use $2m=1$ and $\hbar=1$).
347: \begin{equation}
348: 1 + r = a + b ,
349: \end{equation}
350: \begin{equation}
351: a e^{ikl} + b e^{-ikl} = t ,
352: \end{equation}
353: \begin{equation}
354: ik(1-r) - ik(a-b) = - V (1 + r)\hspace{.5cm}\mbox{and}
355: \end{equation}
356: \begin{equation}
357: ik(a e^{ikl} - b e^{-ikl}) -ikt = - V(a e^{ikl} + b
358: e^{-ikl}) .
359: \end{equation}
360: %%%
361: \PostScript{6}{0.5}{fig4.eps}{0.5}{14}{
362: \hspace{-.5cm}
363: Two identical delta function potentials separated by a length $l$.
364: Strength of each potential is $V$. The thick vertical lines denote the
365: positions of the potentials and the thin horizontal line is the direction
366: of propagation. A plane wave of unit amplitude is incident from the left
367: and wave function in different regions (marked as I, II and III) is
368: written down in the figure. $r$ and $t$ are the reflection and
369: transmission amplitudes, respectively, of the entire system and $ k=\sqrt
370: {E} $ is the incident wave vector.
371: The origin of coordinates is shown in the figure.
372: }{f1}
373: %%%
374: \noindent We will first analyze this system in detail and generalize the
375: results of Refs. \cite{lee99,tan99} further by considering realistic
376: energy dependent $r$ and $t$, that will later help us to accentuate the
377: new features that can be observed in a multi-channel disordered quantum
378: wire.
379:
380: First of all let us calculate the local DOS and
381: global DOS to see how much it agrees
382: with $d \theta /dE$. Although the basic facts
383: discussed in this section is known in the Greens
384: function formalism, to the best
385: of our knowledge, quantitative disagreement (or agreement)
386: has not been shown so far.
387: Using
388: quantum mechanical expression for the local DOS integrated over the region
389: II in Fig.~4, i.e.,
390: $$ \rho_R = {2 \over hv} \int ^l_0 \mid a e^{ikx} + b
391: e^{-ikx} \mid ^2 dx, $$
392: it is easy to show from Eq. 3 that (the Eq. below
393: is consistent with Ref. \cite{tan99} and some hints on its derivation
394: is given in Ref \cite{com})
395: \be
396: {d \theta \over d (kl)} \approx {
397: \bar \rho } + {\rho _q \over l} = \rho '\hspace{.5cm} \mbox{(say)},
398: \ee
399: \be
400: \mbox{where}\hspace{.5cm} { \bar \rho } = \mid a \mid ^{2} + \mid b
401: \mid ^{2} \hspace{.1cm}\mbox{and}
402: \ee
403: \be \rho_{q} = \int ^{l}_{0}
404: \left(ab^* e^{2ikx} +ba^* e^{-2ikx} \right) dx .
405: \ee
406: Here $\frac{\rho_{q}}{l}$ is a term that arises because of quantum
407: mechanical interference and it can be seen that the integrand in Eq.(13)
408: oscillates with $x$. For $|\frac{E}{V}|>1$ (this is the WKB regime
409: when the electron does not feel the potential strongly and is
410: almost entirely transmitted) $\frac{\rho_{q}} {l}$ is
411: negligibly small. This is shown in Fig.5, where we plot $\rho^{'}$ (the
412: dashed curve) and ${\bar \rho }$ (the dotted curve). The two curves are
413: almost the same for $|{E \over V}|>1$, which means ${\rho_q \over l}$,
414: being the difference between the dashed and dotted curves is vanishingly
415: small above this energy. It is known that to get the equality between the
416: LHS and RHS of Eq. 11, it is necessary to drop the term ${\rho_q \over
417: l}$ \cite{gas}.
418: It is also known that this deviation arises because we are
419: considering the local DOS rather than the global DOS.
420: If we consider the global DOS, i.e.,
421: $$\rho(E)={2 \over hv} \int^{\infty}_{-\infty} \psi^*(x) \psi(x) dx$$
422: where $\psi(x)$ is the quantum mechanical wavefunction at x,
423: then instead of Eq. (11) we get
424: $${d \theta \over d(kl)} \approx
425: Lt_{l \rightarrow \infty} (|a|^2 + |b|^2).$$
426: The equality is still approximate although almost exact in the WKB
427: regime because the RHS is positive definite, while it is well known
428: that the LHS can become negative at low energy (non-WKB regime)
429: as will be demonstrated below.
430: To get the equality it is
431: necessary to neglect something else as we shall soon see. This second thing
432: that we shall drop is however not due to the fact that we are considering
433: the local DOS. It is an inherent approximation of the FSR even when we are
434: considering the global DOS and hence also when we are considering the
435: local DOS. The $\rho_q/l$ term or the interference term inside the
436: scatterer does not arise in the case when $l \rightarrow 0$ as in the case
437: to be considered in section IV. All the deviation to be observed there is
438: due to this inherent weakness of the FSR. We shall also see below that
439: this inherent weakness of the FSR is negligible in 1D, 2D or 3D but
440: becomes very formidable in quasi 1D.
441:
442: %%%
443: \PostScript{6}{0}{fig5.ps}{0.5}{14}{
444: \hspace{-.75cm}
445: The system under consideration is shown in Fig.4. The solid curve gives
446: the exact $d \theta/d(kl)$, the dashed curve gives the $\rho^{'}$ and the
447: dotted curve gives ${\bar \rho}$. This plot is done for $Vl^2=-5,\hbar=1,
448: 2m=1$.
449: }{f2}
450: %%%
451:
452: %%%
453: \PostScript{6}{0}{fig6.ps}{0.5}{14}{
454: \hspace{-.5cm}
455: The system under consideration is shown in Fig.4. The plot is of $d
456: \theta/d(kl)$ versus $kl$ for the system for different values of $Vl^2$.
457: The dotted curve is for $Vl^2=-2 $, the dashed curve is for $Vl^2=-2.1$,
458: the solid curve is for $Vl^2=-5$, the long dashed curve for $Vl^2=-8$. We
459: use $\hbar=1, 2m=1$.
460: }{f2}
461: %%%
462: One can prove that \[{ \bar \rho } = \mid a \mid^{2} + \mid b \mid^{2} =
463: \frac{1-\mid r^{'}\mid^{4}}{\mid 1-{r^{'}}^{2}\tau^{2}\mid^{2}} ,\] for
464: any energy dependent reflection amplitude $r^{'}$ of one of the two
465: identical scatterers in Fig.4, where, $\tau=e^{ikl}$.
466: Hence as indicated by Eq. 11, it would be
467: interesting if we can obtain a good estimate of ${\bar \rho}$ or $\rho'$
468: from $\th$. In the appendix I it is explained that if $\frac{d
469: r'}{dE}=\frac{d t'}{dE} \rightarrow 0$ (which means the
470: scatterers are non-dispersive and that
471: only happens at high
472: energy in 1D, 2D and 3D) then, $\frac{d \theta}{d (kl)}$ reduces to the
473: expression $\frac{1-\mid r^{'}\mid^{4}}{\mid
474: 1-{r^{'}}^{2}\tau^{2}\mid^{2}}$, and then therefore, ${d \th \over d(kl)}
475: = {\bar \rho}$. It is shown in section III of Ref. \cite{yeycm} (see Eq. 6
476: and 7 therein), that to relate ${d \theta \over dE}$ to the global DOS, one
477: has to neglect the energy dependence of the self energy, that depends on
478: the coupling of the system to the leads, i.e., $r'$ and $t'$. Thus our
479: results are consistent with that. The exact
480: $\frac{d\th}{d(kl)}$ is shown in Fig.5 by the solid curve. Note that in
481: the relevant energy regime ($|\frac{E}{V}|>1$),
482: the solid curve is very close to the dashed and dotted curves,
483: which means FSR works very well for the local DOS
484: as well as for the global DOS. But for
485: $|\frac{E}{V}|<1$, $\frac{d\th}{d(kl)}$ deviates from ${\bar \rho}$. But
486: since transport effects in weak localization or diffusive or ballistic
487: regime occur at Fermi energies, that is normally higher in semiconductors
488: as well as metals in comparison to the energy where the two curves deviate
489: substantially from each other, Friedel sum rule is often useful in
490: condensed matter to obtain a good estimate of local DOS as well as global
491: DOS.
492:
493: $d \th / dE$ is also well known as Wigner delay time \cite{smi59,kumar}.
494: In the stationary phase approximation, it gives the time spent by the
495: scattered particle at the impurity site. In the low energy regime, where
496: dispersion becomes significant, the stationary phase approximation is
497: not valid and $d \th /dE$ can become negative and does not give a
498: meaningful particle
499: delay time. In this regime $d \th / dE$ becomes negative as
500: the phase velocity becomes larger than the group velocity and even larger
501: than the velocity of light, and although such super-luminous particles can
502: be detected experimentally they cannot carry any signal or information. In
503: Fig.6 we show the negative behavior of $d \theta/d(kl)$. We find that as
504: the strength of the impurities is varied, $d \theta /d(kl)$ can become
505: more or less negative (see Fig.6), maximizing at $Vl^2=-2.1$ for the
506: symmetric delta potentials. The energy
507: regime, where $d \theta /d(kl)$ can be negative remains the same for all
508: $V$ and always $|\frac{E}{V}|<1$. We have checked for all these values of
509: $V$ that apart from this insignificant energy range, FSR works very well.
510: FSR has a close counterpart in quantum mechanics called Levinson's
511: theorem. It is known that Levinson's theorem also breaks down in the
512: presence of zero energy bound states \cite{new82} that can be degenerate
513: with scattering states.
514:
515:
516: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
517: \section{ Wigner delay time in quasi-one-dimensions}
518: \label{s3}
519: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
520:
521: %%%
522: \PostScript{6}{0}{fig7.ps}{0.5}{14}{
523: \hspace{-.5cm}
524: Here we show a quantum wire of width $W$. The dash-dotted curve is a line
525: through the middle of the quantum wire, and it is also taken to be the
526: x-axis. The origin of the coordinates is shown in the figure. A delta
527: function potential $V(x,y)=\gamma \delta(x) \delta(y-y_{i})$ is situated
528: at $x=0$ and $y=y_{i}$ and marked as $\times$. We consider scattering
529: effects when the incident electron is from the left. The sub-bands on the
530: left of the impurity is denoted as 1 for the first mode (i.e., its wave
531: function can be obtained by putting $n$=1 in Eq.(14) with appropriate sign
532: for $k_n$) and 2 for the second mode (i.e., its wave function can be
533: obtained by putting $n$=2 in Eq.(14) with appropriate sign for $k_n$).
534: Similarly the sub-bands on the right of the impurity is denoted as 3 for
535: the first mode (i.e., its wave function can be obtained by putting $n$=1
536: in Eq.(14) with appropriate sign for $k_n$) and 4 for the second mode
537: (i.e., its wave function can be obtained by putting $n$=2 in Eq.(14) with
538: appropriate sign for $k_n$). The impurity at $\times$ mixes these wave
539: functions to give a scattering matrix element $t'_{mn}$ from mode $m$ to
540: mode $n$.
541: }{f2}
542: %%%
543:
544: In Fig.7 we consider a quasi-one-dimensional quantum wire with an
545: attractive impurity at (0, $y_i$), having electrons confined along the
546: y-direction but free to move along the x-direction. While the states far
547: away from the impurity are good momentum states, the impurity can mix the
548: different modes and in this region of mode mixing, the wave function is
549: $\psi (x,y)= \Sigma_n c_n(x) \chi_n(y)$, where $\chi_n(y)$ are the
550: transverse wave functions in the absence of the impurity and $c_n(x)$ are
551: position dependent coefficients that has to be determined by mode
552: matching. The confining potential in the y-direction or the transverse
553: direction is taken to be hard wall. Thus the transverse wave-function is
554: of the form $\chi_n(y)=Sin \frac{n\pi}{W}(y+\frac{W}{2})$. For a given
555: width $W$ of the quantum wire one can choose the energy range of the
556: incident electron such that only two modes are propagating, although, all
557: the other modes (infinite in number, showing that the internal wave
558: function can have infinite degrees of freedom, which makes it difficult to
559: calculate the exact local
560: DOS from the internal wave function) will be present
561: but as evanescent modes. For example, if the energy of the electron be
562: $E$ then for propagation in the n-th transverse mode (in short we will
563: refer this as n-th mode) the wave-function is of the form
564: \be
565: Sin \frac{n\pi}{W}(y+\frac{W}{2})\hspace{.2cm}e^{ik_{n}x}
566: \ee
567: where $k_{n}= \sqrt{E-E_n}$, $E_n$ being
568: $\frac{n^2\pi^2}{W^2}$ \hspace{.2cm} and \hspace{.2cm} $n=1,2,3,
569: \cdots$ $\infty $. Here we have used $\hbar=2m=1$.
570: To have the n-th mode to be propagating it is necessary that
571: $k_{n}^2 > 0 $ or
572: \be
573: n < \frac{W}{\pi} \sqrt{E}.
574: \ee Thus we can choose the energy range where there
575: will be two propagating modes, i.e., $n=1$ and $n=2$ satisfy condition (15).
576: The rest of the modes ($n>\frac{W}{\pi}\sqrt{E}$) will be evanescent, whose
577: wave functions are of the form
578: \be
579: Sin \frac{n\pi}{W}(y+\frac{W}{2})\hspace{.2cm}e^{- \kappa_{n}x} ,
580: \ee
581: where $\kappa_n=\sqrt{E_n-E}$. These evanescent modes just renormalize the
582: scattering matrix elements and drop out of the problem.
583: The transmission amplitude from m-th incident mode to n-th
584: scattered mode is given by \cite{bag90}
585: \bea
586: t_{mn}^{'} =
587: -\frac{i\Gamma_{mn}}{2d\sqrt{k_{m}k_{n}}},\\
588: \mbox{where} \hspace{.5cm} d = 1+\sum ^{e}\frac{\Gamma_{nn}}{2\kappa_{n}}+
589: i\sum ^{p}\frac{\Gamma_{nn}}{2k_{n}}.
590: \eea
591: Here $\sum ^{e}$ denotes the sum over all evanescent modes and $\sum ^{p}$
592: denotes the sum over all propagating modes. Eq.(17) holds only for
593: inter-subband transmission amplitudes and all reflection amplitudes. When
594: we say reflection amplitude we mean the following. For an electron
595: incident from the left, all outgoing channels to the left are reflection
596: channels. According to this convention $t_{11}^{'}$, $t_{22}^{'}$,
597: $t_{21}^{'}$ and $t_{12}^{'}$ are reflection amplitudes. Inter-subband
598: transmission amplitude are then obviously $t_{14}^{'}$ and $t_{23}^{'}$.
599: The intra-subband transmission amplitudes $t_{13}^{'}$ and $t_{24}^{'}$
600: (according to numbering of channels explained in Fig.7) are given by
601: \be
602: t_{13}^{'} = 1+t_{11}^{'}
603: \hspace{.5cm}\mbox{and}\hspace{.5cm}
604: t_{24}^{'} = 1+t_{22}^{'}.
605: \ee
606: Here $ \Gamma_{nm} $ is the strength of coupling between the n-th mode and
607: the m-th mode. If we take the impurity to be a delta function potential
608: i.e., $V(x,y)= \gamma \delta(x) \delta(y-y_i)$, and the confining
609: potential in the y-direction to be hard wall ($V=\infty$ for
610: $-\frac{W}{2}\ge y\ge \frac{W}{2}$, and $0$ everywhere else except the
611: impurity site $\times$ ) (see Fig.7) then
612: \[ \Gamma_{nm} = \gamma Sin \frac{n \pi}{W}(y_{i}+\frac{W}
613: {2})
614: Sin \frac{m \pi}{W}(y_{i}+\frac{W}{2}). \]
615:
616: Apart from the two propagating modes we consider two evanescent modes and
617: truncate the infinite series of evanescent modes (note that although the
618: series is strongly converging, the reason for truncating is different,
619: stronger and explained in more detail after Fig. 12) in Eq. (18) and so
620: Eq. (18) becomes \\
621: \be
622: 1+ \frac{\Gamma_{33}}{2\kappa_{3}}+\frac{\Gamma_{44}}{2\kappa_{4}}
623: +i\left( \frac{\Gamma_{11}}{2k_{1}}+\frac{\Gamma_{22}}{2k_{2}}\right)=d_2
624: \hspace{.2cm} \mbox{(say)}.
625: \ee
626: The lowest evanescent mode (putting $n$=3 in Eq. (16) gives its wave
627: function) has even parity in the transverse direction. For a negative
628: impurity potential $i.e.$, $\gamma < 0$, it also has a bound state at
629: $E=E_{3b}$, where $E_{3b}$ is given by the solution of
630: \be
631: 1+\frac{\Gamma_{33}}{2\kappa_{3}}+\frac{\Gamma_{44}}{2\kappa_{4}}=0.
632: \ee
633: Since $E_{3b}<{9 \pi^2 \over W^2}$, $E_{3b}$ can be degenerate with
634: scattering states (the $n$=1 and $n$=2 modes are the scattering states).
635: The higher evanescent mode (putting $n$=4 in Eq. (16) gives its wave
636: function) has odd parity in the transverse direction and this too has a
637: bound state at $E=E_{4b}$, where $E_{4b}$ is given by the solution of \be
638: 1+ \frac{\Gamma_{44}}{2\kappa_{4}}=0. \ee Once again depending on
639: $\gamma$, $E_{4b}$ can be degenerate with the scattering states. The
640: effect of including more evanescent modes is just to renormalize the
641: strength of the impurity potential and does not give anything new
642: \cite{bag90}.
643:
644: The scattering matrix in this case is
645: \be
646: S = \left[ \begin{array}{llcl}
647: t_{11}^{'} & t_{12}^{'} & t_{13}^{'} & t_{14}^{'} \\
648: t_{21}^{'} & t_{22}^{'} & t_{23}^{'} & t_{24}^{'} \\
649: t_{31}^{'} & t_{32}^{'} & t_{33}^{'} & t_{34}^{'} \\
650: t_{41}^{'} & t_{42}^{'} & t_{43}^{'} & t_{44}^{'}
651: \end{array} \right]
652: =
653: \left[ \begin{array}{cc}
654: r_{2c} & t_{2c} \\
655: \hat{t}_{2c} & \hat{r}_{2c}
656: \end{array} \right] ,
657: \ee
658: \[ \mbox{where}\hspace{.5cm} r_{2c} =
659: \left[ \begin{array}{cc}
660: t_{11}^{'} & t_{12}^{'} \\
661: t_{21}^{'} & t_{22}^{'}
662: \end{array} \right]
663: \]
664: \[\mbox{and} \hspace{.5cm} t_{2c} =
665: \left[ \begin{array}{cc}
666: t_{13}^{'} & t_{14}^{'} \\
667: t_{23}^{'} & t_{24}^{'}
668: \end{array} \right] .
669: \]
670: $\hat{t}_{2c} = t_{2c}$ due to time reversal symmetry and $\hat{r}_{2c} =
671: r_{2c}$ for a symmetric scatterer as that considered here. Now once again
672: due to micro-reversebility $t_{12}^{'} = t_{21}^{'}$. Also $t_{12}^{'} =
673: t_{14}^{'}$ because in both $t_{12}^{'}$ and $t_{14}^{'}$ the density of
674: states in the input as well as the output channel is the same, and also
675: the incident channel momenta and the outgoing channel momenta are the same
676: in the transverse as well as in the propagating direction. Also
677: $t_{23}^{'} = t_{41}^{'}$ because transmission amplitude should be
678: independent of the position of the observer i.e., whether the observer is
679: looking into the plane of the paper or out of the plane of the paper. Thus
680: among the 16 matrix elements in Eq. (23) we are left with only 5 that are
681: distinct. They are $t_{11}^{'},t_{12}^{'},t_{22}^{'},t_{13}^{'}$ and
682: $t_{24}^{'}$.\\ From Eq. (17) and (19),
683: \bea
684: t_{11}^{'} = -\frac{i\Gamma_{11}}{2d_2k_{1}},\\
685: t_{12}^{'} = -\frac{i\Gamma_{12}}{2d_2\sqrt{k_{1}k_{2}}},\\
686: t_{22}^{'} = -\frac{i\Gamma_{22}}{2d_2k_{2}}, \\
687: t_{13}^{'} = \frac{1+ \frac{\Gamma_{33}}{2\kappa_{3}}+\frac{\Gamma_{44}}
688: {2\kappa_{4}}+i\frac{\Gamma_{22}}{2k_{2}}}
689: {d_2} \hspace{.5cm}\mbox{and}\\
690: t_{24}^{'} = \frac{1+ \frac{\Gamma_{33}}{2\kappa_{3}}+\frac{\Gamma_{44}}
691: {2\kappa_{4}}+i\frac{\Gamma_{11}}{2k_{1}}}
692: {d_2}\hspace{.2cm}.
693: \eea
694: Knowing these matrix elements, the scattering matrix is completely known
695: and $\th $ can also be calculated.
696:
697: We find some further relationships between the scattering phase shifts as
698: follows. First of all
699: \be
700: arg(t_{11}^{'}) = arg(t_{22}^{'}) =tan^{-1}\frac{Re(d)}{Im(d)} .
701: \ee
702: Secondly, when $\frac{4\pi^2}{W^2} <E_{3b}<\frac{9\pi^2}{W^2}$, i.e., the
703: bound state of the 3rd subband lies in the energy range where one can have
704: two propagating subbands, then the bound state $E_{3b}$ drastically
705: changes the scattering matrix elements in that energy range. So in this
706: energy range $\frac{4\pi^2}{W^2}$ to $\frac{9\pi^2}{W^2}$ we find
707: \be
708: arg(t_{12}^{'}) \mp \frac{\pi}{2} = \th + \pi .
709: \ee
710: Here negative sign is to be taken when $E_{3b}$ lies in this energy range.
711: Otherwise the positive sign has to be taken. $\th$ is to be calculated
712: from Eq. (2) using Eq. (23). Thirdly we find
713: \be
714: arg(t_{11}^{'}) \pm \pi = arg(t_{12}^{'}) .
715: \ee
716: Note that
717: in contrast to Eq. (30) here the
718: choice of $\pm$
719: sign is arbitrary. However consistent with this choice is the
720: following
721: \be
722: arg(t_{11}^{'}) \pm \frac{\pi}{2} = \th + \pi ,
723: \ee
724: where once again + sign is to be
725: taken when $E_{3b}$ is present in this energy
726: range and - sign is to be taken when absent.
727:
728: We thus find very simple analytical expressions for $\th$ in the sense
729: that one need not calculate it from a $4\times4$ scattering matrix but can
730: calculate it from the argument of a single matrix element like
731: $t_{11}^{'}$ or $t_{12}^{'}$ or $t_{22}^{'}$. These relations are
732: analogous to Eq. (4) in section I obtained for purely one dimensional
733: case, i.e., one need not calculate $\th$ from 2$\times$2 matrix but one
734: can find it from the argument of a single matrix element.
735:
736: %%%
737: \PostScript{6}{0}{fig8.ps}{0.5}{14}{
738: \hspace{-.5cm} The system under consideration is shown
739: in Fig. 7. The plot is of the argument of various transmission amplitudes
740: ($arg(t_{mn}^{'})$ in radians)
741: from incident channel $m$ to propagating channel $n$ versus
742: E$W^2$ . The solid curve gives $arg(t_{11}^{'})$,
743: the long dashed curve gives $arg(t_{13}^{'})$ and
744: the dotted curve gives $arg(t_{24}^{'})$.
745: We use $\gamma=-10$, $y_{i}=.21 W$
746: and $x_i = 0$
747: }{f2}
748: %%%
749:
750:
751: In Fig.~8, we plot only the distinct arguments of the scattering
752: amplitudes versus energy of the incident electron. We find that all of
753: them show negative slopes over a very large range of energy and as already
754: discussed, such negative slopes give rise to fundamental questions in
755: quantum mechanics \cite{smi59,kumar}. Now in Q1D we find that
756: this negative slope is not restricted to low energy but can occur at any
757: arbitrary energy. Notice for example, $arg(t'_{13})$ and $arg(t'_{24})$
758: show larger negative slopes at the highest possible energies for two
759: channel propagation. The rest of this section will be devoted to
760: understanding these negative slopes that at first sight looks very
761: different in nature and character in the three curves in Fig.~8,
762: and also to understanding what will happen when there are more
763: than two propagating modes. We will
764: address the FSR in the next section.
765:
766: It is to be noted that among all these scattering matrix elements
767: $t_{11}^{'}$ and $t_{13}^{'}$ exist in the single channel regime
768: (i.e., $ \pi^2< EW^2 < 4\pi^2$) where
769: $t_{11}^{'} $ is the reflection amplitude and $t_{13}^{'} $ is the
770: transmission amplitude. The phase of $t_{13}'$ in the single channel
771: regime is known to change discontinuously by
772: $\pi$ when $t_{13}'$ is 0, i.e., $t_{13}'$ has a zero
773: in real energy. In the two channel
774: regime if we write from simplifying Eq. (27)
775: \be
776: t_{13}^{'} =\frac{k_2(2\kappa_3+g_3)+i\kappa_3
777: g_2}{k_2(2\kappa_3+g_3)+i\kappa_3(g_2+\alpha
778: g_1)},
779: \ee
780: where $\alpha=\frac{k_2}{k_1}$ and
781: $ g_s = \frac{2\kappa_4}{\Gamma_{44}+2\kappa_4}\hspace{.2cm}
782: \Gamma_{ss}$; with $s$=1,2,3.
783: then interestingly, we see that it has a zero
784: in complex energy and not in real energy.
785:
786: If we modify the Breit-Wigner line shape formula of 1D to
787: include complex zeroes and write
788: \be
789: t_{mbw}(E) = A\hspace{.1cm}\frac{E-E_0+i\Gamma _0}{E-E_p+i\Gamma _p}
790: \hspace{.1cm},
791: \ee
792: where A is a
793: normalization factor, then just as $\Gamma_p$ gives the scale over
794: which $arg[t_{mbw}(E)]$ increase at $E=E_p$, $\Gamma _0$ gives a scale
795: over which $arg[t_{mbw}(E)]$
796: decrease at $E=E_0$ where $|t_{mbw}(E)|^2$ also shows a minimum
797: at $E=E_0$ (but not zero).
798: One can check this very easily (let us say, when
799: $E_0$=2, $E_p$=1 and $\Gamma_0=\Gamma_p=0.5$) and so we do not
800: demonstrate it here. Now from Eq. (33) we
801: see that at an energy which satisfies the condition
802: \be
803: 2\kappa_3+g_3=0 ,
804: \ee
805: the real part of the numerator in Eq. (33) is zero. Condition (35)
806: is the same as the condition (21) for a
807: bound state $E_{3b}$ coming from the 3rd subband
808: that is degenerate with scattering states.
809: So, around this energy where Eq.(35) is satisfied (lets say at $E=
810: E_{3b} \equiv E_0$)
811: $arg(t_{13}')$ will undergo a drop over an energy scale
812: determined by the imaginary part, $\kappa_3 g_2 $, i.e.,
813: $\Gamma _0\equiv\kappa_3 g_2$.
814:
815: It can be seen in Fig.9 that $|t_{13}^{'}|^2$ (dotted curve) shows a
816: narrow minimum around an energy E$W^2\simeq 84$ (which is the solution of
817: Eq. (35) or Eq. (21)) and at this energy $arg(t_{13}^{'})$ shows a very
818: sharp drop over a narrow energy range determined by $\kappa_3 g_2$. Hence
819: by decreasing/increasing this quantity $\kappa_3 g_2$ we can make the
820: phase drop sharper/broader. $g_2$ can be made smaller in two ways, first
821: by decreasing $\gamma$ and second by taking the impurity closer to a node
822: in the transverse wave function. The plot for a decreased value of
823: $\gamma$ is shown in Fig.10 and it confirms this.
824:
825: %%%
826: \PostScript{6}{0}{fig9.ps}{0.5}{14}{
827: \hspace{-.5cm}
828: The system under consideration is shown in Fig.7.
829: The solid curve gives $arg(t_{13}^{'})$ in radians shifted by
830: $\frac{\pi}{4}$ radians in the
831: y-direction and the dotted curve gives $|t_{13}^{'}|^2$.
832: Both the functions are plotted versus
833: $EW^2$ using $x_i=0$, $y_i=.45W$ and $\gamma=-15$.
834: }{f2}
835: %%%
836:
837: Note that the quantity $\kappa_3 g_2$ is actually energy dependent. But
838: in Fig.9 and Fig.10 $\kappa_3 g_2$ is so small that the drop occurs over a
839: scale in which $\kappa_3 g_2$ is roughly constant. For larger values of
840: $\kappa_3 g_2$, the phase drop will be determined by a complex competition
841: between $\kappa_3$ and $g_2$. This is shown in Fig.11. First of all the
842: scale of the phase drop becomes so large that any sensitivity to the
843: position of the bound state can not be seen. Secondly, $\kappa_3 g_2$ can
844: not be taken to be a constant over this large scale and the enhancement of
845: the negative slope for $EW^2 > 79$ is a signature of the fact that here
846: $\kappa_3 \rightarrow 0$ and so $\kappa_3 g_2 \rightarrow 0$ as $EW^2$
847: increases.
848:
849: %%%
850: \PostScript{6}{0}{fig10.ps}{0.5}{14}{
851: \hspace{-.5cm}
852: The system under consideration is shown in Fig.7. The solid curve gives
853: $arg(t_{13}^{'})$ in radians shifted by $\frac{\pi}{4}$ radians in the
854: y-direction and the dotted curve gives $|t_{13}^{'}|^2$. Both the
855: functions are plotted
856: versus $EW^2$ using $x_i=0$, $y_i=.45W$ and $\gamma=-10$
857: }{f2}
858: %%%
859: \noindent Similarly if we rewrite Eq.(28) as
860: \[ t_{24}^{'} = \frac{k_1(2\kappa_3+ g_3)+i\kappa_3 g_1}
861: {k_1(2\kappa_3+ g_3)+i\kappa_3(g_1+\beta g_2)}, \]
862: where $\beta=\frac{k_1}{k_2}$; then it is clear that the behavior of
863: $arg(t_{24}^{'})$ will be qualitatively the same. It is indeed found in
864: Fig.8 that the behavior of $arg(t_{24}^{'})$ is similar to that of
865: $arg(t_{13}^{'})$.
866:
867: $\th = \frac{1}{2i}ln[det[S]]$ is shown in Fig.12 as a function of energy,
868: for different values of $\gamma$. The minimum in $\th$ follows the
869: $E_{3b}$ and so the energy range where the slope of $\th$ versus E is
870: negative is determined by the $E_{3b}$. Note that when $E_{3b}$ goes out
871: of this energy range the $\th$ versus E has a positive slope everywhere.
872: So in Fig.12, the negative slope arises whenever a bound state $E_{3b}$ is
873: degenerate with the scattering states ($n$=1 and $n$=2), and
874: non-monotonously scatter and disperse the scattering states.
875: For weaker impurities in Q1D, the negative slope
876: occur at higher energies and also are steeper as demonstrated in Fig.12.
877: This is in contrast to what happens in 1D and demonstarted in Fig. 6,
878: that the energy where the negative slopes occur is always for
879: $E/V<1$.
880: %%%
881: \PostScript{6}{0}{fig11.ps}{0.5}{14}{
882: \hspace{-.5cm}
883: The system under consideration is shown in Fig.7.
884: The solid curve gives $arg(t_{13}^{'})$ in radians
885: shifted by $2\pi$ radians in the
886: negative y-direction versus $EW^2$ for
887: $\gamma=-47.1371$. The dashed curve is for $\gamma=-25.197$. We
888: use $x_i=0$ and $y_i=.21W$.
889: }{f2}
890: %%%
891:
892: %%
893: \PostScript{6}{0}{fig12.ps}{0.5}{14}{
894: The system under consideration is shown in Fig.7. The plot is of $\th$
895: in radians versus $EW^2 $ for different $\gamma $. The dot-dashed curve is
896: shifted by $\pi$ radians in y-direction for $\gamma = -47.1371$,
897: corresponding $E_{3b}$ is at $EW^2=35$ which is less than the propagating
898: threshold $EW^2 \si 39$ of the second transverse mode. The dashed curve
899: is for $\gamma = -25.197 $, corresponding $E_{3b}$ is at $EW^2=80$. The
900: dotted curve is for $\gamma = -15$, corresponding $E_{3b}$ is at
901: $EW^2=86.606$. The solid curve is for $\gamma = -10 $, corresponding
902: $E_{3b}$ is at $EW^2=87.982$. We use $ y_{i}=.21W$ and $x_i=0$. The arrows
903: accentuate the positions of the minima that is shifting towards higher
904: energies for weaker impurities.
905: }{f2}
906: %%%
907:
908: We have used two evanescent modes in our calculations because one can
909: include as many evanescent modes without changing the nature of the
910: negative slopes as long as the positions of the bound states $E_{3b}$ and
911: $E_{4b}$ remain the same. One can check this that with four evanescent
912: modes and $\gamma =-6.46584$, the negative slopes are the same as in Fig.
913: 8, which means that the third and the fourth evanescent modes just
914: renormalizes $\gamma$ from -6.46584 to -10. The exact renormalization
915: takes place according to a formula $\gamma^h=\gamma^n/d$, where $\gamma^h$
916: is the $\gamma$ value used here and $\gamma^n$ is the renormalized value
917: of $\gamma$ when we use $m$ evanescent modes instead of two. $d=
918: (1+\Gamma_{55}^n/(2 \kappa_5) + \Gamma_{66}^n/(2 \kappa_6)....
919: \Gamma_{mm}^n/(2 \kappa_m)$, where $\Gamma_{mm}^n= =
920: \gamma^n Sin \frac{m \pi}{W}(y_{i}+\frac{W} {2})^2$.
921: Solving this one can find the renormalized value of $\gamma$ i.e.,
922: $\gamma^n$ that keep the minimum of any of the curves for the scattering
923: phase shifts considered here unchanged. It is worthwhile mentioning that
924: at the band edges (i.e., $E \approx 39$ and 89, in the figures considered
925: in this section), the value of any curve is independent of the number of
926: evanescent modes, as all the modes get decoupled there. In other words,
927: number of evanescent modes considered does not change the nature of the
928: negative slopes. Only the positions of the bound states are important.
929:
930: In order to generalize to arbitrary number of propagating channels we
931: change our notations slightly. For electrons incident from the
932: left/right, we call the scattered channels towards the left/right as
933: reflection channels (the rest being transmission channels) and change our
934: notation to
935: \bea
936: {\tilde r_{11}} = t_{11}^{'}\hspace{.2cm}, \\
937: {\tilde r_{22}} = t_{22}^{'}\hspace{.2cm}, \\
938: {\tilde r_{12}} = t_{12}^{'} .
939: \eea
940: Thus all possible reflection channels are distinct. All the intra-subband
941: transmission channels are also distinct and they are denoted as ${\tilde
942: t_{11}}$ and ${\tilde t_{22}}$ where
943: \bea
944: {\tilde t_{11}} = t_{13}^{'} \\
945: \mbox{and} \hspace{.5cm} {\tilde t_{22}} = t_{24}^{'} .
946: \eea
947: All other scattering matrix elements are equal to one of these 5
948: elements. In this notation the only difference is that the lowest channel
949: ($n$=1) on the transmission side is marked 1 instead of 3. So when we say
950: ${\tilde t_{11}}$ we mean transmission amplitude from the $n$=1 channel on
951: the left to the $n$=1 channel on the right. We find from Eqs. (29), (30),
952: (31) and (32)
953: \be
954: \frac{d}{dE}\hspace{.1cm} arg({\tilde r_{mn}}) = \frac{d}{dE}
955: \hspace{.1cm}(\frac{1}{2i}
956: ln[det[S]])
957: \ee
958: We find the above relation to be true for any number of propagating
959: modes. So $m$ and $n$ can take any integer value less than or equal to
960: $p$, where $p$ is the total number of propagating modes. For two
961: propagating modes $p$=2, for three propagating modes $p$=3 and so on. So
962: Eq.(41) is analogous to the 1D case given in Eq.(4). That is when the
963: dimension of the matrix $S$ becomes very large, then it is sufficient to
964: consider the argument of a single matrix element in order to calculate the
965: complicated quantity on the RHS of Eq. 41. In the energy regime where
966: there are two propagating channels, the negative slopes in $\th$ versus
967: incident energy curves are determined by $E_{3b}$, and when there are 3
968: propagating channels then the negative slopes are determined by $E_{4b}$
969: and so on.
970:
971: The scattering phase shifts of transmission channels i.e. $arg({\tilde
972: t_{mn}})$, where again $m$ and $n$ can take all possible integer values
973: less than or equal to $p$, show sharp or gentle phase drops when the
974: scattering states are degenerate with a bound state, depending on the
975: value of the imaginary part in the numerator of ${\tilde t}_{mn}$. In the
976: single channel regime the imaginary part in the numerator is zero and
977: phase drops take the limiting value when the phase drops are absolutely
978: discontinuous by $\pi$. Just as the discontinuous phase drop in single
979: channel case do not affect $\th$ in any way, the phase drops of the
980: $arg({\tilde t}_{mn})$ also do not affect $\th$ in any way and $\th$
981: behaves similarly as $arg({\tilde r_{mn}})$.
982:
983: \section{density of states and Friedel sum rule in quasi-one-dimension}
984:
985: The local DOS is given by the following expression \cite{bag90}
986: \be
987: \rho_R = \int_{R} dx \int^{\frac{W}{2}}_{\frac{-W}{2}} dy
988: \sum_{m,k_m} \delta(E-E_{m,k_m}) \mid \psi_{m,k_m}(x,y)\mid ^2
989: \ee
990: Here $E$ is the incident energy and R is the integration region where
991: modes are mixed. $m$ and $k_m$ are the two quantum numbers that define an
992: incident electron wavefunction, $A_me^{ik_mx}Sin
993: \frac{m\pi}{W}(y+\frac{W}{2})$ whose energy is $E_{m,k_m}$, where we have
994: taken that the electron is incident from the left i.e., $x<0$.
995: $\psi_{m,k_m}(x,y)$ is the wavefunction in the region of mode mixing and
996: $\psi_{m,k_m}(x,y)= \sum_n c_n^{(m)}
997: (x,k_n)Sin \frac{n\pi}{W}(y+\frac{W}{2})$.
998: Here $c_n^{(m)} (x,k_n)=C_ne^{ik_nx}$ for $n=1$ and $n=2$ and
999: $c_n^{(m)}(x,k_n)=C_ne^{-\kappa_nx}$ for $n > 2$; $x$ being greater than or
1000: equal to 0. The coefficients $C_n$ can be determined by using the mode
1001: matching technique. The mode matching has been done in details
1002: by Bagwell \cite{bag90}. Here the delta function potential is taken to be
1003: extending from $-\eps$ to $+\eps$ which has to be set to be tending to 0
1004: in the end. $\rho_{0R}$ can be determined by replacing
1005: $\psi_{n,k_n}(x,y)$ by the plane wave states in absence of the scatterer
1006: and doing the integration again.
1007:
1008: Thus we find that for any non-zero incident energy
1009: $$(\rho -\rho_0)_R =
1010: \frac{2}{hv_1}[|t_{13}|^2+|t_{14}|^2+|t_{15}|^2+\cdots ]$$
1011: \be
1012: +\frac{2}{hv_2}[|t_{23}|^2+|t_{24}|^2+|t_{25}|^2+\cdots ].
1013: \ee
1014: Here $v_1=\frac{\hbar k_1}{m}, v_2=\frac{\hbar k_2}{m}$ and
1015: $t_{mn}=\frac{C_n}{A_m}$. $t_{mn}$ can be obtained by solving the matrix
1016: Eqs. given in Ref. \cite{bag90} (see Eq. 23 therein and we have used the
1017: same notation i.e., $t_{mn}$ here is the same as $t_{mn}$ in Eq. (23)
1018: of Ref. \cite{bag90}).
1019: As can be seen in Fig.12 that
1020: $\frac{d\th}{dE}$ is negative over a very large energy range
1021: while $(\rho - \rho_0)_R$ as given by
1022: Eq.43 is positive.
1023:
1024: %%
1025: \PostScript{6}{0}{fig13.eps}{0.5}{14}{
1026: The system under consideration is shown in Fig.7
1027: with $\gamma = 1 $. The Fig. shows some important scattering
1028: probabilities. The solid curve gives $|{\tilde t_{11}}|^2$
1029: and it shows that
1030: for $EW^2>50$, a particle incident in the first propagating
1031: mode does not feel the scatterer at all, and is almost entirely
1032: transmitted intra-channel, $|{\tilde t_{11}}|^2$ being close to unity.
1033: The dotted curve gives $|{\tilde t_{22}}|^2$ and once again for
1034: $EW^2>50$, it is close to unity signifying that a particle incident
1035: in the second propagating channel is almost entirely transmitted
1036: intra-channel. So $EW^2>50$ is the WKB regime where the potential
1037: scatters the incident electron very weakly.
1038: The dashed curve gives 30 times $|{\tilde t_{12}}|^2$
1039: and shows strong energy dependence not only for $EW^2<50$ but also
1040: around the highest energy ($EW^2 \approx 89$)
1041: or in the extreme WKB limit, its
1042: absolute value being extremely small there signifying extremely low
1043: inter-channel transmission i.e., the incoming particle does not
1044: feel the scatterer.
1045: }{f2}
1046: %%%
1047:
1048: %%
1049: \PostScript{6}{0}{fig14.eps}{0.5}{14}{
1050: The system under consideration is shown in Fig.7 with $\gamma=1$.
1051: The solid curve gives $\pi (\rho(E) - \rho_0(E))$ and the dashed
1052: curve gives -0.5$i {d \over dE} ln(det[S])$. The two curves
1053: deviate from each other, where ever the curves in Fig. 13 are
1054: strongly energy dependent. Otherwise they agree.
1055: }{f2}
1056: %%%
1057:
1058:
1059: Note that if we calculate the global DOS by taking the integration
1060: region to be from $-\infty$ to $\infty$ instead of just the region
1061: R where the modes are mixed then Eq. 43 remain the same. One will
1062: get some extra integrals that are indefinite integrals but using
1063: the current conservation condition it can be analytically proved
1064: (see appendix II)
1065: that they cancel each other. This means the contribution to $\rho(E)$
1066: and that to $\rho_0(E)$ coming from outside the region R cancel each
1067: other in the absence of a term like $\rho_q/l$ as in section II.
1068: Thus in this case (the proof is given in appendix II)
1069: $$\rho(E)-\rho_0(E)=(\rho-\rho_0)_R$$
1070: and both of them deviate identically from ${1 \over \pi}{d \theta \over
1071: dE}$ due to strong dispersion, at any arbitrary energy.
1072:
1073: Since the negative slopes are due to the bound states supported
1074: by the negative delta function potential, one may ask what happens for
1075: a positive delta function potential that does not support any bound
1076: state. This situation is discussed below and it
1077: also elaborates the uniqueness of the Q1D,
1078: with respect to the violation of Friedel sum rule and shows that
1079: large violation can occur in the extreme WKB limit. Fig 13 shows
1080: the energy dependence of some important scattering probabilities.
1081: As in 1D the scattering probabilities are strongly energy dependent
1082: for $EW^2<50$. For $EW^2>50$ all
1083: the curves vary slowly with energy (non-dispersive scattering).
1084: However,
1085: one scattering probability $|{\tilde t_{12}}|^2$ is also very
1086: strongly energy dependent at the highest energy of $EW^2 \approx 89$.
1087: ($k_m^2/V>>1$). In this regime scattering is almost entirely intra-channel
1088: as can be seen that $|{\tilde t_{11}}|^2$ and $|{\tilde t_{22}}|^2$ are
1089: both almost unity. All other scattering probabilities like reflection
1090: probabilities and inter-channel scattering probabilities like
1091: $|{\tilde t_{12}}|^2$ are extremely small. There are several other
1092: scattering matrix elements that are identical to ${\tilde t_{12}}$.
1093: Although being small,
1094: they can have strong energy dependence, or their energy derivative
1095: can be very large. As can be seen in Fig. 13 that the dashed curve
1096: bends down with a steep slope at $EW^2 \approx 89$.
1097: It is found in Fig. 14 that there is a strong violation
1098: of Friedel sum rule in the non-WKB regime as in 1D and also in the
1099: extreme WKB regime ($EW^2 \approx 89$ where $k^2/V>>1$ and scattering
1100: is almost completely intra-channel)
1101: quite unlike that in 1D. It can be seen in Fig. 13
1102: that in the mid-energy range, the scattering probabilities are
1103: not very energy dependent and the Friedel sum rule holds in this
1104: mid-energy range as can be seen in Fig. 14. But at regimes
1105: where the scattering leads to strong dispersion, FSR is violated.
1106: As the strength of the
1107: positive delta function potential is increased, this regime where
1108: the scattering probabilities are not strongly energy dependent becomes
1109: narrower and Friedel sum rule is violated at all energies.
1110:
1111:
1112: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1113: \section{ phase behavior at critical energies }
1114: \label{s4}
1115: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1116:
1117: Very interesting phase behaviors can be seen at energies where the
1118: $S$-matrix changes dimension. For example for $E\le \frac{4\pi ^2}{W^2}$
1119: there is only one propagating mode and the $S$-matrix is 2$\times$2. But
1120: for $E > \frac{4\pi ^2}{W^2}$, there are two propagating modes and the
1121: $S$-matrix is 4$\times$4. The matrix element $t_{11}^{'}$ exists on either
1122: side of the energy $\frac{4\pi ^2}{W^2}$ and in Fig.15 we show the
1123: behavior of $arg(t_{11}^{'})$ in the energy range that includes $EW^2 =
1124: 4\pi ^2 $. Note that it exhibits a discontinuous phase drop by
1125: $\frac{\pi}{2}$ at $EW^2 = 4\pi ^2 $. So far only discontinuous phase
1126: drops of $\pi$ has been observed but never $\frac{\pi}{2}$.
1127: From the properties of a 2$\times$2 $S$-matrix it follows that if there
1128: is a discontinuous phase change then it can only be of $\pi$ \cite
1129: {lee99,tan99}. So had the $S$-matrix been 2$\times$2 on either side of
1130: $EW^2 = 4\pi ^2$ the phase drop would have been $\pi$. But since the
1131: $S$-matrix is 2$\times$2 only on one side, including $EW^2 = 4\pi ^2$,
1132: i.e., $E\le \frac{4\pi ^2}{W^2}$, the phase drop is also one half of
1133: $\pi$. $|t_{11}^{'}|^2$ also has a zero at $EW^2 = 4\pi ^2$ for all
1134: possible choice of parameters \cite{bag90}, and this zero is associated
1135: with a $\frac{\pi}{2}$ phase jump instead of a $\pi$ phase jump.
1136:
1137: %%%
1138: \PostScript{6}{0}{fig15.ps}{0.5}{14}{
1139: \hspace{-.5cm}
1140: The system under consideration is shown in the Fig.7. The plot is of
1141: $arg(t_{11}^{'})$ in radians versus E$W^2$. This plot is for
1142: $\gamma=-25.197$, $x_i=0$ and $y_i=.45W$.
1143: }{f2}
1144: %%%
1145:
1146: %%%
1147: \PostScript{6}{0}{fig16.ps}{0.5}{14}{
1148: \hspace{-.5cm}
1149: The system under consideration is shown in the Fig.7. The solid curve
1150: gives $|t_{13}^{'}|^2$. The dashed curve is after subtracting 5.27183
1151: radians from $arg(t_{13}^{'})$ in radians. We use $\gamma=+25.197$,
1152: $x_i=0$ and $y_i=.45W$.
1153: }{f2}
1154: %%%
1155:
1156: Next we take a repulsive $\delta$ function potential. It is known
1157: \cite{bag90} that at critical energies like $EW^2=4\pi^2$,
1158: $|t_{13}^{'}|^2$ shows discontinuities. Here $|t_{13}^{'}|^2$ does not
1159: have a zero but exhibits a discontinuous jump. At these points
1160: $arg(t_{13}^{'})$ also shows non-analytic behavior as demonstrated in
1161: Fig.16. In this case $\frac{d}{dE} arg(t_{13}^{'})$ is discontinuous.
1162:
1163: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1164: \section{Conclusions}
1165: \label{s5}
1166: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1167:
1168: In a multichannel quantum wire with attractive impurities, negative slopes
1169: in the scattering phase shift versus incident energy curves can occur at
1170: all possible energies. For weaker defects it happens at higher energies
1171: and the negative slopes are more pronounced.
1172: Such negative slopes
1173: mean super luminescence \cite{smi59,kumar} that can be observed
1174: experimentally. Although such a super luminescent particle will not give
1175: any information about the particle delay or information delay, they are of
1176: interest because they demonstrate fundamental principles in quantum
1177: mechanics. Hence Eq. (41) derived in this paper may be of use to
1178: experimentalists and theoreticians. Calculations of variation of DOS due
1179: to the presence of the scatterer show that FSR can be violated at any
1180: arbitrary energy. That is violation of FSR is not restricted to
1181: non-WKB regime (the regime where anyway transport does not occur)
1182: as in 1D, 2D or 3D.
1183: Rather the violation is related to strong energy dependence of the
1184: scattering matrix elements (dispersive behavior).
1185: While in 1D, 2D and 3D the strong
1186: energy dependence of scattering occur only in the non-WKB regime,
1187: in Q1D there is no systematics. A Q1D system in the extreme
1188: WKB limit can also exhibit strongly energy dependent scattering
1189: and there is no definable regime where the FSR will work and
1190: hence it may not work even in regimes where transport occurs.
1191: For attractive impurities, weaker the impurity, stronger the
1192: violation of FSR while for repulsive impurities, stronger the
1193: impurity, stronger the violation of FSR.
1194: We also show that the discontinuous phase drops in the
1195: single channel case have a counterpart in the multichannel case wherein
1196: the drops can be continuous and we propose a line shape formula for them
1197: in Eq. 34. When there is a third channel of escape
1198: for the electron, apart from the channel along which
1199: it is incident and the channel where its scattering
1200: phase shift is measured, the phase drop becomes
1201: continuous. However, these phase drops do not
1202: affect ${1 \over 2i} ln Det [S]$ and hence Friedel sum rule.
1203: Finally, we discuss some novel scattering phase shifts at
1204: energies where the $S$ matrix changes dimension.
1205:
1206: \section{ACKNOWLEDGMENTS}
1207:
1208: We thank Dr. A. M. Jayannavar for useful discussions.
1209:
1210:
1211: \section{APPENDIX I}
1212:
1213: Considering the symmetric scattering potential in Fig.4, the scattering
1214: matrix $ S $ of the structure can be found by cascading the scattering
1215: matrices of different parts, i.e.,\\
1216: \[ S = \left(\begin{array}{cc}
1217: \displaystyle r & \displaystyle t \\
1218: \displaystyle t & \displaystyle r
1219: \end{array}\right) =S_{1} \otimes S_{2} \otimes S_{3} ,\]
1220: \[ \mbox{where} \hspace{.4cm} S_{1} = S_{3} = \left(\begin{array}{cc}
1221: \displaystyle r' & \displaystyle t' \\
1222: \displaystyle t' & \displaystyle r'
1223: \end{array}\right) \]
1224: \hfill\\
1225: \[ \mbox{and} \hspace{.4cm} S_{2} = \left(\begin{array}{cc}
1226: 0 & \displaystyle \tau \\
1227: \displaystyle \tau & 0
1228: \end{array}\right)\hspace{.1cm}.\]
1229: Here $\tau = e^{i \phi}$, $\phi = kl$ and $k = \sqrt{\frac{2m}{\hbar^{2}}
1230: E }$. $S_{2}$ is the scattering matrix for the free region II of length
1231: $l$ between the two scatterers. $ r' $ \& $ t' $ are the reflection \&
1232: transmission amplitudes due to one of the two potentials when isolated.
1233:
1234: After cascading these three matrices the resultant scattering matrix of
1235: our system becomes
1236: \[ S = \left(\begin{array}{cc}
1237: \displaystyle r'+\frac{t'^{2}{\tau}^{2}r'}{1-r'^{2}{\tau}^{2}} &
1238: \displaystyle \frac{t'^{2}{\tau}}{1-r'^{2}{\tau}^{2}}\\
1239: \displaystyle \frac{t'^{2}{\tau}}{1-r'^{2}{\tau}^{2}} &
1240: \displaystyle r'+\frac{t'^{2}{\tau}^{2}r'}{1-r'^{2}{\tau}^{2}}
1241: \end{array}\right) \hspace{.1cm}. \]
1242: And so,
1243: \[ det [S] = \left(\displaystyle r'+\frac{t'^{2}{\tau}^{2}r'}{1-r'^{2}{\tau}
1244: ^{2}}\right)^{2} - \left(\displaystyle \frac{t'^{2}{\tau}}{1-r'^{2}{\tau}^{2}}
1245: \right)^{2} \hspace{.5cm} \mbox{(i)}\]
1246: \[ \hspace{1cm}= \frac{1}{(1-r'^{2}{\tau}^{2})^2} \hspace{.1cm} (M+N)
1247: \hspace{.1cm},\hspace{2.5cm} \mbox{(ii)}\]
1248: \[ \mbox{where}\hspace{1cm}M = r'^{2}(1-r'^{2}{\tau}^{2})^2 \hspace{2.0cm}
1249: \mbox{(iii)} \]
1250: \[ \mbox{and} \hspace{.5cm} N = (1-r'^{2}{\tau}^{2})\hspace{.1cm} (2t'^{2}r'^{2}
1251: {\tau}^{2}-t'^{4}{\tau}
1252: ^{2}) . \hspace{1cm} \mbox{(iv)} \]
1253: From (i),
1254: \[ \frac{\pa det [S] }{\pa \phi} = 2\left( A \frac{\pa A}{\pa \phi} -B
1255: \frac{\pa B}{\pa \phi} \right) , \hspace{2.0cm}\mbox{(v)}\]
1256: \[ \mbox{where}\hspace{1cm} A = r'+\frac{t'^{2}
1257: {\tau}^{2}r'}{1-r'^{2}{\tau}^{2}} \hspace{2cm}\mbox{(vi)}\]
1258: \[ \mbox{and} \hspace{.5cm} B =
1259: \frac{t'^{2}{\tau}}{1-r'^{2}{\tau}^{2}}\hspace{.5cm}
1260: .\hspace{2cm}\mbox{(vii)}\]
1261: Using (v), (vi) and (vii) we find
1262: $$ \hspace{-5cm}\frac{\pa det [S] }{\pa \phi} = $$
1263: \[ \hspace{1.0cm} 2\left[A
1264: \left(2r'B+\frac{2r'^{3}{\tau}^{2}
1265: B}{1-r'^{2}{\tau}^{2}}\right) -
1266: B\left(\frac{t'^{2}+r'^{2}t'^{2}{\tau}^{2}}
1267: {(1-r'^{2}{\tau}^{2})^2}\right)\right] \frac{\pa \tau }{\pa \phi} \]
1268: \[ \hspace{1cm} + 2\left[A
1269: \left(1+B{\tau}+\frac{2r'^{2}
1270: {\tau}^{3}B}{1-r'^{2}{\tau}^{2}}\right)-B
1271: \left(\frac{2r'{\tau}^{2}B}{1-r'^{2}{\tau}^{2}}\right)\right] \frac{\pa r'}{\pa \phi} \]
1272: \[ \hspace{1cm} + 2\left[2r'{\tau}AB
1273: -B\left(\frac{2{\tau}t'}
1274: {1-r'^{2}{\tau}^{2}}\right)\right] \frac{\pa t'}{\pa \phi} .\hspace{1.5cm} \mbox
1275: {(viii)} \]
1276:
1277: We will neglect the last two terms in comparison to the 1st term in
1278: Eq.(viii) but we will retain the energy dependence of $ r' $ and $ t' $
1279: and the 1st term in Eq.(viii) is actually varying with energy very
1280: strongly. Hence our results correspond to real potentials and we do not
1281: parameterize the $ S $ matrix in a special way. We apply this result to
1282: the case of double delta function potential in Fig.5 and illustrate the
1283: significance of the last two terms in comparison with the 1st one. We
1284: stress that this calculation in this appendix holds even if the $\delta
1285: $-function potential is replaced by the square-well or any arbitrary
1286: potential. Thus in the regime where $ \frac{\pa r'}{\pa \phi} \to 0 $ and
1287: $\frac{\pa t'}{\pa
1288: \phi} \to 0 $ and using $\frac{\pa \tau }{\pa \phi} = i\tau$,
1289: $$ \hspace{-7cm} \frac{\pa det [S] }{\pa \phi} $$
1290: \[ = 2\left[A\left(2r'B+\frac{2r'^{3}{\tau}^{2}
1291: B}{1-r'^{2}{\tau}^{2}}\right) -
1292: B\left(\frac{t'^{2}+r'^{2}t'^{2}{\tau}^{2}}
1293: {(1-r'^{2}{\tau}^{2})^2}\right)\right] i\tau \hspace{.2cm}. \]
1294: At this point we substitute the values of $A$ and $B$ from (vi) and (vii)
1295: to get
1296: \[ \frac{\pa det [S] }{\pa \phi} = 2i\hspace{.2cm}
1297: \frac{1}{(1-r'^{2}{\tau}^{2})^3}\hspace{.2cm}N \hspace{.2cm}. \hspace{1.5cm}
1298: \mbox{(ix)}\]
1299: \[ \mbox{Now},\hspace{.2cm} \th = \frac{1}{2i} ln (det [S] )\hspace{4cm}
1300: \mbox{(x)} \]
1301: \[ \mbox{and so} \hspace{.4cm} \frac{\pa \th}{\pa E} = \frac{\pa \th}{\pa
1302: \phi} \hspace{.2cm}
1303: \frac{\pa \phi}{\pa E} . \]
1304: From (x), (ii) and (ix)\\
1305: \[ \frac{\pa \th}{\pa \phi} = \frac{1}{2i} \hspace{.2cm} \frac{1}{det [S] }
1306: \hspace{.2cm} \frac{\pa det [S] }{\pa \phi} \]
1307: \[ \hspace{1.5cm} = \frac{1}{1-r'^{2}{\tau}^{2}}\hspace{.2cm}\frac{1}
1308: {\frac{M}{N}+1} \hspace{.2cm}. \]
1309: Multiplying the numerator and denominator by $\frac{(1-r'^{2}{\tau}^{2})^{*}}
1310: {1-\mid r^{'} \mid^{4}}$, we get
1311: \[ \hspace{.2cm} \frac{\pa \th}{\pa \phi} = \frac{1-\mid r' \mid ^{4}}
1312: {\mid 1-r'^{2}{\tau}^{2} \mid
1313: ^{2}} \hspace{.2cm} \frac{(1-r'^{2}{\tau}^{2})^{*}}{1-\mid r' \mid ^{4}}
1314: \hspace{.2cm}\frac{1}{\frac{M}{N}+1} \]
1315: \[ \hspace{.2cm} = \frac{1-\mid r' \mid ^{4}}{\mid 1-r'^{2}{\tau}^{2} \mid
1316: ^{2}} \hspace{.2cm}Q \hspace{.2cm},\hspace{4cm}\mbox{(xi)}\]
1317: \[ \mbox{where,} \hspace{.5cm} Q=\frac{(1-r'^{2}{\tau}^{2})^{*}}{1-\mid r'
1318: \mid ^{4}}
1319: \hspace{.2cm}\frac{1}{\frac{M}{N}+1} \hspace{2cm} \mbox{(xii)}\]
1320: \[ \hspace{.2cm} =\left[\frac{(1-{r'^{*}}^{2}{{\tau}^{*}}^{2})-1+\mid
1321: r'\mid^{4}\mid
1322: \tau \mid^{4}}{1-\mid r' \mid^{4}}+1\right]\hspace{.2cm}
1323: \frac{1}{\frac{M}{N}+1} \hspace{.2cm},\]
1324: \[ \hspace{5cm} \mbox{as}\hspace{.4cm} |\tau |^4 = 1\hspace{.2cm}. \]
1325: As \hspace{.5cm} $1-|r'|^2=|t'|^2$
1326: \[ \hspace{.2cm} Q =\left[\frac{-{r'^{*}}^{2}{{\tau}^{*}}^{2}+\mid r'\mid^{4}
1327: \mid \tau \mid^{4}}{(1+\mid r' \mid^{2})|t'|^2}+1\right]\hspace{.2cm}\frac{1}
1328: {\frac{M}{N}+1} \hspace{.2cm}.\]
1329: Now substituting the values of $M$ and $N$ from (iii) and (iv)
1330: \[ Q = \frac{\frac{-{r'^{*}}^2{{\tau}^{*}}^2(1-r'^{2}
1331: {\tau}^{2})+\mid t' \mid ^{2}(1+\mid r' \mid ^{2})}{\mid t' \mid ^{2}
1332: (1+\mid r' \mid ^{2})}}{
1333: \frac{r'^{2}(1-r'^{2}{\tau}^{2})-t'^{2}{\tau}^{2}(t'^{2}-2r'^{2})}{-t'^{2}
1334: {\tau}^{2}(t'^{2}-2r'^{2})}} \hspace{.2cm}.\]
1335: \hfill\\
1336: Using, \hspace{.4cm} $ r' = \mid r' \mid e^{i\th _{r}} $ \hspace{.4cm} and
1337: \hspace{.4cm} $ t'= \mid t' \mid e^{i\th _{t}} $,
1338: $$Q=[|r'|^2|t'|^2|\tau |^4e^{2i\th _{t}}e^{2i(\th _{t}-\th _{r})}
1339: -|r'|^4|t'|^2|\tau
1340: |^4\tau ^2e^{4i\th _{t}}$$
1341: $$-|t'|^4\tau^2e^{4i\th _{t}}
1342: -|t'|^4|r'|^2\tau^2e^{4i
1343: \th _{t}}$$
1344: $$-2|r'|^4|\tau |^4e^{2i\th _{t}}
1345: +2|r'|^6|\tau |^4\tau^2e^{2i(\th _{t}
1346: +\th _{r})}$$
1347: $$+2|r'|^2|t'|^2\tau^2e^{2i(\th _{t}+\th _{r})}+2|r'|^4||t'|^2\tau^2
1348: e^{2i(\th _{t}+\th _{r})}]/D ,$$
1349: \[\mbox{where},\hspace{.2cm} D = (1+\mid r' \mid^{2})(-\mid t' \mid ^{4}{\tau}
1350: ^{2}e^{4i\th _{t}}-\mid r' \mid ^{4}{\tau}^{2}e^{4i\th _{r}}\]
1351: \[ \hspace{.2cm} +2\mid t' \mid ^{2}\mid r' \mid ^{2}\tau ^2 e^{2i
1352: (\th _{r}+\th _{t})}+\mid r' \mid ^{2}e^{2i\th _{r}}) . \hspace{1cm}
1353: \mbox{(xiii)}\]
1354: \hfill\\
1355: Now it follows from unitarity that
1356: $e^{2i(\th _t - \th _r)} = e^{i \pi} = -1$ and
1357: $|\tau |^4 = 1$ and so
1358: \hfill\\
1359: \[ Q = [-|r'|^2 |t'|^2 e^{2i \th _t } - |r'|^4 |t'|^2 \tau ^2 e^{4i \th _t }
1360: -|t'|^4 \tau ^2 e^{4i \th _t } \]
1361: \[ - |t'|^4 |r'|^2 \tau ^2 e^{4i \th _t } - 2
1362: |r'|^4 e^{2i \th _t } + 2 |r'|^6 \tau ^2 e^{2i ( \th _t + \th _r)} \]
1363: \[ + 2 |r'|^2 |t'|^2 \tau ^2 e^{2i ( \th _t + \th _r)} + 2 |r'|^4 |t'|^2
1364: \tau ^2 e^{2i ( \th _t + \th _r)}]/D \]
1365: \[\hspace{1cm} = [- e^{2i \th _t } |r'|^2 \{|t'|^2 + 2 |r'|^2 \} - (|t'|^2 +
1366: |r'|^2 \{|r'|^2 + |t'|^2\}) \]
1367: \[ (e^{4i \th _t } |t'|^2 \tau ^2 - e^{2i ( \th _t +
1368: \th _r)} 2 |r'|^2 \tau ^2 )]/D \hspace{.2cm}.\]
1369: Now inside the $\{ \} $ brackets if we use the fact that $ \mid r' \mid ^{2}+
1370: \mid t' \mid ^{2}=1$, then\\
1371: \[ Q = [- e^{2i \th _t } |r'|^2 (1+|r'|^2) - e^{4i \th _t }|t'|^2 \tau ^2
1372: + 2 e^{2i (\th _t + \th _r)} |r'|^2 \tau ^2]/D .\]
1373: Multiplying numerator and denominator above by $e^{-2i(\th _{r}+\th _{t})}$ and
1374: putting $ e^{2i (\th _t - \th _r)} = e^{i \pi} = -1$, we get
1375: \[ Q = \frac{-e^{-2i \th _{r}} |r'|^2 (1+|r'|^2 )+\tau ^2 (|t'|^2 +|r'|^2)+
1376: |r'|^2
1377: \tau ^2}{D'} ,\]
1378: \[ \mbox{where,}\hspace{.1cm} D'= De^{-2i(\th _{r}+\th _{t})} .\hspace{2cm}
1379: \mbox{(xiv)} \]
1380: Again using $|r'|^2 + |t'|^2 =1$,
1381: \[ Q = \frac{(1+|r'|^2 ) (\tau ^2 - |r'|^2 e^{-2i\th _{r}})}{D'} . \hspace{1cm}
1382: \mbox{(xv)} \]
1383: Now from (xiv) and (xiii)
1384: \[ D' = (1+|r'|^2)[|r'|^2 e^{-2i \th _t} - |r'|^4 \tau ^2 e^{-2i(\th _t -
1385: \th _r)} \]
1386: \[-|t'|^4 \tau ^2 e^{2i(\th _t - \th _r)} + 2|t'|^2 |r'|^2 \tau ^2] . \]
1387: As $\th _t - \th _r = \frac{\pi}{2}$,
1388: \[ \hspace{2cm} D' = (1+|r'|^2)[|r'|^2 e^{-2i \th _t} + \tau ^2 (|r'|^2 +
1389: |t'|^2 )^2 ] \]
1390: where of course $\mid r' \mid^{2}+\mid t' \mid ^{2}=1$.
1391: Substituting $D'$ in Eq.(xv),
1392: \[ Q = \frac{{\tau}^{2} - \mid r' \mid ^{2}e^{-2i\th _{r}}}{{\tau}^{2}+\mid
1393: r' \mid ^{2}e^{-2i\th _{t}}} .\]
1394: Multiplying numerator and denominator of $Q$ by
1395: $e^{2i\th_{t}}$ and using $e^{2i
1396: (\th _t - \th _r)} = e^{i \pi} = -1$,
1397: we get from (xi)
1398: \[ \frac{\pa \th}{\pa \phi} =\frac{1-\mid r' \mid ^{4}}{\mid 1-r'^{2}{\tau}
1399: ^{2} \mid
1400: ^{2}} \]
1401:
1402:
1403: \section{APPENDIX II}
1404:
1405: The global density of states is given by
1406: \bea
1407: \rho(E)&=&\int_{-\infty}^\infty dx \int_{\frac{-W}{2}}^{\frac{W}{2}}dy\nn\\
1408: && \sum_{m,k_m}\left|\psi_{m,k_m}(x,y)\right|^2 \delta (E-E_{m,k_m})
1409: \hspace{1.6cm}\mbox{(i)}\nn
1410: \eea
1411: where \hspace{0.2cm} $\psi_{m,k_m}(x,y) = \sum_n c_n^{(m)}(x)\chi_n(y)$
1412: and $E_{m,k_m}$ is the energy of an electron in the leads.\\
1413: \hfill\\
1414: $E_{m,k_m}=\displaystyle \frac{m^2\pi^2\hbar ^2}{2m_eW^2}+ \frac{\hbar ^2k_m^2}{2m_e},\hspace{0.2cm}\mbox{where}\hspace{0.2cm}m=\pm 1, \pm 2,$
1415: as there are two propagating modes in the leads.\\
1416: \hfill\\
1417: As $\chi_n(y)$'s form an orthonormal set,
1418: $$
1419: \rho(E)=\sum_{m,k_m}\delta (E-E_{m,k_m})\int_{-\infty}^\infty
1420: dx \sum_n \left|c_n^{(m)}(x)\right|^2$$
1421: \bea
1422: \rho(E)&=&\frac{2}{h v_1}\int_{-\infty}^\infty dx
1423: \sum_n \left|c_n^{(1)}(x)\right|^2\nn\\
1424: &+& \frac{2}{h v_2}\int_{-\infty}^\infty dx \sum_n \left|c_n^{(2)}(x)\right|^2
1425: \hspace{2.5cm}\mbox{(ii)}\nn
1426: \eea
1427: Here, $v_1 = \displaystyle \frac{\hbar k_1}{m_e}$ and $v_2 =
1428: \displaystyle \frac{\hbar k_2}{m_e}$.\\
1429: \hfill\\
1430: Now,
1431: \bea
1432: \int_{-\infty}^\infty dx \sum_n \left|c_n^{(1)}(x)
1433: \right|^2 &=& \int_{-\infty}^\infty \left|c_1^{(1)}(x)\right|^2 dx \nn\\
1434: &+& \int_{-\infty}^\infty \left|c_2^{(1)}
1435: (x)\right|^2 dx\nn\\
1436: &+& \int_{-\infty}^\infty \left|c_3^{(1)}(x)\right|^2 dx +
1437: \cdots \nn\\
1438: &=& T1 \mbox{(say)},\nn
1439: \eea
1440: where electron is incident in the fundamental mode,
1441: \bea
1442: c_1^{(1)}(x) &=& e^{ik_1x} +
1443: {\tilde r_{11}} e^{-ik_1x}
1444: \hspace{.2cm}\mbox{for}\hspace{.2cm} x < 0 \nn\\
1445: &=& {\tilde t_{11}} \hspace{.2cm}e^{ik_1x}
1446: \hspace{.2cm}\mbox{for}\hspace{.2cm}
1447: x > 0 \nn\\
1448: c_2^{(1)}(x) &=& {\tilde r_{12}}
1449: \hspace{.2cm}e^{-ik_2x}
1450: \hspace{.2cm}\mbox{for}\hspace{.2cm} x < 0 \nn\\
1451: &=& {\tilde t_{12}} \hspace{.2cm}e^{ik_2x}
1452: \hspace{.2cm}\mbox{for}\hspace{.2cm}
1453: x > 0 \nn
1454: \eea
1455: and for $n > 2$,
1456: \bea
1457: c_n^{(1)}(x) &=& t_{1n} \hspace{.2cm}e^{\kappa_n x}\hspace{.2cm}
1458: \mbox{for}\hspace{.2cm}x < 0 \nn\\
1459: &=& t_{1n} \hspace{.2cm}e^{-\kappa_n x}\hspace{.2cm}\mbox{for}
1460: \hspace{.2cm}x >
1461: 0 \nn
1462: \eea
1463: \bea
1464: T1 &=& \int_{-\infty}^0 dx
1465: \left[1+|{\tilde r_{11}}|^2+2|{\tilde r_{11}}| \hspace{.2cm}Cos(2k_1x+\eta_1)
1466: \right] \nn\\
1467: &+& \int_0^\infty dx|{\tilde t_{11}}|^2
1468: + \int_{-\infty}^0 dx|{\tilde r_{12}}|^2 +
1469: \int_0^\infty dx|{\tilde t_{12}}|^2\nn\\
1470: &+&|t_{13}|^2 + |t_{14}|^2 + \cdots \cdots \nn
1471: \eea
1472: Here,\hspace{.3cm}$\eta_1$ is defined as $\displaystyle{\tilde r_{11}}=|{\tilde r_{11}}|e^{-i\eta_1}$\\
1473: \hfill\\
1474: Similarly, for electron incident in the first excited mode,\\
1475: \bea
1476: T2 &=& \int_{-\infty}^\infty dx \sum_n \left|c_n^{(2)}(x)\right|^2\nn\\
1477: &=& \int_{-\infty}^0 dx \left[1+|{\tilde r_{22 }}|^2+2|{\tilde r_{22}}| \hspace{.2cm}Cos(2k_2x+\eta_2)\right]\nn\\
1478: &+& \int_0^\infty dx|{\tilde t_{22}}|^2 + \int_{-\infty}^0 dx|{\tilde r_{21}}|
1479: ^2 + \int_0^\infty dx|{\tilde t_{21}}|^2\nn\\
1480: &+&|t_{23}|^2 + |t_{24}|^2 + \cdots \cdots ,\nn
1481: \eea
1482: Here,\hspace{.3cm}$\eta_2$ is defined as $\displaystyle{\tilde r_{22 }}= |{\tilde r_{22 }}| e^{-i\eta_2}$\\
1483: \bea
1484: \mbox{Therefore,}\nn\\
1485: \rho (E)&=& 2\left[\frac{1+|{\tilde r_{11}}|^2}{h v_1}\int_{-\infty}^0 dx \hspace{.2cm}+\hspace{.2cm} \frac{1+|{\tilde r_{22}}|^2}{h v_2}\int_{-\infty}^0 dx \right.\nn\\
1486: &+& \left. \frac{|{\tilde r_{12}}|^2}{h v_1}\int_{-\infty}^0 dx +
1487: \hspace{.02cm}\frac{|{\tilde r_{21}}|^2}{h v_2}\int_{-\infty}^0 dx \right.\nn\\
1488: &+& \left. \frac{2|{\tilde r_{11}}|}{h v_1}\int_{-\infty}^0 dx \hspace{.2cm}Cos(2k_1x+\eta_1)\right. \nn \\
1489: & + & \left. \hspace{.02cm}\frac{2|{\tilde r_{22}}|}{h v_2}\int_{-\infty}^0 dx \hspace{.2cm}Cos(2k_2x+\eta_2)\right. \nn \\
1490: &+& \left. \frac{|{\tilde t_{11}}|^2+|{\tilde t_{12}}|^2}{h v_1}\int_0^{\infty}dx + \frac{|{\tilde t_{22}}|^2+|{\tilde t_{21}}|^2}{h v_2}\int_0^{\infty}dx \right. \nn \\
1491: &+& \left.\frac{1}{h v_1}\left(|t_{13}|^2+|t_{14}|^2+ \cdots \right)\right. \nn \\
1492: &+& \left.\frac{1}{h v_2}\left(|t_{23}|^2+|t_{24}|^2+ \cdots \right) \right] \nn
1493: \eea
1494: Due to time reversal symmetry, ${\tilde r_{12}}={\tilde r_{21}}$ \& ${\tilde t_{12}}={\tilde t_{21}}$.\\
1495: Inside the square bracket $\displaystyle[\hspace{.2cm}]$,\\
1496: in the 3rd term we put ${\tilde r_{12}}={\tilde r_{21}}$, \\
1497: in the 4th term we put ${\tilde r_{21}}= {\tilde r_{12}}$,\\
1498: in the 7th term we put ${\tilde t_{12}}={\tilde t_{21}}$, \\
1499: in the 8th term we put ${\tilde t_{21}}= {\tilde t_{12}}$.
1500: \bea
1501: \mbox{Therefore},\nn\\
1502: \rho (E) &=& 2\left[\frac{1+|{\tilde r_{11}}|^2+|{\tilde r_{21}}|^2}{h v_1}\int_{-\infty}^0 dx \right.\nn\\
1503: &+&\left. \frac{1+|{\tilde r_{12}}|^2+|{\tilde
1504: r_{22}}|^2}{h v_2}\int_{-\infty}^0 dx \right.\nn \\
1505: & + &\frac{|{\tilde t_{11}}|^2+|{\tilde t_{21}}|^2}{h v_1}\int_0^{\infty}dx
1506: \hspace{.02cm}+\hspace{.02cm}\frac{|{\tilde t_{12}}|^2+|{\tilde t_{22}}|^2}{h v_2}\int_0^{\infty}dx\nn \\
1507: & + &\frac{2|{\tilde r_{11}}|}{h v_1}\int_{-\infty}^0 dx \hspace{.2cm}Cos(2k_1x+\eta_1)\nn \\
1508: &+&\frac{2|{\tilde r_{22}}|}{h v_2}\int_{-\infty}^0 dx \hspace{.2cm}Cos(2k_2x+\eta_2)\nn \\
1509: & + & \frac{1}{h v_1}\left(|t_{13}|^2\hspace{.02cm}+\hspace{.02cm}|t_{14}|^2 \hspace{.02cm}+ \cdots \right)\nn \\
1510: & + &\left. \frac{1}{h v_2}\left(|t_{23}|^2\hspace{.02cm}+\hspace{.02cm}|t_{24}|^2\hspace{.02cm}+ \cdots \right)\right]\nn
1511: \eea
1512: Now adding and subtracting the following terms inside $\displaystyle[\hspace{.2cm}]$,\\
1513: \hfill\\
1514: $\displaystyle{\frac{|{\tilde t_{11}}|^2}{h v_1}\int_{-\infty}^0 dx}$,
1515: $\displaystyle{\frac{|{\tilde t_{21}}|^2}{h v_1}\int_{-\infty}^0 dx}$,
1516: $\displaystyle{\frac{|{\tilde t_{22}}|^2}{h v_2}\int_{-\infty}^0 dx}$,
1517: $\displaystyle{\frac{|{\tilde t_{12}}|^2}{h v_2}\int_{-\infty}^0 dx}$,
1518: \bea
1519: \mbox{we get},\nn\\
1520: \rho (E) &=& 2\left[\frac{1 + |{\tilde r_{11}}|^2 + |{\tilde r_{21}}|^2 + |{\tilde t_{11}}|^2 + |{\tilde t_{21}}|^2}{h v_1}\int_{-\infty}^0 dx \right. \nn\\
1521: & + & \frac{1+|{\tilde r_{12}}|^2+|{\tilde r_{22}}|^2+|{\tilde t_{12}}|^2+
1522: |{\tilde t_{21}}|^2}{h v_2}\int_{-\infty}^0 dx\nn\\
1523: & + &\frac{2|{\tilde r_{11}}|}{h v_1}\int_{-\infty}^0 dx \hspace{.2cm}Cos(2k_1x+\eta_1)\nn\\
1524: & + &\frac{2|{\tilde r_{22}}|}{h v_2}\int_{-\infty}^0 dx \hspace{.2cm}Cos(2k_2x+\eta_2)\nn \\
1525: & + & \frac{1}{h v_1}\left(|t_{13}|^2\hspace{.02cm}+\hspace{.02cm}|t_{14}|^2 \hspace{.02cm}+ \cdots \right)\nn \\
1526: & + & \left. \frac{1}{h v_2}\left(| t_{23}|^2\hspace{.02cm}+\hspace{.02cm}|t_{24}|^2\hspace{.02cm}+ \cdots \right) \right]\nn
1527: \eea
1528:
1529: Now \hspace{.2cm}$|{\tilde r_{11}}|^2 + |{\tilde r_{21}}|^2 + |{\tilde t_{11}}|^2 + |{\tilde t_{21}}|^2=1$ \\
1530:
1531: and \hspace{.2cm}$|{\tilde r_{12}}|^2+|{\tilde r_{22}}|^2+|{\tilde t_{12}}|^2+
1532: |{\tilde t_{21}}|^2=1$ \\
1533: \bea
1534: \mbox{Thus},\nn\\
1535: \rho (E) & = & \frac{2}{h v_1}\int_{-\infty}^{\infty}dx \hspace{.2cm} + \hspace{.2cm}\frac{2}{h v_2}\int_{-\infty}^{\infty}dx\nn\\
1536: & + &\frac{2|{\tilde r_{11}}|}{h v_1}\int_{-\infty}^{\infty} dx
1537: \hspace{.2cm}Cos(2k_1x+\eta_1)\nn\\
1538: & + &\frac{2|{\tilde r_{22}}|}{h v_2}\int_{-\infty}^{\infty} dx \hspace{.2cm}Cos(2k_2x+\eta_2)\nn \\
1539: & + &\frac{2}{h v_1}\left(|t_{13}|^2\hspace{.02cm}+\hspace{.02cm}| t_{14}|^2 \hspace{.02cm}+ \cdots \right)\nn \\
1540: & + &\frac{2}{h v_2}\left(t_{23}|^2\hspace{.02cm}+\hspace{.02cm}|t_{24}|^2\hspace{.02cm}+ \cdots \right)\hspace{2.0cm}\mbox{(iii)}\nn
1541: \eea
1542: Now $\displaystyle \frac{2}{h v_1}\int_{-\infty}^{\infty}dx + \frac{2}{h v_2}\int_{-\infty}^{\infty}dx = \rho_0 (E)$ i.e. DOS in the absence of scatterer.
1543: Also according to some text books \cite{gha},
1544: for plane wave states the current conservation condition
1545: $$\frac{\pa}{\pa t}\int_{\Omega} \psi^*\psi \hspace{.2cm}d\tau \hspace{.2cm}+\hspace{.2cm}\int_{S} \vec J.\hat n \hspace{.2cm}ds \hspace{.2cm}=\hspace{.2cm}0$$
1546: can be satisfied if and only if the wave function vanishes at $\pm \infty$ for
1547: all energy. Thus the 3rd and 4th term of Eq.(iii) of this
1548: section is $0$. Besides, neglecting these oscillatory
1549: terms in the leads is very standard in the derivation of Friedel
1550: sum rule \cite{yeycm}, where it is assumed that the carrier
1551: concentration is so high in the leads that the leads
1552: are non-polarizable. However, assuming non-polarizable leads
1553: alone does not eradicate all energy dependence of the self energy
1554: as thought in Ref. \cite{yeycm} (see Eqs. 2 and 6 therein, the
1555: parameter $t_\alpha$ can be strongly energy dependent, which
1556: is dispersive behavior).
1557: So we get\\
1558: \bea
1559: \rho (E) & = &\rho_0 (E)\hspace{.2cm}+\hspace{.2cm}\frac{2}{h v_1}\left(|t_{13}|^2\hspace{.2cm}+\hspace{.2cm}|t_{14}|^2 \hspace{.2cm}+ \cdots \right)\nn\\
1560: &+&\frac{2}{h v_2}\left(|t_{23}|^2\hspace{.2cm}+\hspace{.2cm t_{24}}|^2\hspace{.2cm}+ \cdots \right)\nn
1561: \eea
1562: \bea
1563: \mbox{Thus} \hspace{.4cm} \rho(E)-\rho_0(E)&=& \frac{2}{h v_1}\left(|t_{13}|^2\hspace{.2cm}+\hspace{.2cm}|t_{14}|^2 \hspace{.2cm}+ \cdots \right)\nn\\
1564: &+&\frac{2}{h v_2}\left(|t_{23}|^2\hspace{.2cm}+\hspace{.2cm t_{24}}|^2\hspace{.2cm}+ \cdots \right)\nn\\
1565: &=& \left(\rho-\rho_0\right)_R\nn
1566: \eea
1567:
1568:
1569:
1570:
1571: \begin{thebibliography}{99}
1572: \bibitem{new82} R.G.Newton, Scattering theory of
1573: wave and particles (Springer -
1574: Verlag, New York, 1966), pg. 440.
1575: \bibitem{yac95} A.Yacoby et al.,
1576: Phys. Rev. Lett. {\bf 74}, 4047 (1995)
1577: \bibitem{sch97} R. Schuster et al.,
1578: Nature(London) {\bf 385}, 417 (1997)
1579: \bibitem{CER97} Cernicchiaro et. al, Phys. Rev. Lett. {\bf 79}, 273 (1997)
1580: \bibitem{deo96} P. Singha Deo,
1581: Phys. Rev.B {\bf 53}, 15447 (1996)
1582: \bibitem{comment} In some cases these NSDN can lead to a spontaneously
1583: broken discrete symmetry like parity. P.Singha Deo et al., (to be published)
1584: \bibitem{leg91} A.J. Leggett, in Granular Nano-Electronics, Vol.251 of NATO
1585: Advanced Studies Institute, Series B: Physics, edited
1586: by D.K. Ferry, J.R. Barker and C. Jacoboni (Plenum, New York,
1587: 1991), p.297
1588: \bibitem{fan} U.Fano
1589: Phys. Rev. {\bf 124}, 1866 (1961)
1590: \bibitem{bag90} P.F.Bagwell,
1591: Phys. Rev.B {\bf 41}, 10354(1990)
1592: \bibitem{tek93} E.Tekman and P.F.Bagwell
1593: Phys. Rev.B {\bf 48}, 2553 (1993);
1594: W. Porod, Z. Shao and C.S. Lent, Phys. Rev. B {\bf 48}, 8495 (1993).
1595: \bibitem{deo98} P. Singha Deo,
1596: Solid State Commun. {\bf 107}, 69 (1998)
1597: \bibitem{yey95} L.Yeyati and M.B{\"u}ttiker,
1598: Phys. Rev.B {\bf 52}, R14360 (1995)
1599: \bibitem{jay96} P. Singha Deo and A.M.Jayannavar
1600: Mod. Phys. Lett. B {\bf 10}, 787 (1996)
1601: \bibitem{cho98} C.-Mo Ryu and S.Y.Cho,
1602: Phys. Rev.B {\bf 58}, 3587 (1998)
1603: \bibitem{xu98} H.Xu and W.Sheng,
1604: Phys. Rev.B {\bf 57}, 11903 (1998)
1605: \bibitem{lee99} H.W.Lee,
1606: Phys. Rev. Lett. {\bf 82}, 2358 (1999)
1607: \bibitem{tan99} T.Taniguchi and M.Buttiker,
1608: Phys. Rev.B {\bf 60}, 13814 (1999)
1609: \bibitem{yeycm} A.L.Yeyati and M.B{\"u}ttiker, Phys. Rev. B
1610: {\bf 62}, 7307 (2000).
1611: \bibitem{fri52} J.Friedel, Philos. Mag. {\bf 43}, 153 (1952);
1612: J.M. Ziman, Principles of the Theory of Solids, 2nd ed.,
1613: Cambridge University Press, 1972.
1614: \bibitem{lan61} J.S.Langer and V.Ambegaokar, Phys. Rev. {\bf 121}, 1090
1615: (1961)
1616: \bibitem{smi59} F.T.Smith Phys. Rev. {\bf 113}, 349 (1960)
1617: \bibitem{kumar} R. Landauer and T. Martin, Rev. Mod. Phys. {\bf 66},
1618: 217 (1994).
1619: \bibitem{gri} S.Griffith
1620: Trans. Faraday Soc. {\bf 49}, 650 (1953)
1621: \bibitem{deo93} P.S.Deo and A.M.Jayannavar
1622: Mod. Phys. Lett. {\bf B 7}, 1045 (1993);
1623: %\bibitem{jay}
1624: P. Singha Deo and A. M. Jayannavar, cond-mat/0006035 and
1625: references therein.
1626: \bibitem{com} From the derivation of the FSR (one can see the book by Ziman
1627: in Ref. \cite{fri52}), that the scattering phase shift is defined such
1628: that when we set the scattering potential to 0 the scattering
1629: phase shift is also 0. However, in the double delta function potential
1630: the $arg(t)=kl$ when the strength of the delta function potentials
1631: are set to zero. Hence the scattering phase shift is $arg(t)-kl$.
1632: However, the term coming from the derivative of $kl$ cancels with the
1633: term coming from $\rho_0$, leading to the Eq. in (11).
1634: \bibitem{gas} V. Gasperin, T. Christen and M. B{\"u}ttiker, Phys.
1635: Rev. A {\bf 54}, 4022 (1996).
1636: \bibitem{gha} Quantum Mechanics by A.K. Ghatak and S. Loknathan,
1637: third edition, Macmillan India Ltd., pg. 57.
1638: \end{thebibliography}
1639: \end{multicols}
1640:
1641: \end{document}
1642:
1643:
1644:
1645: