cond-mat0703216/8.tex
1: \documentclass[aps,amsmath,prb,twocolumn,a4paper,showpacs,floatfix]{revtex4}
2: \usepackage{graphicx}
3: \newcommand{\comment}[1]{}
4: \newcommand{\beq}{\begin{equation}}
5: \newcommand{\eneq}{\end{equation}}
6: \newcommand{\bea}{\begin{eqnarray}}
7: \newcommand{\enea}{\end{eqnarray}}
8: \newcommand{\bc}{\begin{center}}
9: \newcommand{\ec}{\end{center}}
10: \newcommand{\vt}{ \varphi}
11: \newcommand{\ct}{ \cos  \Theta}\newcommand{\st}{ \sin  \Theta}
12: \newcommand{\cz}{ \cos  \chi}
13: \newcommand{\sz}{ \sin  \chi}
14: \newcommand{\cphi}{ \cos  \Phi}
15: \newcommand{\sphi}{ \sin  \Phi}
16: \newcommand{\nn}{\nonumber}
17: \newcommand{\wc}{\omega_c}
18: \newcommand{\met}{\frac{1}{2}}
19: \newcommand{\ddt}[1]{\frac{#1}{dt}}
20: \newcommand{\df}{\dot {\varphi}}
21: \newcommand{\dP}{\dot {\Phi}}
22: \newcommand{\dTh}{\dot {\Theta}}
23: \newcommand{\spin}[2]{|{#1}_1,{#2}_2 \rangle }
24: \newcommand{\mr}{\mathrm}
25: \newcommand{\cmt}{ \cos \frac{ \vartheta}{2}}
26: \newcommand{\smt}{ \sin \frac{ \vartheta}{2}}
27: \newcommand{\cmqt}{ \cos ^2\frac{ \vartheta}{2}}
28: \newcommand{\smqt}{ \sin ^2\frac{ \vartheta}{2}}
29: \newcommand{\bu}{\bar{u}}
30: \newcommand{\bv}{\bar{v}}
31: \newcommand{\UU}{\breve{{\cal{U}}}}
32: \newcommand{\GG}{\breve{G}}
33: \newcommand{\ovr}{\overline{R}}
34: \newcommand{\wo}{\omega_o}
35: \newcommand{\wu}{\omega_1}
36: \newcommand{\drond}[1]{\frac{\partial}{\partial #1} }
37: \newcommand{\EE}{\mathbf{E}}
38: \newcommand{\pp}{\mathbf{p}}
39: \newcommand{\sig}{\mathbf{\sigma}}
40: \newcommand{\rr}{\mathbf{r}}
41: \newcommand{\CRE}[1]{C^{\dagger}_{#1}}
42: \newcommand{\DES}[1]{C_{#1}}
43: \newcommand{\LAG}[3]{L_{#1}^{#2}\left(#3 \right)}
44: \newcommand{\COE}{\mathcal{C}}
45: \newcommand{\h}{\hbar}
46: \newcommand{\bd}{\begin{displaymath}}
47: \newcommand{\ed}{\end{displaymath}}
48: \newcommand{\de}{\delta}
49: \newcommand{\f}{\varphi}
50: \newcommand{\ww}{\omega}
51: \newcommand{\DD}{\Delta}
52: \newcommand{\la}{\vec \lambda}
53: \newcommand{\ug}{\underline{\hat{g}}}
54: \newcommand{\up}{\uparrow}
55: \newcommand{\da}{\downarrow}
56: \newcommand{\hd}{\hat d}
57: \newcommand{\hsp}[2]{\langle {#1}_1,{#2}_2 | }
58: \newcommand{\T}{\mathcal{T}}
59: \newcommand{\ang}{ i\omega _o t }
60: \newcommand{\bwt}{\begin{widetext}}
61: \newcommand{\ewt}{\end{widetext}}
62: \begin{document}
63: 
64: \title{Quantum Rings with Rashba spin orbit coupling: a path integral
65: approach}
66: 
67: \author{P. Lucignano$^{1,2}$}\author{D. Giuliano$^{1,3}$} 
68: \author{A. Tagliacozzo$^{1,2}$}
69: 
70: \affiliation{$^1$ Dipartimento di Scienze Fisiche Universit\`a degli
71:              studi di Napoli "Federico II", Napoli, Italy}
72: \affiliation{$^2$ Coherentia-INFM,
73:              Monte S.Angelo - via Cintia, I-80126 Napoli, Italy}
74: \affiliation{$^3$ Dipartimento di Fisica, Universit\`a della Calabria and
75:              I.N.F.N., Gruppo collegato di Cosenza, Arcavacata di Rende
76:              I-87036, Cosenza, Italy}
77: 
78: 
79: 
80: \pacs{03.65.Vf,% Phases: geometric, dynamic, topological
81:       72.10.-d,% Electronic transport and scattering mechanisms
82:       73.23.-b,% Electronic transport in mesoscopic system
83:       71.70.Ej % SO Coupling, Zeeman, Stark...           
84:               }
85: \date{\today}
86: \begin{abstract}
87: We employ a path integral real time approach to compute the DC
88: conductance and spin polarization for electrons transported across a
89: ballistic Quantum Ring with Rashba spin-orbit interaction.  We use a
90: piecewise semiclassical approximation for the particle orbital motion
91: and solve the spin dynamics exactly, by accounting for both Zeeman
92: coupling and spin-orbit interaction at the same time. Within our
93: approach, we are able to study how the interplay between Berry phase,
94: Ahronov Casher phase, Zeeman interaction and weak localization
95: corrections influences the quantum interference in the conductance
96: within a wide range of externally applied fields. Our results are
97: helpful in interpreting recent measurements on interferometric rings.
98: \end{abstract}
99: 
100: \maketitle
101: 
102: \section{Introduction}
103: 
104: In classical physics, a charged particle moving in an external
105: magnetic field $B$ feels a (Lorentz) force only in the regions in
106: which $B$ is different from zero. Instead, in 1959 Ahronov and Bohm
107: (AB)\cite{aharonovbohm} showed that the wave packet representing the
108: state of a quantum particle can be influenced by an external vector
109: potential $\vec{A}$, even if the corresponding magnetic field is zero,
110: provided that the particle is moving in a space with a nontrivial
111: topology (holes). To be more specific, the wavefunction of the charged
112: particle may acquire a nonzero phase, when undergoing a closed path in
113: a space threaded by an external magnetic flux.
114: 
115: Nowadays, mesoscopic quantum rings (QR's) allow to have direct access to
116: the phase of the electron wavefunction, since their size is smaller
117: than the distance over which the phase randomizes, as a consequence of
118: scattering against impurities and interactions. The total rate of
119: particles coming from the two arms of the ring and interfering at the
120: exit contact can be directly probed by measuring the QR conductance.
121: Indeed, interference effects have been observed in metal QRs many years
122: ago\cite{webb}.  In addition, since electrons are spinful particles,
123: the spin part of the wavefunction is influenced by the magnetic field
124: via the Zeeman term in the Hamiltonian.  Moreover, a more subtle
125: effect arises when there is a magnetic field non-orthogonal to the
126: plane of the orbiting particle because, as a consequence of the
127: orbital motion, its spin dynamics is instantaneously governed by a
128: time-dependent Hamiltonian \cite{anandans}.  This time dependence ends
129: up in an extra phase acquired by the particle wavefunction which is
130: named after Berry
131: \cite{berry,anandan}, who put in foreground its topological properties
132: when the orbits are closed.
133: 
134: Recently, semiconductor technology allows to grow samples (for
135: instance made by InAs or InGaAs) with sizeable spin orbit interaction
136: (SOI). Rashba\cite{rashba} first pointed out that the SOI strength can be
137: controlled by means of voltage gates.  This feature has been recently
138: found experimentally \cite{meijer,miller}. Because of the SOI, an electric 
139: field $E$ orthogonal to the orbit provides a momentum dependent effective 
140: magnetic field felt by the electron spin in orbital motion.  Such a
141: field can be equally well described by a vector potential that adds an
142: extra phase to the wavefunction\cite{aronov}. This phase was spelled
143: out by Ahronov and Casher \cite{ac}, as a ``dual'' AB effect, with
144: charge and spin interchanged (together with $B$ and $E$).
145: 
146: During the last years, the effects of SOI on the AB oscillations have
147: been observed in semiconductor based QR by several groups
148: \cite{yau,nitta,morpurgo,kato}.  It has been recognized that, in the
149: presence of both $B$ and $E$ orthogonal to the orbit plane, the
150: effective, momentum dependent  total $B$ field is tilted w.r. to the vertical
151: direction. The resulting Berry phase influences the interference
152: pattern. Such a result is of the utmost interest because the Rashba-SOI
153: (RSOI) turns out to be a tool to tune the Berry phase and, ultimately,
154: the conductance, as well as the spin polarization of the outgoing
155: electrons.  In Ref. \cite{molenkamp}, it is clearly shown that the AB
156: interference fringes can be modified by tuning an external
157: electrostatic potential.
158: 
159: In a recent publication\cite{noiletter} we have included all these
160: phenomena affecting the interference of an electron ballistically
161: transported in a ring, also accounting for some dephasing at the
162: contacts. We have also shown that the AB peak in the Fourier transform
163: of the magnetoconductance displays satellite peaks due to the RSOI. It
164: has been suggested that satellite peaks observed in recent
165: experiments\cite{shayegan} have the same origin. In the present work
166: we review the theory by presenting the full calculation. In addition,
167: we include the semiclassical paths leading to weak localization
168: corrections\cite{nonloso}, and we discuss the rotation of the spin
169: polarization, during the transport across the ring.
170: 
171: During the last years, several theoretical techniques have been
172: employed to study quantum transport in a QR. In Ref.s \cite{loss}, an
173: imaginary time path integral approach\cite{feynman} is developed to
174: study the conductance of a strictly one dimensional (1D) QR and its
175: conductance fluctuations in the diffusive limit. In
176: Ref. \cite{tserkov} a real time path integral approach is applied in
177: the limit of negligible Zeeman splitting.  Several papers have
178: discussed the conductance properties and the spin selective transport
179: of QR's in the strictly 1D ballistic limit, by means of a spin
180: dependent scattering matrix approach
181: \cite{molnar,frustaglia,shen,dario,citro}. In the absence
182: of the magnetic flux, the conductance shows quasi-periodic
183: oscillations in the SOI stength, which can be modified by switching
184: the magnetic field on.  Numerical calculations \cite{souma,lozano,wu}
185: have shown that in the 2D case there are only quantitative
186: modifications of the 1D results that do not qualitatively affect the
187: physics.
188: 
189: In this paper we extend the real time path integral approach
190: previously developed in Ref.\cite{noiletter} to study the conductance
191: and the spin transport properties of a ballistic quantum ring in the
192: presence of both RSOI and of an external magnetic flux orthogonal to
193: the ring plane. We use a ``piecewise'' saddle point approximation for
194: the orbital motion, keeping the full quantum dynamics of the
195: spin. This approach allows us to take into account, in a
196: nonperturbative way, both the RSOI and AB phase and to include also
197: the Zeeman spin splitting.
198: 
199: Our numerical approach evaluates all paths contributing to the quantum
200: propagator. The scattering at the leads can be forward or backward,
201: according to the probability amplitudes given by the S-matrix. Weak
202: localization corrections can be easily extracted from our result. We
203: also allow for some diffusiveness at the contacts by adding a random
204: phase factor in the motion.
205: 
206: The DC conductance is derived from the Landauer formula
207: \cite{landauer} $ {\cal{G}} ={e^2}/{\h} \sum_{\sigma\sigma'}
208: \left|A(\sigma; \sigma'|E)\right|^2$, where $A(\sigma ;\sigma'|E)$ is
209: the probability amplitude for an electron entering the ring with
210: energy $E$ and spin polarization $\sigma '$ to exit with spin
211: polarization $\sigma $. We also report the change in the spin
212: polarization the electron transported across the ring.
213: 
214: The structure of the paper is as follows:
215: \begin{itemize}
216: \item
217: In Section II we introduce the Feynman propagator for a spinful
218: electron injected at the Fermi energy in the ring.
219: \item 
220: In Section III we discuss the topology of the allowed paths and the
221: scattering of the electron at the leads.
222: \item
223: In Section IV we represent our path integral in the coherent spin
224: basis\cite{haldane,plet} and derive the saddle point equations of motion,
225: whose classical counterpart is described in detail in Appendix A. This
226: allows us to justify the choice of a piecewise semiclassical
227: approximation for the orbital motion of the electron in the ring.
228: \item
229: In Section V we present the details of the calculation by rewriting
230: the path integral as a collection of single arm propagators. These are
231: the building blocks to be calculated in the next section.
232: \item
233: In Section VI we analyze how the orbital motion affects the full
234: quantum dynamics of the electron spin for each arm of the ring and
235: chirality. The spin propagator is derived in Appendix B, in the basis
236: corresponding to the rotating reference frame in the spin space.
237: \item
238: In Section VII we discuss the dependence of the conductance on the external
239: fields and on the overall transmission across the ring.
240: \item
241: In Section VIII we focus on the spin polarization of the outcoming
242: electron.
243: \item
244: Section IX includes a short summary and our conclusions.
245: \end{itemize} 
246: %
247: \section{The transmission  amplitude}
248: 
249: Our model Hamiltonian will be the two-dimensional Hamiltonian for a
250: particle with spin-1/2 $\vec{S}$, in an orthogonal magnetic field,
251: with spin-orbit coupling to an orthogonal electric field (Rashba
252: coupling).  It is given by
253: %
254: \bea
255: H[ \vec{p} , \vec{r} , \vec{S} ]  =
256: \frac{1}{2m}\left(\vec p+\frac{e}{c}\vec A_0\right)^2
257: -  \omega_{c}\: S_{z}+{H}_{so}
258:  \label{hamilt} \\
259:  H_{so}= \frac{2\alpha}{\hbar ^2}\left( {\hat z} {\times} 
260: \left({\vec p+\frac{e}{c} \vec A_0}\right) \right){\cdot}
261: {\vec{S}}  \:\:\:\: , \nonumber
262: \enea
263: %
264: where $\alpha$ is the spin orbit coupling constant, in units $ eV
265: \:$\AA, $\vec{S} = \hbar \vec{\sigma } /2 $
266: (${\sigma_x,\sigma_y,\sigma_z}$ are the Pauli matrices),
267: $\vec{A}_0(\vec{r}) = \frac{B}{2} ( - y , x , 0 ) $ is the vector
268: potential generating the uniform field $B$, normal to the ring
269: surface, taken in the symmetric gauge, $\omega_c = ge B / 2mc$ is the
270: cyclotron frequency. In real nanostructures based on InAs or InGaAs
271: the $g$ factor can strongly deviate from the value of two.  However
272: the result we present here, are fully general as they depend on the
273: ratio $\alpha/\hbar \omega_c R$ which can be tuned by acting on
274: $\alpha$.  Since we will assume only a single channel to be avaliable
275: for electron propagation across the ring, we will picture the single
276: channel ring as a $1-d$ circle of radius $R$, connected to two
277: leads. Accordingly, the position of the particle within the ring is
278: parametrized by the angle $\vt$ and the vector potential has just the
279: azimutal component $A_\vt = \phi / 2\pi R $, where $\phi$ is the
280: magnetic flux threading the ring.
281: 
282: In order to study the conduction properties of the ring, one needs the
283: propagation amplitude for an electron entering the ring with spin
284: polarization $\mu_0$ to exit with spin polarization $\mu_f$, at energy
285: $E_0$. This is given by
286: 
287: \beq
288: A(\mu _f ; \mu _0 | E_0 )  =
289: \int_{0}^{\infty} \: \frac{d t_f}{\tau _0}  \:  e^{i \frac{E_{0}
290: t_f}{\hbar}}  \:
291: \langle {\vec r}_{f} , \mu_{f} , t_{f}|{\vec r}_{0} , \mu_{0 } , t_{0} \rangle
292:  \:\:\:\: ,
293: \label{ampclas1}
294: \eneq
295: %
296: 
297: where $\langle {\vec r}_{f} , \mu_{f} , t_{f}|{\vec r}_{0} , 
298: \mu_{0 } , t_{0} \rangle$ is the  amplitude for  a particle entering the
299: ring  at the point $\vec{r}_0$ and  at the time $t_0$ with spin 
300: polarization $\mu_0$ to exit at the point $\vec{r}_f$ at the
301: time $t_f$ with spin polarization $\mu_f$. In our tensor product notation,
302: we define $|{\vec r} , \mu \rangle=|{\vec r}\rangle\:\otimes\:|\mu\rangle\ $.
303:  $\tau _0 =mR^2 /(2\hbar) $ is the
304: time scale for the quantum motion.
305: 
306: In order to compute $\langle {\vec r}_{f} , \mu_{f} , t_{f}|{\vec
307: r}_{0} , \mu_{0 } , t_{0} \rangle$ , we resort to a path integral
308: representation for the orbital part of the amplitude. Since we
309: parametrize the orbital motion of the particle in terms of the angle
310: $\varphi$, we provide the pertinent Lagrangian, ${\mathcal{L}}_{orb
311: }$, as a function of $\varphi , \dot{\varphi}$.  It is given by 
312: %
313: \beq
314: {\mathcal{L}}_{orb}[\varphi(t), \dot \varphi(t),\vec \sigma]=
315: \frac{m}{2}R^2\dot\varphi^2(t)-\frac{\phi}{\phi_{0}} \hbar
316: \dot\varphi(t) + \frac{\alpha^2\:m}{2 \hbar^2}+ \frac{\hbar^2}{8 m
317: R^2} \:\:\:\: .
318: \label{lag}
319: \eneq 
320: %
321: The last two contributions to Eq.(\ref{lag}) are a constant, coming from
322: the spin-orbit term, and the Arthurs\cite{morette} term, which is
323: required when a path integration is performed in cylindrical
324: coordinates. Since both contributions are constant, they can be lumped
325: into the incoming energy $E_0$ and therefore they will be omitted
326: henceforth.
327: %
328: By taking into account the spin degree of freedom, as well, we represent
329: the propagation amplitude as
330: %
331: \bwt  
332: %
333: \beq 
334: \langle {\vec r}_{f}
335: , \mu_{f} , t_{f}|{\vec r}_{0} , \mu_{0 } , t_0 \rangle = \langle
336: {\vec r}_{f} , \mu_{f} | e^{ - i \int_{t_0}^{t_f } \: d t \: H } | {\vec
337: r}_{0} , \mu_{0 } , t_0 \rangle =\int_{\varphi ( t_0 ) =
338: \varphi_0}^{\varphi ( t_f ) = \varphi_f} \!\!\!\!{\cal D} \varphi\: e
339: ^{- i \int_{0}^{t_f} d t \; \left[ \tau_0 \dot{\varphi} ^2 - q
340: \dot{\varphi}\right ]} \: \langle \mu _f | \hat{U}_{spin} ( t_f, t_0 ) |
341: \mu _0 \rangle \:\:\:\: ,
342: \label{parte}
343: \eneq  
344: \ewt
345: %
346: where $q=\phi / \phi_o $,  $\phi_o $ being the
347: flux quantum $hc/e$. 
348: %
349: \beq
350: \hat{U}_{spin} ( t_f, t_0 ) = {\bf T} \exp \left [ -\frac{i}{\hbar} \:
351: \int _{t_0}^{t_f} dt' \: \hat{H} _{spin} (t') \right ] \:\:\: .
352: \eneq
353: %
354: is the full  spin  propagator and the spin Hamiltonian  
355: $ \hat{H} _{spin} (t)$ is given by 
356: %
357: \beq
358: \frac{1}{\hbar } \hat{H}_{spin} ( t ) =
359:  \left[ \begin{array}{cc}  \frac{  \omega _c}{ 2}  & 
360:  \gamma  \dot\varphi  e^{ - i \varphi ( t ) } \\
361: \gamma  \dot\varphi e^{  i \varphi (t)  } &
362: -  \frac{ \omega _c}{ 2}  \end{array} \right]
363: \:\:\:\: , 
364: \label{three}
365: \eneq
366: 
367: %
368: with $\gamma = 2\alpha \tau _0 /(\hbar R ) $\cite{nota}.
369: %
370: In Sections IV and V we show that the amplitude of Eq.(\ref{parte}) can
371: be approximated by choosing a piecewise semiclassical orbital motion
372: for the particle in each arm of the ring, while keeping the full
373: quantum dynamics of the spin.
374: %
375: In particular, we will see that, within the physically relevant range
376: of parameters, the orbital motion can be approximated as a uniform
377: rotation (with constant angular velocity), which makes the spin
378: dynamics to be the one of a spin-1/2 in an effective, rotating,
379: external magnetic field.  Yet, in order to explain how we deal with
380: quantum backscattering at the contacts between ring and arms and
381: corresponding dephasing effects, we will introduce our formalism in
382: the next section, by discussing a simplified version of our problem: a
383: spinless electron propagating across a mesoscopic ring.
384: %
385: \section{Feyman's paths for a Spinless particle transmitted across  a  ring}
386: %
387: In this section we introduce our formalism by computing the 
388: transmission amplitude for  a spinless electron of  mass $m$ and charge  $-e$, 
389: traveling   across the ring   in  an 
390: orthogonal magnetic field. 
391: %
392: For a realistic device, at each lead one has to take into account three 
393: possible scattering processes, consistently with the conservation of the
394: total current. This is  described in terms of a unitary $S-$matrix
395: that, in the  symmetric case in which  the two arms are symmetric, is
396: given by
397: %
398: \beq
399: \mathcal S=\left(
400: \begin{array} {c c c}
401:          -\met (1+\bar r) & \met (1-\bar r)          & \sqrt{\met(1-\bar r^2)}\\
402:           \met (1-\bar r) &-\met (1+\bar r)          & \sqrt{\met(1-\bar r^2)}\\
403:  \sqrt{\met(1-\bar r^2)}  & \sqrt{\met(1-\bar r^2)}  &  \bar r
404: \end{array}\right)
405: \label{matr}
406: \:\:\:\: .
407: \eneq
408: %
409: The numerical labeling of the S-matrix elements referring to the three
410: terminals of each contact fork, are explained in Fig.(\ref{noWLpaths},
411: 1a).  Assuming, for simplicity, that the scattering matrix is the same
412: for both leads, Eq.(\ref{matr}) will hold both at the left-hand lead,
413: and at the right-hand lead of the ring.
414: 
415: In particular, $\mathcal S_{3,3}=\bar r$ is the reflection amplitude
416: for a wave coming from the left lead, $\mathcal S_{1(2),1(2)}=-\met
417: (1+\bar r)$ the reflection amplitude for a wave incoming from the
418: upper/lower arm, $\mathcal S_{1(2),2(1)}= \met (1-\bar r)$ is the
419: transmission amplitude from the upper (lower) to the lower (upper) arm
420: and $\mathcal S_{1(2),3}=\mathcal S_{3,1(2)}=\sqrt{\met(1-\bar r^2)}$
421: is the transmission amplitude from the upper/lower arm to outside of
422: the ring.  
423: 
424: 
425: In Fig.(\ref{noWLpaths}) we show the simplest paths of the electrons
426: in the ring including forward scattering at the contacts, only.
427: The contribution to the total amplitude coming from such paths, in
428: which the electron enters the ring at an angle $\varphi_0$ at time
429: $t_0$ and exits at $\varphi_f$ at time $t_f$, is given by
430: %
431: %\bwt
432: \bea 
433: {\cal{A}}(\vt_{f}, t_f; \vt _0, t_0  ) =  \sum _{n=-\infty}^{+\infty} 
434: \int _{\vt (t_0)=
435:  \vt _i}^{\vt (t_f)= \vt _f +2\pi n }\!\!\!\!\!\!\! {\cal{D}}\vt (\tau ) \: 
436: e^{-i{\cal{S}}[\vt[t)]/\hbar }\nonumber\\
437:   =\sum _{n=-\infty}^{+\infty} 
438: e^{-iq \:(\vt_f-\vt_i +2\pi n ) }
439: \int _{\vt (t_0 )=
440:  \vt _i}^{\vt (t_f)= \vt _f +2\pi n }\!\!\!\!\!\!\! 
441: \!\!\!\!\!\!\! 
442: \!\!\!\!\!\!\! 
443: \!\!\!\!\!\!\! 
444: {\cal{D}}\vt (t ) \: 
445: e^{-i\frac{mR^2}{2\hbar }  \int _{t_0}^{t_f} dt \: \dot \vt ^2 (t )}  
446: \:,   
447: \label{spinless}
448: \enea
449: %\ewt
450: %%
451: where we have summed over paths in which the electron winds $n+1/2$ times in 
452: the ring, before exiting it. 
453: Positive (negative) $n$ values imply clockwise (counterclockwise)
454: propagation along the ring.  
455: 
456: This propagator can be evaluated exactly
457: \cite{moran}.  However we report here just the saddle point
458: evaluation, for comparison with the spinful case.  Minimizing the 
459: action gives the classical equation of motion (together with the pertinent
460: boundary conditions for a path winding  $n+1/2$ times): 
461: 
462: \beq 
463: \ddot \vt (t ) 
464: = 0 \:\:\:
465: ; \vt (t_i ) = \vt _i \: , \vt (t_f ) = \vt_f + 2\pi n 
466: \:\:\:\: . 
467: \eneq 
468: %
469: Let us assume that the particle is injected in the ring at $\vt_0 =0 $
470: and comes out at $\vt_f = \pi $ in a time $T = t_f$.
471: Eq.(\ref{spinless}) gives:
472: %
473: \bea {\cal{A}}(\pi ,0, t_f ) =
474: e^{i\pi q } \: \sqrt{\frac{\tau_0}{\pi i t _f}} \: {\sum
475: _{n=-\infty}^{+\infty }}' e^{-i \pi ^2 (2|n|-1) ^2 \tau_0  / t_f -i
476: 2\pi n q } \nonumber\\
477: \:\:\:\: , 
478: \label{less} 
479: \enea
480: %
481: where the prime in the sum takes into account the fact that one does not
482: sum over  $n=0$, and the square root at the prefactor accounts for the
483: gaussian fluctuations. Of course, this propagator is periodic in $q$ of
484: period $q=1$ up to a minus sign. 
485: 
486: More involuted paths arise if one takes into account backscattering
487: processes in which the electron can get backscattered within the same
488: ring's arm from which it is coming. For instance, the paths $(2f)$ and
489: $(2h)$, as well as $(2g)$ and $(2i)$ in Fig.(\ref{WLpaths}), include
490: looping in opposite directions around the ring's hole. Interference
491: between clockwise and counterclockwise windings leads to weak
492: localization corrections. We denote these corresponding paths
493: (including also $(2c)$ and $(2d)$ ) as ``reversed paths''$(RP)$. In our
494: approach, all order paths are numerically generated up to the
495: convergency and the $S-$matrix of Eq.(\ref{matr} ) is implemented in
496: the numerical algorithm.
497: %
498: \begin{figure}[!htp]
499:     \centering \includegraphics[width=\columnwidth]{paths2b.eps}
500: \caption{(color online) First and second order paths included in the
501: calculation of the transmission amplitude across the ring, 
502: from left to right, including
503: forward scattering only. Numbers $1,2,3(1',2',3')$ in Fig. 1a refer to
504: the labeling of the terminals in Eq.(\ref{matr}.)}
505: \label{noWLpaths}
506: \end{figure}
507: %
508: %
509: \begin{figure}[!htp]
510:     \centering
511:     \includegraphics[width=\columnwidth]{paths3.eps}
512: \caption{Second order paths of the transmission amplitude from left to
513:   right including backscattering at the leads. Paths $(2f)$ and
514:   $(2h)$, as well as $(2g)$ and $(2i)$ contribute to the weak
515:   localization corrections.}
516: \label{WLpaths}
517: \end{figure}
518: In conclusion, we have established that the paths contributing to the
519: transmission amplitude can be built up by adding four types of
520: elementary paths: $u_\rightarrow$ forward orbiting in the upper arm of
521: the ring ($\vt \in ( 0,\pi )$), $ u_\leftarrow$ backward orbiting in
522: the upper arm of the ring ($\vt \in ( \pi,0 )$), $d_\rightarrow$
523: forward orbiting in the lower arm of the ring ($\vt \in ( 2 \pi,\pi )$), $
524: d_\leftarrow$ backward orbiting in the lower arm of the ring ($\vt \in
525: ( \pi, 2 \pi )$). In Section V, we will generalize such an approach, while
526: in the next Section we discuss the Feynman path-integral
527: representation of the propagation amplitude for a spinful electron
528: propagating along one of these elementary paths.
529: %
530: \section{Quantum amplitude and semiclassical  orbital  motion 
531: for a spinful electron in the ring}
532: %
533: In this section, we construct the Feynman representation in the basis
534: of the coeherent spin states, for the propagation amplitude of an
535: electron moving along one of the arms of the ring with a given
536: chirality.  To discuss the saddle-point approximation we need the
537: equations of motion coming from the condition that the action is
538: stationary.  The coherent spin state basis provides a straightforward
539: route to perform a semiclassical approximation involving both orbital,
540: and spin degrees of freedom, at the same time. In particular, we will
541: show under which conditions the classical orbital motion $\dot{\vt}=
542: cnst $, can be retained, as in the spinless case.  Accordingly, the
543: spin dynamics will be that of a ``quantum magnetic moment'' in a
544: time-dependent external magnetic field.
545: 
546: Let $\Omega$ denote the orientation of the spin $\vec S$ and $|\Omega
547: \rangle$ be the coherent state such that:
548: %
549: \beq
550: \langle \Omega | \hat{\vec{S}}|  \Omega \rangle  \equiv \vec{S}[\Omega ] = 
551: \hbar \: S \:  \left[ \begin{array}{c} \sin  \Theta  \cos  \Phi  \\
552: \sin  \Theta  \sin  \Phi  \\ \cos  \Theta  \end{array} \right]
553: \:\:\:\: , 
554: \label{ringo2.2.2}
555: \eneq
556: \noindent
557: %
558: for the three components of the spin vector, respectively. 
559: %
560: The full propagator in the coherent spin state representation is given
561: by:
562: %
563: \bwt
564: %
565: \bea
566: \langle \varphi_f , \Omega _f , t_f \: |  \varphi_0 , \Omega _0 , 0
567: \rangle =\int_{\varphi ( 0 ) = \varphi_0}^{\varphi ( t_f ) = \varphi_f} 
568: \!\!\!\!{\cal D} \varphi\:   e ^{- i \int_{0}^{t_f} d t \;
569:  \left[  \tau_0  \dot{\varphi} ^2 -  q \dot{\varphi}\right ]}
570: \int_{\Theta ( 0 ) = \Theta_0}^{\Theta ( t_f )
571:  = \Theta_f} \!\!\!\!
572: {\cal D} \Theta \int_{\Phi ( 0 ) = \Phi_0}^{\Phi ( t_f ) = \Phi_f} 
573: \!\!\!\!{\cal D} \Phi\: 
574: \: e^{-\frac{i}{\hbar} 
575: {\cal{S}}_{spin}\left [ \Theta , \dot{\Theta} ; \Phi , 
576: \dot{\Phi} | \vt , \dot{\vt} \right ] }
577: \:\:\:\: , 
578: \label{propo}
579: \enea
580: %
581: where:
582: %
583: \bea
584: {\cal{S}}_{spin} \left [ \Theta , \dot{\Theta} ; \Phi , 
585: \dot{\Phi} | \vt , \dot{\vt} \right ]  /\hbar  = \int_{0}^{t_f} d t \;
586:  \left \{ \frac{ (1-\cos\Theta )}{2} \dot \Phi + {\cal {L}}_{spin} 
587:  \left [ \Theta , \dot{\Theta} ; \Phi , 
588: \dot{\Phi}| \varphi , \dot{\varphi}   \right ] \right \} 
589: \:\:\:\: . 
590: \label{prospin}
591: \enea
592: 
593: %
594: with:
595: %
596: \bea
597:   {\cal L}_{spin}  \left [ \Theta , \dot{\Theta} ; \Phi , 
598: \dot{\Phi} | \varphi , \dot{\varphi} \right  ] =
599:   \frac{ \omega_c}{2} \cos \Theta - \gamma
600: \dot{\varphi} \sin \Theta \cos ( \varphi - \Phi  ) 
601: \:\:\:\: . 
602: \label{venti}
603: \enea
604: 
605: %
606: \ewt
607: %
608: The Lagrangian of Eq.(\ref{venti}) corresponds to the coherent
609: spin-state representation of the classical Lagrangian derived from
610: Eq.(\ref{hamilt})
611: %
612: \bea
613:  {\cal L} [ \varphi , \dot{\varphi} , \vec{S}  ] =
614:  \nonumber \\
615:  \frac{m}{2}   \dot\vec{r} ^2  - 
616: \frac{e}{c}\:   \dot\vec{r} \cdot \vec{A}   + 2m \alpha  \left [
617:  \hat z\cdot  \left ( \dot\vec{r}  \times \vec{S}\right ) \right ] +
618:  \frac{ m \alpha^2}{\hbar^2} +  \omega_c S_z \: .  
619:  \label{sette} 
620: \enea
621: %
622: In Eq.(\ref{sette}) we have introduced the constraint that the orbital electron
623: motion takes place along a one-dimensional circle by parametrizing the
624: trajectories with the angle $\varphi$ as $ x = R \cos \varphi ; \; y =
625: R \sin \varphi $.
626: %
627: The additional {\sl Berry phase } term \cite{aurbach} in
628: Eq.(\ref{prospin}) arises from the fact that different spin coherent
629: states are not orthogonal to each other since, to leading order in
630: $\epsilon$, the scalar product between spin-coherent states at times
631: $t_j , t_j +\epsilon $, $ | \Omega ( t_j ) \rangle$ and $ | \Omega (
632: t_{ j } + \epsilon ) \rangle$ is given by
633: %
634: \beq
635: \langle \Omega ( t_j + \epsilon ) | \Omega ( t_j ) \rangle \approx 
636: \exp \left[ \frac{i}{2} [ 1 - \cos \Theta ( t_j ) ] \dot{\Phi} ( t_j ) 
637: \epsilon \right]
638: \:\:\:\: . 
639: \label{ringostar}
640: \eneq
641: 
642: %
643: Let us look for the trajectories in orbital and spin space which make
644: the action stationary.  This requires solving the Eulero-Lagrange
645: equations for the Lagrangian $ {\cal L} [ \varphi , \dot{\varphi} ;
646: \Theta , \dot{\Theta} ; \Phi , \dot{\Phi} ]$ appearing in Eq.(\ref{propo}),
647: which are given by
648: %
649: \begin{eqnarray} 
650: \frac{\dot{\Theta}}{2} \sin \Theta  + \gamma 
651: \dot{\varphi} \: \sin  \chi \:  \sin  \Theta  &=& 0 
652: \label{em1}\:, \\
653: \sin  \Theta  [- \dot{\Phi} + \omega_c ] +
654: 2  \gamma  
655: \dot{\varphi} \: \cos \Theta\: \cos \chi  &=& 0 
656:  \label{em2}\:, \\ 
657: \ddot{ \varphi} - \frac{ \gamma}{2\tau  _0} [ \dot{\Theta} \: \cos \Theta
658: \: \cos \chi + \dot{\Phi} \: \sin \Theta\:  \sin \chi  ]
659: &=& 0 \label{em3}
660: \: ,
661: \end{eqnarray}
662: 
663: %
664: with $\chi = \varphi -\Phi $. 
665: 
666: In order to extract the relevant physics from the above equations, we
667: multiply Eq.(\ref{em3}) by $\dot{\varphi}$, and, by use of
668: Eqs.(\ref{em1},\ref{em2}), we rewrite Eq.(\ref{em3}) as:
669: %
670: \beq \dot{\varphi} \ddot{
671: \varphi} =- \frac{ \omega _c}{4\tau _0} \: \dot{\Theta} \: \sin\Theta
672: \:\:\:\: .
673:  \label{orbo} 
674: \eneq
675: %
676: A straightforward time integration gives:
677: %
678: \beq
679: \tau_0 \frac{ ( \dot{\varphi} )^2 }{2} - \frac{\omega_c}{4} \cos \Theta = cnst
680: \:\:\:\: , 
681: \label{orbo2}
682: \eneq
683: 
684: which states that the total energy is conserved. According to
685: Eq.(\ref{orbo2}), the particle energy only includes the orbital kinetic
686: term and the Zeeman term. This is obvious, since the force associated
687: to the spin orbit coupling, being gyroscopic, does no work.
688: %
689: A change in the precession angle $\Theta$ implies acceleration in the
690: orbital motion.  According to the spin Hamiltonian of Eq.(\ref{three})
691: the RSOI is responsible for flipping of the spin, while the Zeeman
692: coupling tends to stabilize the spin direction.  Hence, there are two
693: physically relevant limits, in which the orbital motion fully
694: decouples from the spin dynamics, according to the inequality
695: $\omega_c/2 <(>) \dot \varphi$, as we are going to show next.
696: Both limits involve  an orbital  motion with constant velocity 
697: $\dot \varphi $ and a spin orientation  given by the angles 
698: $\Theta = {\rm constant} \:\:\: , \:\:
699: \varphi - \Phi = 0 \: (mod.\: \pi ) $.
700: 
701: {\sl a) Vanishing Zeeman coupling}:  This case has already been considered 
702:  and it has been shown that it can be exactly solved
703: analytically\cite{tserkov}. In this case Eq.(\ref{orbo})
704: allows for the semiclassical solution $\dot\varphi =cnst $ and quantum
705: fluctuations of the spin do not interfere with the orbital motion.
706:  The spin is tilted by the constant angle $\tan \Theta = 2\gamma $ and 
707: precesses with a constant frequency $\dot \Phi = \dot \varphi $. 
708: 
709: {\sl b) Large Zeeman coupling}: flipping of the spin and quantum
710: fluctuations of the spin are strongly disfavoured, again corresponding 
711: to the classical
712: saddle point configuration $\Theta = {\rm constant} \:\:\: , \:\:
713: \varphi - \Phi = 0, \pi  $  with 
714: %
715: \beq
716: \pi/t_f=\dot{\varphi} = \pm \dot{\Phi} = 
717:  \omega_c \:\frac{1}{1-2\gamma \:  {\rm ctan} \Theta}\:\:\:.
718: \label{orbo3}
719: \eneq
720: %
721: The spin precesses around the $z$-axis, with the same angular velocity
722: as the orbital motion. When $\gamma \to 0 $, $\Theta $  can  be
723: vanishingly small, with $\dot\Phi =\omega _c  $.
724: 
725: It may appear that the simple spin precession provided by
726: Eq.(\ref{orbo}) with $\Theta = cnst $ is a solution of
727: Eq.s(\ref{em1},\ref{em2},\ref{em3}) for any ratio $\omega _c
728: /\dot\varphi $. Were this the case, there would be no need to invoke
729: the limitations of case $b)$ as stated above.  However, a careful
730: analysis of the stability of this saddle point solution shows that the
731: simple spin precession is a minimum of the action only when the Zeeman
732: coupling is strong. In particular, if $\omega _c /\dot\varphi >>1 $,
733: an additional condition $\omega _c \tau _o/\gamma ^2 >1 $ has to be
734: satisfied. Otherwise the frequency of the fluctuations around the
735: saddle point solution does not keep real and the analysis of small
736: oscillations in the parameter space around the saddle point solution
737: breaks down.
738: 
739: This shows that a classical orbital motion with constant velocity is
740: compatible with both limits of large and small ratios of the Zeeman
741: coupling to the RSOI strength. However the two limiting cases do not
742: belong to the same saddle point. In particular there will be a
743: crossover region connecting the two limits in which quantum
744: fluctuations of the spin may induce changes in the orbital velocity of
745: the particle.
746: 
747: In the rest of the paper we will choose $\dot \varphi=const$ piecewise
748: and parametrize the quantum dynamics of the spin with the value of the
749: velocity obtained by the stationary phase condition discussed in the
750: next section.  As discussed above, our approximation cannot reproduce
751: faithfully the intermediate region of parameters ranging the two
752: limits of large and small Zeeman coupling $\omega_c / \gamma
753: \dot\varphi <(>) 1$.  We checked numerically the reliability of our
754: approximation, by numerically integrating
755: Eqs.(\ref{em1},\ref{em2},\ref{em3}) and have have found that it is
756: satisfactory in most of the parameter range.
757: 
758: By putting $\dot\varphi =cnst $ into Eqs.(\ref{em1},\ref{em2}), they
759: become, as we show in detail in Appendix A, the classical equations of
760: motion for a magnetic moment in a time dependent magnetic field
761: ${\cal{B}}
762: \equiv ({\cal{B}}_+,{\cal{B}}_-,{\cal{B}}_z ) = \left ( \gamma
763: \dot{\vt} \: e^{i\vt}, \gamma \dot{\vt} \: e^{-i\vt},-\omega _c/2
764: \right ) $. This can be easily understood from the fact that
765: minimizing the action in Eq.(\ref{propo}) directly w.r. to the spin
766: components provides:
767: %
768: \beq 
769: \vec{S}\times \frac{d\vec{S}}{dt} = - \frac{\delta H }{\delta
770: \vec{S}} \:\:\:\: , 
771: \eneq 
772: %
773: which has to  be solved together with the constraint of constant 
774: modulus: $\vec{S}\cdot d\vec{S} /dt = 0$ ( see Eq.(\ref{appe1}).
775: %
776: 
777: Among the other possible saddle-points, a solution of the motion
778: equations can be found with the particle trapped within the ring arm
779: (turning points of $\varphi(t)$ are at $\varphi = 0,\pi $) if $\gamma
780: $ is large enough. This solution doesn't seem to be practical as it
781: requires fine tuning of the external parameters with transfer of
782: energy from the spin motion to the orbital motion and viceversa. We
783: have not investigated it in detail, but we expect that it could
784: provide resonant tunneling across the ring.
785: %
786: \section{Saddle point approximation and looping in the ring}
787: %
788: In this Section we show how we implement the piecewise saddle point
789: solution for the orbital motion, $\dot{\vt} = cnst$, to study the
790: coherent propagation of the electron inside the ring.
791: %
792: In the following, we will denote by ``loop'' and ``looping
793: trajectory`` both trajectories that wind around the ring (closed), and
794: paths in which the particle moves forth and back in one of the ring
795: arms (open) (see Fig.(\ref{WLpaths})).  Of course, the amplitudes
796: differs very much in these two types of looping . Indeed, the net spin
797: rotation is different between the two paths and, also, the
798: Ahronov-Bohm ($AB$) phase is absent in the latter ones .
799: %
800: 
801: %
802: \begin{figure}[!htp]
803:     \centering \includegraphics[width=\columnwidth]{tree.eps}
804: \caption{(color online) Cayley tree describing the way in which higher
805:   order paths are built in the numerical code. Full lines correspond to
806:   propagation in the upper arm ( $u$ ) or lower arm ( $d$) in the
807:   forward ($\rightarrow$) or backward ( $\leftarrow$) direction. The
808:   nodes represent the leads of the ring. The exit nodes are marked by
809:   a broken line at each order $n$.  The heavy line correspond to the
810:   path reported in Eq.(\ref{ugly}). }
811: \label{tree}
812: \end{figure}
813: %
814: In general, at order $n$, we will have $2^{(2|n|-1)}$ trajectories of
815: a particle entering the ring at $\vt = 0 $ and exiting at $\vt = \pi $
816: and each of them will include $2|n|-1 (n = \pm 1,\pm 2,...) $
817: elementary paths (or ``stretches'') of the type
818: $u_\rightarrow,u_\leftarrow,d_\rightarrow,d_\leftarrow$, as classified
819: at the end of Section II.  Looping at order $|n|$ involves $2 |n|$
820: scattering processes at the contacts.  In the case of spinful
821: electrons the $S$-matrix is doubled ($6\times 6$) w.r. to the one
822: given in Eq.\ref{matr}.  Of course, the $S$-matrix at the contacts is
823: sample dependent and any special choice is arbitrary.  In the
824: following we neglect possible asymmetries in the up-down channel as
825: well as accidental spin flipping in traversing the contacts. Hence, we
826: make the simplifying assumption that the $S$-matrix is block diagonal
827: of the form given in Eq.\ref{matr} for both spins.
828: 
829: Each time the trajectory impinges at a contact without leaving the
830: ring, the $S-$ matrix of Eq.(\ref{matr}) alledges for either forward,
831: or backward scattering.  Let us denote with
832: $u_{\rightarrow(\leftarrow)}(t_i,t_j)$ the forward (backward)
833: propagation amplitude in the upper arm from time $t_j$ to time $t_{i}$
834: ($d_{\rightarrow(\leftarrow)}(t_i,t_j)$ for the lower arm).  The
835: $u(d)$ amplitudes include the Ahronov-Bohm phase and the spin
836: evolution, but not the dynamical phase, which is factored out (see
837: below).  As shown in section II, to first order ($|n|=1$) there are
838: only two possible paths (see Fig.(\ref{noWLpaths} 1a,1b)), the
839: propagation amplitude is, then, the sum of the the corresponding two
840: amplitudes:
841: %
842: \bwt
843: %
844: \beq
845: A_1(\mu _f ; \mu _0 | E_0 ) = \int_0^\infty \: \frac{d t_f}{ \tau_0} 
846: e^{ i \frac{E_0 t_f}{\hbar}}  
847: \left\langle\mu_f,t_f\right|\biggl [ ( \mathcal S_{3'1} u_\rightarrow(t_f,t_0) \mathcal S_{13} +  \mathcal
848: S_{3'2} d_\rightarrow(t_f,t_0) \mathcal S_{23} ) \biggr ] 
849: \left|\mu_0,t_0\right\rangle \: 
850: e^{-i\frac{mR^2}{2\hbar }  \int _{t_0}^{t_f} dt \: \dot \vt ^2 (t )}
851: \label{prio}
852: \eneq
853: %
854: The amplitudes have to be summed all together, order by order.  The
855: key observation appearing in the symbolical writing of Eq.(\ref{ugly})
856: is that the dynamical phase, at a given order, does not depend on the
857: chirality of the motion and can be factored out.  On the contrary, the
858: Ahronov-Bohm phase depends on the chirality, and the spin evolution
859: depends on both the chirality and on the modulus of the propagation
860: velocity.
861: %
862: Beyond the first order, there is a net increase in the number and type
863: of the paths to be summed together. In Fig.(\ref{tree}), we
864: pictorially sketch all the possible paths by mean of a Cayley tree.
865: Each node represents a lead of the ring and, according to the
866: scattering matrix of Eq.(\ref{matr}), the electron can be
867: backreflected into the ring's arm it is coming from with an amplitude
868: $S_{ii}$, or, with an amplitude $S_{ij}$, it can be either transmitted
869: to the other arm, or outside of the ring.  In the tree, the
870: transmission out of the ring is possible at all the nodes crossing the
871: dashed lines. Each dashed line is labeled by the order $n$ of the
872: interference in the ring. As an example, we explicitely write down one
873: of the possible second order paths (the bold red line marked by $\cal
874: {P}$ in Fig.(\ref{tree}), which corresponds to
875: Fig.(\ref{WLpaths},2h)):
876: %
877: \bea
878:  A_2^{\cal{P}} (\mu _f ; \mu _0 | E_0 )=\!
879: \int_0^\infty \!\! \frac{d t_f}{ \tau_0} \! e^{ i \frac{E_0 t_f}{\hbar}}  
880: \int _{t_0}^{t_f} \!\!\!dt_2 \! 
881: \int _{t_0}^{t_2}\!\! dt_1 \!    
882: \left\langle\mu_f,t_f\right|  
883: \mathcal S_{3'2} d_\rightarrow(t_f,t_2) \mathcal S_{22} d_\leftarrow(t_2,t_1) \mathcal
884: S_{21} u_\rightarrow (t_1,t_0)\mathcal S _{13}
885: \left|\mu_0,t_0\right\rangle
886: e^{-i\frac{mR^2}{2\hbar }  \int _{t_0}^{t_f} dt  \dot \vt ^2 (t )} \:.
887: \label{ugly}
888: \enea
889: %
890: 
891: 
892: At the saddle point with uniform velocity, $\dot{\vt} = 2 \pi \:
893: (2|n|-1) / t_f $, we classify the collection of sequences belonging to
894: a certain $n$ as $\{{\cal{C}}_n\} $, and denote the superposition of
895: amplitudes (e.g., the ones in the big parenthesis of Eq.(\ref{prio})
896: to first order ) as $ {\cal{F}}
897: \left [\mu _f, \mu _0; q, \dot{\vt} | {\cal{C}}_n \right ] $.  In this
898: way, we can rewrite the full amplitude of Eq.(\ref{ampclas1}) as:
899: %\bwt
900: \beq
901: \mathcal A(\mu _f ; \mu _0 | E_0 )  = 
902: \int_{0}^{\infty} \: d t_f  \:  e^{-i E_{0}
903: t_f/\hbar}
904: \sqrt{\frac{\tau _0}{\pi i t_f } } \:
905:  \: {\sum _{n=-\infty}^{+\infty }}' \:  \sum _{\{{\cal{C}}_n\}}
906:  \: {\cal{F}} \left [ \mu _f, \mu _0; q,  \dot{\vt} = 2 \pi \: (2 |n|-1) / t_f
907:  |  {\cal{C}}_n \right ]   \:
908: e^{  -i  \pi ^2 (2|n| -1) ^2 \tau _0 / t_f  }
909:  \: .
910: \label{coll}
911: \eneq 
912: \ewt
913: %
914:  The series in Eq.(\ref{coll}) is  uniformly convergent. Thus, we may
915: swap the integral with the sum, and integrate term by term.  
916: The integral contributing to order $n$ is given by
917: %
918: \bea I_n = \int_{0}^{\infty} \: d t_f \:
919: \sqrt{\frac{\tau _0}{\pi i t_f }} \: e^{-i E_{0} t_f/\hbar -i \pi ^2
920: (2|n| -1) ^2 \tau _0 /t_f } \: \nonumber\\ =\: \sqrt{\frac{1}{\pi i }}
921: \: \int_{0}^{\infty} \: d x \: e^{-i\epsilon \: x^2 -i \pi ^2 (2|n|
922: -1) ^2 /x^2 }
923: \label{stat}
924: \enea 
925: %
926: with $\epsilon = E_{0} \tau _0/\hbar $.  We compute it approximately
927: within stationary phase contribution. Since the phase of the exponent
928: of the integrand is stationary at $\bar{t}_n = \epsilon ^{(-1/2)} \:
929: \pi \:(2|n|-1) \tau _0$, by inserting this value in the phase and
930: integrating out the gaussian fluctuations, we readily get:
931: %
932: \bea
933: I_n \approx  e^{-i \sqrt{\epsilon}  2 \pi (2|n|-1) } 
934: \sqrt{\frac{1}{\pi i  }} \:
935:   \int_{-\infty}^{\infty} \: d (\delta x ) \:
936:   e^{-i \epsilon \:( \delta x )^2} \nonumber\\
937:   \sim 
938: \frac{1}{i \sqrt{2\epsilon  }}
939:  \: e^{-i \sqrt{\epsilon}  2 \pi  (2|n|-1)^2  } \:\:\: .
940: \enea 
941: %
942: ($ -i \epsilon ^{-1/2} $ is the usual factor appearing in the
943: one-dimensional free particle Green's function in real space and
944: energy).  This approximation, when plugged into Eq.(\ref{coll}),
945: provides the final result:
946: %
947: \bwt
948: \beq
949: A(\mu _f ; \mu _0 | E_0 )  = 
950:  \frac{1}{i \sqrt{ 2\epsilon } } \: 
951:  {\sum _{n=-\infty}^{+\infty }}' \: \biggl (  \sum _{\{{\cal{C}}_n\}}
952:  \: {\cal{F}} \biggl [ \mu _f, \bar{t}_n , \mu _0; q,  \dot{\vt} =
953:  2 \pi \: ( 2|n|-1 ) / \bar{t}_n
954:  \biggl | \biggr .   {\cal{C}}_n \biggr ]  \: \biggr )  \: 
955:     \: e^{-i \sqrt{\epsilon}  2 \pi  (2|n|-1)  }
956:  \: .
957: \label{colf}
958: \eneq 
959: \ewt
960: %
961: The enumeration of the trajectory configurations belonging to the
962: collection $ {\cal{C}}_n $, to order $n$, is numerically performed
963: order by order.
964: 
965: In the next Section, we discuss the elementary
966: propagators for each of the four stretches,
967: $u_{\rightarrow},u_{\leftarrow},d_{\rightarrow},d_{\leftarrow} $, as
968: defined in Section III.  This allows us to construct the functional $
969: {\cal{F}} $ for each incoming and outgoing spin polarization.
970: 
971: 
972: \section{Quantum spin  dynamics of the electron  propagating in the ring} 
973: In this Section, we provide the explicit formula for the spin
974: contribution to the total propagation amplitude, given by
975: \beq
976: \hat{U}_{\rm spin} ( t , t' ) = {\bf T} \exp \left[ - \frac{i}{ \hbar}
977: \int_{ t' }^{ t } \: d \tau \: H_{\rm spin} ( \tau ) \right] 
978: \:\:\:\: .
979: \label{a.2.1}
980: \eneq
981: 
982: As discussed in detail in appendix A, within the saddle point
983: approximation, $ H_{\rm spin} ( t )$ corresponds to the Hamiltonian of
984: a spin-1/2 in a time-dependent external magnetic field. It can be
985: written as (from now on, we will denote by $\omega_o$ the frequency of
986: the orbital motion, that is, the stationary phase value of
987: $\dot{\varphi}$)
988: 
989: \beq
990: \hat{H}_{\rm spin} ( t ) = \left[ \begin{array}{cc} r \cos  \vartheta  & 
991: r \sin  \vartheta  e^{ - i \omega_o t } \\
992: r \sin  \vartheta  e^{  i \omega_o t  } &
993: - r \cos  \vartheta  \end{array} \right]
994: \label{three_1}
995: \:\:\:\: , 
996: \eneq
997: 
998: with
999: \begin{eqnarray}
1000:  r \cos \vartheta =   \frac{ \omega _c}{ 2}\: ,  \:\:\:\:\:
1001: r \sin \vartheta = \gamma \omega_o \: ,   \:\:\:\:\:
1002: \varphi (t) =  \omega_o  t \: ,    \label{def}\\ 
1003: r =  \met \sqrt{ \omega _c^2+4 \gamma ^2  \omega_o ^2} \: ,
1004:   \:\:\:\:\: \tan \vartheta  =   \frac{2\gamma  \omega_o }{ \omega _c}
1005: \:\: .
1006:  \nonumber
1007: \end{eqnarray}
1008: 
1009: Including only the AB phase implies $\vartheta = 0 $, while including
1010: only RSOI implies $\vartheta \to \pi /2 $.
1011: 
1012: It is useful to solve for the spin dynamics in the representation of
1013: the instantaneous eigenstates of $\hat{H}_{\rm spin} (t)$. At fixed
1014: $t$, its eigenvalues are given by $\pm \epsilon = \pm r$, while the
1015: corresponding eigenvectors take the form:
1016: \begin{eqnarray}
1017: | + , t \rangle = \left ( \begin{array}{c} 
1018:     \cmt  \\  
1019:  \smt \: e^{\ang }     \end{array} \right ) \: ,  \:\:\:\:\:\:
1020: | - , t \rangle = \left ( \begin{array}{c} 
1021:     -\smt \: e^{ - \ang}  \\  
1022:  \cmt   \end{array} \right ) \: . 
1023: \label{four}
1024: \end{eqnarray}
1025: 
1026: Thus, the matrix diagonalizing $\hat H_{\rm spin} ( t ) $ at time $t$
1027: is
1028: \begin{eqnarray}
1029: \hat B ( t ) \equiv \left[ \begin{array}{cc} 
1030:  \cmt &   - \smt \: e^{ - \ang} \\
1031:  \smt \: e^{  \ang}  &  \cmt
1032:  \end{array} \right] 
1033: \:\:\:\: .
1034: \label{five}
1035: \end{eqnarray}
1036: The matrix $\hat B ( t )$ encodes the adiabatic contribution to the
1037: evolution of $ | \Psi ( t ) \rangle$. Therefore, in order to write
1038: down the Schr\"odinger equation with Hamiltonian $\hat{H}$,
1039: \[
1040: \biggl\{ i \frac{ \partial}{ \partial t} - \hat{H} ( t ) \biggr\} 
1041: | \Psi ( t ) \rangle = 0 
1042: \]
1043: 
1044: in the adiabatic basis, one has to strip off from the state $ | \Psi (
1045: t ) \rangle$ its adiabatic evolution, operating with $\hat{B}^\dagger
1046: (t)$, so to get:
1047: \beq
1048:  \biggl[ i \frac{ \partial}{ \partial t} 
1049: -   \hat{B}^\dagger ( t ) \hat{H} ( t ) \hat{B} ( t ) +  
1050:  \hat{B} ( t )^\dagger  i \frac{ \partial}{ \partial t} \hat{B} ( t ) 
1051: \biggr]  \hat{B}^\dagger ( t ) | \Psi ( t ) \rangle  = 0 \:\: .
1052: \label{addi2}
1053: \eneq
1054: 
1055: Eq.(\ref{addi2}) may be rewritten in a 2$\times$2 matrix formalism.
1056: Let $ \left( \begin{array}{c} u_+ \\ u_- \end{array} \right )$ be the
1057: components of $ | \Psi ( t ) \rangle$ in the adiabatic basis. The
1058: corresponding system of differential equations reads
1059: 
1060: \beq
1061: i \frac{d}{dt}\left ( \begin{array}{c}
1062:     u_+  \\
1063:     u_-     \end{array} \right ) \:=
1064: \hat{H} _A(t) \: 
1065: \left ( \begin{array}{c}
1066:     u_+  \\
1067:     u_-
1068: \end{array} \right )\label{addi6}
1069: \eneq
1070: where we have defined
1071:  
1072: \bea
1073:  \hat{H}_A= 
1074: \left( \begin{array}{cc} 
1075: r+ \wo \smqt & \met  \wo \sin\vartheta   e^{-i \wo t} \\
1076: \met \wo \sin\vartheta  e^{i \wo t} & - r - \wo \smqt \end{array}
1077: \right)\:.
1078: \label{berryham}
1079: \enea
1080: The  extra  term  appearing on the diagonal w.r.to the hamiltonian of 
1081: Eq.(\ref{three_1}) is just the Berry  phase:
1082: \beq
1083:  \langle + , t | i \frac{d}{ d t } | + , t \rangle = -
1084: \langle - , t | i \frac{d}{d t } | - , t \rangle =  \omega _o \: \smqt \:\: . 
1085: \eneq
1086: %
1087: Eq.(\ref{addi6}) is solved in Appendix B  and   the full propagator 
1088: in the representation  of the instantaneous eigenvectors  reads: 
1089: \bwt 
1090: 
1091: \beq
1092: U(t,t')\:=\:
1093: \left ( \begin{array}{cc}
1094: (\cos(\epsilon (t-t'))- i \eta \sin(\epsilon (t-t')))e^{i/2 \varphi
1095:  (t-t')}  &
1096: -i \beta \sin (\epsilon (t-t'))e^{i/2 \varphi(t+t')}   \\
1097:  -i \beta \sin (\epsilon (t-t'))e^{-i /2  \varphi (t+t')}    &
1098: (\cos(\epsilon (t-t'))+ i \eta \sin(\epsilon (t-t'))) e^{-i/2  \varphi
1099:  (t-t')}
1100:  \end{array} \right )\;.
1101: \label{propagatorespin}
1102: \eneq
1103: %
1104: \ewt
1105: where  $ \epsilon=\pm\sqrt{ (r + \frac{\wo}{2}\cos
1106: \vartheta )^2+s^2}$  and $s = \frac{\omega _o}{2} \sin \vartheta $. Also:
1107: %
1108: \[\beta=\frac{\wo}{2\epsilon } \sin\vartheta \:\:\: , 
1109: \:\:\:\eta =\frac{ r + \frac{\wo}{2}\cos
1110:   \vartheta  }{ \epsilon}\: . \]
1111: %
1112: This is the propagator in the adiabatic basis.  Therefore, in order to
1113: switch to the fixed  spin basis, one should write
1114: $U_{spin}(t,t')=B(t)U(t,t')B^\dagger(t') $, where $B(t) $ is given by
1115: Eq.(\ref{five}).
1116: The four elementary  stretches imply  the following substitutions in 
1117:  the propagator  of Eq.(\ref{propagatorespin}):
1118:   
1119:  $u_\rightarrow$)  {\sl   forward orbiting in the upper arm of
1120: the ring :}   $\:\:\:\:\:\:\varphi  (t) =   \omega _o t $ and  $\vartheta \to  \vartheta $.
1121: 
1122:  $ u_\leftarrow$)  {\sl   backward orbiting in
1123: the upper arm of the ring :}    $\:\:\:\:\:\:\varphi  (t) =  \pi  - \omega _o t $ and 
1124:  $\vartheta \to  -\vartheta  $.
1125: 
1126:  $d_\rightarrow$)  {\sl   forward orbiting in the lower arm of the ring :} 
1127:      $\:\:\:\:\:\:\varphi  (t) =  2\pi  - \omega _o t $ and  $\vartheta \to  -\vartheta  $.
1128: 
1129:  $ d_\leftarrow$)  {\sl  backward orbiting in the lower arm of the ring :}
1130:     $\:\:\:\:\:\:\varphi  (t) =  \pi  + \omega _o t $ and  $\vartheta \to  \vartheta  $.
1131:  
1132: \section{the  conductance}
1133: %
1134: In this section, we derive the DC conductance  ${\cal{G}}$ across the ring, at
1135: the Fermi energy.  Within Landauer's approach, ${\cal{G}}$ is given by
1136: %
1137: \beq
1138: {\cal{G}}=\frac{e^2}{\h}\sum_{\sigma,\sigma'}
1139:  \left| \mathcal A (\sigma;\sigma'|E_0)\right|^2
1140: \:\:\:\: . 
1141: \label{ec.1}
1142: \eneq
1143: %
1144: We will here consider the dependence on the external magnetic field
1145: ($\phi /\phi _0$) and on the spin-orbit strength $k_{SO} R
1146: $\cite{nota} both in absence and in presence of dephasing at the
1147: contacts.  The various amplitudes in Eq.(\ref{ec.1}) have been
1148: numerically computed, as discussed in Sec.s(III-VI).
1149: %
1150: \begin{figure}[!htp]
1151:     \centering \includegraphics[width=\columnwidth]{GvsB.eps}
1152: \caption{(color online) $a) $ Magnetoconductance of an ideal ring as a
1153:   function of the magnetic field $\phi/\phi_0$. The Fourier transform
1154:   (FT) of the magnetoconductance ({\sl in the inset}) shows only the
1155:   AB peak at freq. $\phi_0^{-1} $.  $b) $ Magnetoconductance of a
1156:   realistic ring in which the S matrix of Eq.(\ref{matr}) regulates the
1157:   scattering of the electron at the leads. As a consequence of the
1158:   backscattering, the FT of the magnetoconductance ({\sl inset})
1159:   shows higher order frequencies.}
1160: \label{GvsB}
1161: \end{figure}
1162: %
1163: In Fig.(\ref{GvsB}) we show the magnetoconductance of the ring in the
1164: absence of RSOI ($k_{SO}R =0$).
1165: %
1166: In panel a) of Fig.(\ref{GvsB}), only the path of
1167: Fig. (\ref{noWLpaths}, $1a)$ has been  considered,
1168:  i.e., full transmission across the ring is
1169: assumed, as it would be the case for ideal coupling to the leads.  The
1170: corresponding Fourier spectrum is showed in the corresponding inset. 
1171: The well known
1172: Ahronov-Bohm sinusoidal pattern implies that just the fundamental
1173: frequency $\phi_0^{-1}$ appears.
1174: 
1175: To make the model more realistic, we allow for higher order looping of
1176: the electron within the ring.  In Ref.\cite{noiletter}, only the paths
1177: of the kind of Fig.(\ref{noWLpaths},2a),\ref{noWLpaths},2b)) were
1178: included.  Here, we consider also the paths of the kind of
1179: Fig.(\ref{WLpaths}) in which the electron can be backscattered at the
1180: leads.  We  use here $\bar r=0$ in the scattering matrix between the
1181: arms and the leads, which means that no backreflection in the incoming
1182: lead is present. The magnetoconductance of the ring, in this regime,
1183: is showed in Fig.  (\ref{GvsB}b). Because of the inclusion of time
1184: reversed  paths (TRP) within the ring, we see that higher
1185: order frequencies appear in the Fourier spectrum. In particular, the
1186: inset shows a peak at $2/\phi_0$ which is the signature of weak
1187: localization\cite{nonloso}.
1188: %
1189: \begin{figure}[!htp]
1190:     \centering \includegraphics[width=\columnwidth]{barge.eps}
1191: \caption{(color online)Conductance as a function of $\phi/\phi_0$ {\sl
1192: (left panels)} and $k_{SO}R$ {\sl (right panels)} for ideal {\sl
1193: (black curves)} and realistic {\sl (red curves)} contacts. An
1194: increasing amount of dephasing at the contacts is also included: {\sl
1195: from top to bottom:} $\zeta=\pi/3,\pi,2\pi$}.
1196: \label{GvsBSO}
1197: \end{figure}
1198: %
1199: We also include some dephasing due to diffusiveness in the contacts by
1200: adding a random phase $z\in (-\zeta,\zeta ) $ for each scattering at
1201: the leads.  In Fig.(\ref{GvsBSO}), we report the conductance {\cal{G}}
1202: {\it vs.}  $ \phi / \phi_0$ , with $k_{\text {SO}}R = 0$ {\sl (left
1203: panels)} or {\it vs.}  $k_{SO} R$ with $\phi/\phi_0=0$ {\sl (right
1204: panels)}. These are averaged over $N=1000$ realizations of disorder,
1205: and plotted increasing the window of phase randomness ($\zeta = \pi/3
1206: ,\pi, 2\pi$ from top to bottom). The black curves refer to the case of
1207: Fig(\ref{GvsB} $a)$) (ideal contacts) while the red curves refer to
1208: the case of realistic contacts (Fig(\ref{GvsB} $b)$), with
1209: $\bar{r}=0$.
1210: 
1211: By comparing the top left panel of Fig.(\ref{GvsBSO}) with
1212: Fig.(\ref{GvsB}), we see that the ring is rather insensitive to small
1213: dephasing at the contacts. By increasing the amount of dephasing 
1214: {\sl (middle and bottom left panels in Fig.(\ref{GvsBSO}))}  we find that the
1215:  sensitiveness is larger  in the case of realistic contacts.
1216:  This is due  to the fact that for realistic coupling, the
1217: electrons in the ring can experience higher order paths, since it scatters
1218:  with the leads many times.  
1219: 
1220: In the right panel of Fig.(\ref{GvsBSO}), we plot the DC conductance
1221: {\it vs.}  $k_{\text {SO}}R$ at $\phi/\phi_0 =0 $ for both ideal
1222: contacts and realistic contacts (and $\bar{r}=0$) {\sl (black
1223: and red lines in each box)}, with an increasing phase randomization 
1224: {\sl (boxes
1225: from top to bottom with $\zeta = \pi/3 ,\pi, 2\pi$)}, averaged over
1226: $N=1000$ realizations. In the case of ideal contacts and little
1227: dephasing {\sl (top right panel black curve)}, we see again the
1228: quasiperiodic oscillation of the conductance reproducing the
1229: localization conditions at the expected values of $k_{\text {SO}}R$
1230: \cite{frustaglia,molnar,souma,dario,noiletter}.  When including higher
1231: order processes, interference effects give rise to a slightly
1232: different pattern.  In the case of realistic contacts, we note that
1233:  the device is seriously  affected by dephasing, mainly because
1234: including  the TRPs contributing to the transmission   amplitude
1235:  increases the number of scattering processes
1236: at the leads. Indeed, when  the 
1237: dephasing is  quite large, it gives rise to
1238:  random oscillations that  are not
1239: averaged out,  so  that they  wash  out the conductance oscillations.
1240:  The effect takes 
1241: place  for  
1242: $\zeta\sim\pi$  when TRPs are included, in contrast to $\zeta \sim 2\pi$ 
1243: when  the TRPs are  absent.  As regular magnetoconductance oscillations 
1244: are experimenally
1245: observed \cite{nitta,yau,morpurgo,kato,molenkamp} with  little
1246: precentage of contrast between maxima and minima, we conclude that, in
1247: real samples, dephasing is ubiquitous.
1248: 
1249: \section{Spin Transmission}
1250: 
1251: In this Section we calculate the rotation of the spin of the electron
1252: transmitted through the ring.  We first consider an incoming electron
1253: beam with in-plane spin polarization (let's say, polarized along the
1254: $x$ direction ). The spin rotation is measured by calculating 
1255:  the average value of the outgoing spin:
1256: %
1257: \beq
1258: \langle S_z\rangle= \frac{\left\langle\Psi_{out}\right|S_z\left|\Psi_{out}\right\rangle}{\left\langle
1259: \Psi_{out}\right.\left|\Psi_{out}\right\rangle }\;.
1260: \label{szout}
1261: \eneq
1262: %
1263: Since in the previous Section we have shown that higher order looping
1264: just adds subleading higher order harmonics to the conductance, here we
1265: focus on the case of ideal contacts, that is, we include in the
1266: calculation only paths as the ones of Fig.(\ref{noWLpaths},1a).  The
1267: in plane polarization can be considered as a superposition of equal
1268: weighted z-polarized spin components.  In the absence of RSOI
1269: ($ k_{SO} R =0 $),
1270: opposite spin polarizations do not interfere with each other. As a
1271: consequence, the total expected $\langle S_z \rangle $ component keeps
1272: zero at the exit.  Fig.(\ref{spintramiscela}) shows the
1273: magnetoconductance for increasing  $ k_{SO} R$, and the
1274: corresponding expected spin component polarized along the $z-$axis 
1275:  at the exit of the ring. The
1276: Zeeman term is on the diagonal of the spin  Hamiltonian of Eq.(\ref{three_1})
1277:  and operates to keep the $z-$components of the spin polarization fixed,
1278:  while the RSOI is offdiagonal and tends to favor inplane spin polarization.
1279: This implies that when the magnetic field increases 
1280: ($\phi /\phi _0 >> k_{SO} R$)  the spin polarization gets frozen  to the
1281:  incoming polarization.
1282: This is the reason why,in Fig.(\ref{spintramiscela}), at high fields,
1283:   the   trasmitted spin polarization is inplane. Incidentally  we observe 
1284: that this result should not be expected  in real systems in which spin 
1285: relaxation can occur due to electron-phonon  interaction or other mechanisms
1286:  as hyperfine interaction with nuclear spins.
1287: Spin relaxation would induce flipping of that spin component that is 
1288: energetically  unfavourable and the final
1289: transmission of the spin will result  to be partly out of the $x-y$ plane.
1290:  On the contrary,
1291: when  $\phi /\phi _0 \sim
1292: k_{SO} R $ the competition of the Zeeman and the RSOI induces  
1293: a coherent rotation of the  spin while the electron travels
1294: along the ring. 
1295: Fig.(\ref{spintramiscela})  shows that, when $\phi /\phi _0 \sim
1296: k_{SO} R$, the spin is moved significantly out of the $x-y $ plane,
1297: consistently affecting the AB oscillations.
1298: %
1299: \begin{figure}[!htp]
1300:     \centering
1301:     \includegraphics[width=\columnwidth]{spintramiscela.eps}
1302: \caption{Magnetoconductance and expectation value of the outgoing $\hat
1303: z $ spin component for an incoming spin in the $x$ direction at
1304: different values of the RSOI strength ($k_{SO}R = 0,1,2,4 $ indicated
1305: in the pictures).}
1306: \label{spintramiscela}
1307: \end{figure}
1308: %
1309: To better understand what happens when $\phi /\phi _0 \sim k_{SO} R$,
1310: we isolate the  spin ``up'' polarization for the incoming particle 
1311: in what follows ($\left\langle\Psi_{in}\right|S_z\left|\Psi_{in}\right\rangle/\left\langle
1312: \Psi_{in}\right.\left|\Psi_{in}\right\rangle=1$)  and   
1313: we separately plot in Fig.(\ref{spintraup}) the two contributions to
1314: the conductance $G_{up-up} $ ({\sl full line}) and $ G_{down-up} $
1315: ({\sl dotted line}), for opposite polarizations of the outgoing
1316: electron. $G_{up-up} $ is the contribution to the conductance 
1317: arising from the particle flux that mantains the same polarization at the exit as the incoming one, while  $ G_{down-up} $ refers to a particle flux 
1318: having  the opposite polarizations at the exit with respect to the one 
1319: at the entrance.  Of course, when $ k_{SO} R = 0 $, the electron spin is in
1320: the ``up''direction for any $\phi/\phi_0$. When both RSOI and magnetic
1321: field are present, with $\phi /\phi _0 >> k_{SO} R$, the spin
1322: polarization is steadly in the $z-$ direction, except for sharp
1323: reversals at flux values $\phi _0 m/2 $ ($m$ integer).  However,
1324: $G_{down-up} $ is vanishingly small at these places, together with
1325: $G_{up-up} $. Therefore the conductance vanishes at these points
1326: anyhow, and the transmitted spin polarization is fully up, except for these
1327: points.  On the contrary, in the parameter intervals characterized by
1328: $\phi /\phi _0 \sim k_{SO} R$, both $ G_{up-up} $ and $ G_{down-up} $
1329: are non vanishing (see Fig.\ref{spintraup}), so that the spin is rotated 
1330: at the exit, with
1331: nonvanishing transmission amplitude.
1332: %
1333: \begin{figure}[!htp]
1334:     \centering \includegraphics[width=\columnwidth]{spintraup.eps}
1335: \caption{Separate contributions to the magnetoconductance for
1336: different outcoming spin polarizations, $ G_{up-up} $ ({\sl full line})
1337: and $ G_{down-up} $ ({\sl dotted line} ) compared to the expected
1338: value of the outgoing $\hat z $ spin component. The incoming spin is
1339: polarized ``up'' .  Different values of the RSOI strength are reported
1340: ($k_{SO}R = 0,1,2,4 $ indicated in the picture ).}
1341: \label{spintraup}
1342: \end{figure}
1343: %
1344: %
1345: 
1346: We now examine in more detail the expected dependence of the outgoing
1347: spin polarization on the RSOI, at zero magnetic field.  For an
1348: incoming electron polarized with spin ``up'', Fig.(\ref{spinwire}(a))
1349: shows that a large enough RSOI produces a rotation of the spin which
1350: points down at the exit for any $k_{SO}R$ value. Meanwhile, the total
1351: conductance oscillates with $k_{SO}R$. This finding also appears in
1352: Ref.\cite{souma} and is quite remarkable, because it is the
1353: consequence of the interference between the two arms of the ring.  In
1354: order to point out the role of quantum interference, we discuss here,
1355: for comparison, what happens by transporting a spinful electron along
1356: a single arm of the ring (the upper one).  When just one arm is
1357: considered, (see Fig(\ref{spinwire}b)) spin polarization oscillates,
1358: as a function of $k_{SO} R $ \cite{datta}, while the conductance is
1359: always unitary, because of the conservation of the particle flux. In
1360: the limit of large $k_{SO} R$ the propagation amplitude for the
1361: travelling electron acquires a simple analytical form. From
1362: Eq.(\ref{propagatorespin}) we see that, in the representation of the
1363: instantaneous spin eigenvectors, the spin propagator at the exit time
1364: $t_f$ (with $\omega _o t_f = \pi )$, in this limit is diagonal:
1365: %
1366: \beq
1367: U^{u\rightarrow}(t_f,0)=
1368: \left(
1369: \begin{array}{c c}
1370: i e^{-i\pi\gamma}    &  0\\
1371: 0 & -i e^{i\pi\gamma}   
1372: \end{array}
1373: \right) \:\: ,
1374: \eneq
1375: %
1376: so that the spin appears not to be rotated at the exit in the rotating
1377: reference frame. However, in order to move from the representation of
1378: the instantaneous spin eigenvectors to the reference basis, one has to
1379: perform the transformation with the unitary $B$ matrix of
1380: Eq.(\ref{five}) , with $\vartheta = \pi /2$. The spin part of the
1381: propagator for an electron travelling into the upper arm (upper path
1382: of Fig.(\ref{noWLpaths},1a)) is then:
1383: %
1384: \bea
1385: U^{u\rightarrow}_{spin}(t_f,0)=
1386: B(t_f)U^{u\rightarrow}(t_f,0)B^\dagger(0)
1387: \nonumber\\
1388: =\left(
1389: \begin{array}{c c}
1390: i \cos(\pi \gamma)    &   \sin(\pi \gamma)\\
1391:  - \sin(\pi \gamma) & -i \cos(\pi \gamma)   
1392: \end{array}
1393: \right) \:\: .
1394: \enea
1395: %
1396: If we inject $up$ electrons in the upper arm only, the expectation
1397: value of the the outcoming $S_z$ defined in Eq.(\ref{szout}) is:
1398: $\langle S_z\rangle=\cos(2\pi\gamma)=\cos(\pi k_{SO} R)$ (note that
1399: for large enough $\gamma$ this result well approximates the red-full
1400: line in Fig.(\ref{spinwire}b)). The conductance is unitary, $G=2e^2/h$
1401: (the black-dashed line in Fig.(\ref{spinwire}b), as no interference
1402: takes place.
1403: %
1404: 
1405: We now go back to the transmission along both arms simultaneously and
1406: examine the resulting interference.  According to the rules given
1407: after Eq.(\ref{propagatorespin}), in the same limiting case as above,
1408: the propagator accounting for transmission of incoming $up$ spins
1409: across the ring is:
1410: %
1411: \beq
1412: U^{u\rightarrow+d\rightarrow}_{spin}(t_f,0)=
1413: \left(
1414: \begin{array}{c c}
1415: 0    &  2 \sin(\pi \gamma)\\
1416:  - 2\sin(\pi \gamma) & 0
1417: \end{array}
1418: \right) \:\: ,
1419: \eneq
1420: %
1421: so that the spin at the exit is reversed.  In fact, in the expectation
1422: value of Eq.(\ref{szout}), the oscillations in the numerator
1423: compensate those in the denominator, eventually giving $\langle
1424: S_z\rangle=-1$ (for large enough $\gamma$ this result well approximate
1425: the red-full line in Fig.(\ref{spinwire}a)).  The conductance however
1426: oscillates according to $G/(2e^2/h)= 2\sin^2(\pi k_{SO}R/2)$, as
1427: plotted in the black-dashed line in (Fig.\ref{spinwire}a)).
1428: 
1429: It is quite remarkable that this result is only found at zero magnetic
1430: field.  Indeed, no matter how small $B$ is, the time reversal symmetry
1431: is broken and the spin oscillates with $k_{SO}R $ (see
1432: Fig.(\ref{brokensymmetry},a).  However, for very small magnetic field
1433: these oscillations are confined close to special values of $ k_{SO} R
1434: = 2l $($l$ integer) and display a Lorentzian shape around these
1435: points.  The role of the magnetic field is to make the oscillations
1436: broader.
1437: 
1438: To summarize, there are two limiting conditions in the outgoing spin
1439: polarization, for incoming ``up'' spin polarization: $a)$ {\sl the
1440: zero RSOI } which leaves the incoming spin polarization unchanged,
1441: provided no relaxation takes place; $b)$ {\sl the zero magnetic flux
1442: case } in which the RSOI produces a flip of the spin at the exit. It
1443: is interesting that when the flux $\phi $ is an integer number of flux
1444: quanta $ \phi _0$, the crossover between case $a)$ and case $b) $,
1445: with $k_{SO} R$ increasing from the value zero to values $ k_{SO} R >>
1446: \phi /\phi _0 $ is rather sharp.  This is shown in
1447: Fig. (\ref{brokensymmetry},b) where the expectation value of the
1448: outcoming spin is plotted vs. $k_{SO} R$ for different integer values
1449: of $\phi/\phi_0$.  For values of $\phi/\phi_0 >> k_{SO} R $ the
1450: outgoing spin polarization is the same as that at the entrance, ( $up$
1451: in the picture).  On the contrary, by increasing $k_{SO} R$ at non
1452: zero $B$ field, we see again a pattern similar to the one of
1453: Fig. (\ref{brokensymmetry},a), but shifted to higher values of $k_{SO}
1454: R$.
1455: 
1456: \begin{figure}[!htp]
1457:     \centering
1458:     \includegraphics[width=\columnwidth]{spinwire.eps}
1459: \caption{ (color online) Conductance ({\sl broken line}) and spin
1460: polarization ({\sl full line}) of the outgoing electron for $B=0$, as
1461: a function of the RSOI. The spin of the incoming electrons is
1462: polarized ``$up$'' : $a)$ The ring case with ideal contacts. $b)$ a
1463: single wire of the same length and curvature as one of the ring arms.}
1464: \label{spinwire}
1465: \end{figure}
1466: %
1467: \begin{figure}[!htp]
1468:     \centering
1469:     \includegraphics[width=\columnwidth]{SvsAvBs.eps}
1470: \caption{(color online) Expectation value of the outgoing spin for
1471: incoming spin ``$up$'' polarized electrons as a function of the RSOI
1472: for different values of the magnetic flux : $a)$ $\phi/\phi_0$ zero or
1473: very small; $b)$ increasing integer values of $\phi/\phi_0$. }
1474: \label{brokensymmetry}
1475: \end{figure}
1476: 
1477: \section{Conclusions}
1478: 
1479: 
1480: To conclude, we have employed a path integral real time approach to
1481: compute the DC conductance of a ballistic one dimensional mesoscopic
1482: ring in both external electrical and magnetic fields orthogonal to the
1483: ring plane.The spinful electron experiences a Rashba spin orbit
1484: interaction and the Zeeman term. We employ a piecewise saddle point
1485: approximation for the orbital motion, but we implement the full
1486: scattering matrix at the leads and sum over all the possible higher
1487: order paths up to convergency of the result. Our approach goes beyond
1488: other recent semiclassical calculations.  Our theory is
1489: nonperturbative and separates the adiabatic spin dynamics from the non
1490: adiabatic one by using the rotating frame for the spin travelling
1491: around the ring. In practice, we diagonalize the time dependent spin
1492: Hamiltonian in the representation of the spin eigenvectors of the
1493: instantaneous Hamiltonian.  This allows us to explore a wide range of
1494: Hamiltonian parameters, ranging from the limit of strong magnetic
1495: field and weak Rashba SOI to the opposite case. In both extreme
1496: regimes our piecewise saddle point approximation is very efficient as
1497: quantum fluctuations with flipping of the spin has little influence on
1498: the orbital motion. This is also seen from the number of paths
1499: required to gain full convergency.  As explained in Sec. $V$ the
1500: separation of adiabatic from non adiabatic spin dynamics shows that in
1501: the intermediate regimes our approximation is less justified, but
1502: nevertheless, the results it produces seem to be in rather good
1503: agreement with recent numerical calculations
1504: \cite{frustaglia, molnar, shen,
1505: dario,souma} and experiments\cite{nitta, yau,
1506: morpurgo,kato,molenkamp}.  When we include also time reversed paths,
1507: the Fourier transform of the magnetoconductance shows the typical
1508: $\phi_0/2$ peak due to weak localization\cite{nonloso}.This would be the 
1509: only surviving contribution if an ensemble average or different rings 
1510: were measured\cite{koga}.
1511: 
1512: We have also allowed for nonideal couplings between ring and leads as
1513: we account for dephasing effects due to diffusiveness at the
1514: contacts. The results satisfactorily compare with experiments where
1515: the contrast between maxima and minima in the interference fringes is
1516: always few tens of percentage of the background DC signal.
1517: 
1518: \begin{acknowledgments}
1519: We acknowledge financial support by the  Italian Ministry of Education
1520: (PRIN) and by the CNR within ESF Eurocores Programme FoNE (contract
1521: N. ERAS-CT-2003-980409).
1522: \end{acknowledgments}
1523: 
1524: 
1525: \appendix 
1526: \section{ Motion of  a classical spin in  a rotating magnetic field}
1527: 
1528: In this appendix we derive the classical equations of motion for the
1529: spin from the Lagrangian in Eq.(\ref{venti}), by assuming for the
1530: orbital coordinate the saddle point solution $\dot{\varphi}$ =
1531: constant. Once the orbital motion is dealt with in this way, the
1532: Lagrangian for the spin degrees of freedom is given by (besides a
1533: constant contribution)
1534: %
1535: \beq
1536: \tilde{\cal L} [ \Theta , \Phi , \dot{\Phi } ]  / \hbar = \left( \frac{1 -  
1537: \cos \Theta }{2} \right) \dot{\Phi} + \vec{{\cal B}} \cdot \vec{S}
1538: \:\:\:\: , 
1539: \label{al1}
1540: \eneq
1541: 
1542: where  the effective time  dependent  magnetic field is 
1543: ${\cal{B}} \equiv  ({\cal{B}}_+,{\cal{B}}_-,{\cal{B}}_z ) = 
1544: \left ( \gamma \dot{\vt} \: e^{i\vt}, \gamma \dot{\vt} \: 
1545: e^{-i\vt},-\omega _c/2 \right ) $.   
1546: %
1547: To derive the equations of motion from a variational principle, we 
1548: write the Berry phase term in the total spin action as
1549: %
1550: \beq
1551: \tilde{S}_B = 
1552: \int_{ \Phi ( 0 ) }^{ \Phi ( t_f )}  \: d t \:  \left( \frac{1 -  
1553: \cos \Theta }{2} \right) d \Phi = \int_\Sigma \sin \Theta \: d \Theta \wedge
1554: d \Phi 
1555: \:\:\:\: , 
1556: \label{al3}
1557: \eneq
1558: 
1559: %
1560: where $\Sigma$ is the spherical triangle with vertices given by the
1561: north pole on the sphere and by the points with coordinates $( \Theta
1562: ( 0 ) , \Phi ( 0 ) ) $ , $ ( \Theta ( t_f ) , \Phi ( t_f ) )$.  Let $
1563: ( t , u ) $ be a parametrization of the spherical triangle, such that
1564: $ \vec{S} ( t , 1 ) = \vec{S} ( t )$, and $\vec{S} ( t , 0 ) = (0 , 0
1565: , 1)$. Thus, one may rewrite the action $\tilde{S}_B$ in
1566: Eq.(\ref{al3}) as
1567: %
1568: \beq
1569: \tilde{S}_{B} = \int_0^T \: d t \: \int_0^1 \: d u \: {\bf S} \cdot \left[ 
1570: \frac{ \partial {\bf S}}{ \partial t} \times \frac{ \partial {\bf S}}{
1571: \partial u} \right]
1572: \:\:\:\: . 
1573: \label{e.9}
1574: \eneq
1575: 
1576: %
1577: To derive the equations of motion, we consider a variation
1578: $ \vec{S} ( t , u ) \rightarrow \vec{S} ( t , u ) + \delta \vec{S} 
1579: ( t , u )$ such that $\vec{S} ( T , u ) $ 
1580: and $\vec{S} ( 0 , u )$ are "locked", that is, $ \delta \vec{S}  ( 0 , u ) 
1581: = \delta \vec{S} ( T , u ) = 0$. 
1582: %
1583: Since $ [ \vec{S} ( t , u ) ]^2 = 1$ $\forall t , u $, one gets $\vec{
1584: S} \cdot \frac{ \partial \vec{ S}}{ \partial t} = \vec{ S} \cdot
1585: \frac{ \partial \vec{ S}}{ \partial u} = 0 $, As a consequence,
1586: $\frac{ \partial \vec{ S}}{ \partial t} \times \frac{ \partial \vec{
1587: S}}{ \partial u} $ is parallel to $\vec{S}$.  As $\delta \vec{S} \cdot
1588: \vec{S} = 0$, this implies that
1589: %
1590: \beq
1591: \int_0^T \: d t \: \int_0^1 \: d u \: \delta \vec{S} \cdot \left[ 
1592: \frac{ \partial \vec{ S}}{ \partial t} \times \frac{ \partial \vec{ S}}{
1593: \partial u} \right] = 0 
1594: \:\:\:\: . 
1595: \label{e.9.b}
1596: \eneq
1597: 
1598: %
1599: Thus, by integrating by parts we see that the only nonzero variation of 
1600: $\tilde{S}_B$ is given by the boundary term 
1601: %
1602: \beq
1603: \delta \tilde{S}_B = \int_0^{t_f } \: d t \: \delta \vec{S} ( t ) 
1604: \cdot \left[ \vec{S} ( t  ) \times \frac{ \partial  \vec{S} ( t ) }{
1605: \partial t} \right]
1606: \:\:\:\: ,
1607: \label{e.12b}
1608: \eneq
1609: 
1610: %
1611: where we have used the fact that  $\vec{S} ( t , 1 ) = \vec{S} ( t )$. 
1612: %
1613: By equating to zero the total variation of the action,  one obtains
1614: %
1615: \beq
1616: \vec{S}\times \frac{d\vec{S}}{dt} =\vec{\mathcal B}
1617: \:\:\:\: , 
1618: \label{appe1}
1619: \eneq
1620: that is, the classical equations of motion we used in section IV.
1621: %
1622: To show that Eqs.(\ref{appe1}), when the spin components are expressed
1623: in polar coordinates, are equivalent to Eqs.(\ref{em1},\ref{em2}), 
1624: let us set  $\omega =  \gamma \dot{\vt} $ and $ \Omega = -\omega _c/2$. 
1625: Also, we define
1626: %
1627: \bea
1628: S_{+}=S_{x}+iS_{y}\hspace*{2cm}{\mathcal B}_{+}={\mathcal B}_{x} +
1629: i{\mathcal B}_{y}\nonumber\\ S_{-}=S_{x}-iS_{y}\hspace*{2cm}{\mathcal
1630: B}_{-}={\mathcal B}_{x}-i{\mathcal B}_{y}\vspace*{1cm}\nonumber
1631: \:\:\:\: .
1632: \enea
1633: In terms of the new variables, the equations of motion are given by 
1634: %
1635: \bea
1636: \frac{d S_{z}}{dt}=\frac{i}{2}({\mathcal B}_{+}S_{-}-{\mathcal B}_{-}S_{+})
1637: \nonumber\\
1638: \frac{d S_{+}}{dt}= i({\mathcal B}_{z} S_{+}-{\mathcal B}_{+}S_{z})\nonumber\\
1639: \frac{d S_{-}}{dt}=-i({\mathcal B}_{z} S_{-}-{\mathcal B}_{-}S_{z}) \label{e2}
1640: \:\:\:\: . 
1641: \label{appe2}
1642: \enea
1643: %
1644: or:
1645: %
1646: \bea
1647: \frac{d m(t)}{dt} =i\:\left ( \Omega -\dot\varphi \right )\:p(t)-2\:i
1648:   \omega  \:S_{z} \nonumber \\
1649: \frac{d p(t)}{dt} =i\:\left (\Omega -\dot{\varphi} \right )\:m(t)
1650:  \nonumber\\
1651: \frac{dS_{z}}{dt} =-\frac{i\omega }{2}\:{m(t)}\:
1652: \label{e4}
1653: \enea
1654: %
1655: where
1656: %
1657: \bea
1658: p(t)=S_{+}e^{(-i\varphi)}+S_{-}e^{(i\varphi)}\nonumber \\
1659: m(t)=S_{+}e^{(-i\varphi)}-S_{-}e^{(i\varphi)}\nonumber
1660: \:\:\:\: , 
1661: \enea
1662: %
1663: and  $ 1= 4|S|^2=4 {S_{z}}^2+ (p^2-m^2 ) $.
1664: %
1665: By introducing $b= \Omega -\dot \vt $,  we obtain:
1666: %
1667: \bea
1668: \frac{d(m(t) + p(t))}{dt}  &=&  i b \left( m(t) + p(t) \right) -2 i\omega
1669:   \: S_z (t)   \label{m3.1}\\
1670: \frac{d {S}_z(t)}{dt}   &=& - i \omega \:  m(t)          \label{m3.2}
1671: \enea
1672: %
1673: Resorting to the polar coordinates  $(\Theta  ,\Phi )$ for the spin $\vec{S}$,
1674: we get:
1675: %
1676: \bea
1677: \left[\dTh \ct +i \left( \dP- \Omega  \right)\st \right] e^{i \chi}
1678:  + i \ct \alpha \df=0 \label{una} \\
1679: \left[\dTh - \omega \:  \sin{\chi}\right]  \st = 0 
1680: \:\:\:\: . 
1681: \label{due}
1682: \enea
1683: %
1684: Eq.(\ref{due}) is the same as Eq.(\ref{em1}). The real part 
1685:  of Eq.(\ref{una})  is proportional  to the imaginary part:  both give  
1686: Eq.(\ref{em2}) when  equated to  zero,  which completes the proof. 
1687: %
1688: \section{The  spin propagator} 
1689: %
1690: In order to find the propagator of the Berry Hamiltonian $ \hat{H}_A $
1691: of Eq.(\ref{berryham}),
1692:  we  solve
1693: the system of differential  Eq.(\ref{addi6}), in the 
1694: representation of the instantaneous eigenvectors:
1695: \bwt
1696: \beq
1697: i \frac{d}{dt}\left ( \begin{array}{c}
1698:     u_+  \\
1699:     u_-     \end{array} \right ) \:=
1700: \left ( \begin{array}{cc}
1701:    r+ \wo \smqt   & \wo \smt \cmt e^{-i \wo t}  \\
1702:    \wo \smt \cmt  e^{i \wo t}&  -r- \wo \smqt
1703: \end{array} \right )
1704: \left ( \begin{array}{c}
1705:     u_+  \\
1706:     u_-
1707: \end{array} \right )\label{addi6_1}
1708: \:\:\:\: . 
1709: \eneq
1710: %
1711: \ewt 
1712: %
1713: To solve Eq.(\ref{addi6_1}), first of all, we switch to a
1714: time-independent coefficient matrix by defining:
1715: %
1716: \beq
1717: \left ( \begin{array}{c}
1718:     y_+  \\
1719:     y_-     \end{array} \right ) \:=
1720: \left ( \begin{array}{cc}
1721:      e^{ +i \frac{  \wo }{2} t  } & 0  \\
1722:      0 &   e^{- i \frac{\wo}{2}t   }
1723: \end{array} \right )
1724: \left ( \begin{array}{c}
1725:     u_+  \\
1726:     u_-
1727: \end{array} \right )\:.
1728: \label{T}
1729: \eneq
1730: %
1731: By setting
1732: \[Y=
1733: \left ( \begin{array}{c}
1734:     y_+  \\
1735:     y_-     \end{array} \right ) \:\:\: ; \:\: 
1736: W=\left ( \begin{array}{c}
1737:     u_+  \\
1738:     u_-
1739: \end{array} \right )\:,\]
1740: we define the matrix $T$ through
1741: \beq Y\:=\:T\:W,\;\;\;W\:=\:T^{-1}\:Y\:.\eneq
1742: 
1743: Eqs.(\ref{addi6_1}) now read:
1744: \bea
1745: i \frac{ d y_+}{ d t } ( t ) = (r-\frac{\wo}{2}cos(\vartheta)) y_
1746: + ( t ) + \frac{  \wo}{ 2 }
1747: \sin  \vartheta   y_- ( t ) \:,\nonumber \\
1748: i \frac{ d y_-}{ d t } ( t ) =
1749: + \frac{  \wo}{ 2 }  \sin  \vartheta   y_+ ( t )
1750: +(-r+\frac{\wo}{2}cos(\vartheta)) y_- ( t )\:.
1751: \enea
1752: Now we define $ r' = r - \frac{ \wo}{2} \cos \vartheta $ and $s=
1753: \frac{ \wo}{2} \sin \vartheta $, so that in matrix form we have:
1754: \begin{eqnarray}
1755: i \frac{d}{dt}
1756: \left ( \begin{array}{c}
1757:  y_+  \\
1758:  y_-     \end{array} \right ) \:=
1759: \left ( \begin{array}{cc}
1760: r' & s \\
1761: s  &  -r'
1762:  \end{array} \right )
1763: \left ( \begin{array}{c}
1764:  y_+  \\
1765:  y_-     \end{array} \right ) \:,
1766: \label{dicia}
1767: \end{eqnarray}
1768: in a compact form we can  rewrite the last equation as:
1769: \beq
1770: i \frac{d}{dt}Y\:=\:C\:Y\:,
1771: \label{matr1}
1772: \eneq
1773: which defines the  matrix $C$.\\ We now decouple the
1774: previous system of equation by diagonalizing the matrix $C$.
1775: Its eigenvalues are $\lambda=\pm \epsilon=\pm\sqrt{r'^2+s^2}$ and the
1776: matrix that diagonalizes $C$ is
1777: \beq
1778: P=\left ( \begin{array}{cc}
1779: 1 & \frac{r'-\epsilon}{s} \\
1780: \frac{\epsilon-r'}{s}  &  1
1781:  \end{array} \right ) \:.
1782: \eneq
1783: Its inverse is
1784: \beq
1785: P^{-1}=\left ( \begin{array}{cc}
1786:  \frac{s^2}{2\epsilon(\epsilon -r')}  & \frac{(\epsilon-r')s}{2
1787: \epsilon(\epsilon -r')} \\
1788: -\frac{(\epsilon-r')s}{2\epsilon(\epsilon -r')}  & \frac{s^2}{2\epsilon(
1789: \epsilon -r')}
1790:  \end{array} \right ) \:.
1791: \eneq
1792: 
1793: Eq.(\ref{matr1}) now reads:
1794: \beq
1795: i \frac{d}{dt} P^{-1}\: Y\:=\:P^{-1}\:C\:P\:P^{-1}\:Y\:,
1796: \eneq
1797: which, by defining $V\:=\:P^{-1}\:Y$, becomes
1798: \beq
1799: i \frac{d}{dt} V\:=\left ( \begin{array}{cc}
1800: \epsilon  & 0          \\
1801:     0     &  -\epsilon
1802:  \end{array} \right ) \:      V\:.
1803: \eneq
1804: Its formal solution is:
1805: \beq
1806:  V(t)\:=\left ( \begin{array}{cc}
1807: e^{-i\epsilon (t-t')}  & 0          \\
1808:     0     &  e^{i\epsilon (t-t')}
1809:  \end{array} \right ) \:      V(t')\:,
1810: \eneq
1811: or, in matrix form
1812: \beq
1813: V(t)\:=\:S(t-t')\:V(t')\:.
1814: \eneq
1815: Now we apply inverse transformations, in order to obtain the full
1816: Schr\"odinger propagator, that is the matrix transformation between
1817: ($W(t')$ and $W(t)$).
1818: \[
1819: W(t)\:=\:T^{-1}(t)\:P\:S\:P^{-1}\:T(t')\:W(t')
1820: \;\;,\]
1821: where $P$ is time independent. 
1822: The full evolution operator in the adiabatic basis is:
1823: \beq
1824: U(t,t')\:=\:T^{-1}(t)\:P\:S\:P^{-1}T(t')\:;
1825: \eneq
1826: By performing all the matrix products, we obtain  the result 
1827: of  Eq.(\ref{propagatorespin}), as 
1828: given  in  the text.
1829: 
1830: 
1831: \begin{thebibliography}{99}
1832: 
1833: \bibitem{aharonovbohm} Y. Aharonov, D. Bohm, Phys. Rev. {\bf 115} 485
1834:   (1959).
1835: 
1836: \bibitem{webb} S. Washburn and R. A. Webb, Rep. Prog. Phys. {\bf 55}, 1311 
1837: (1992).
1838: 
1839: \bibitem{anandans} J. Anandan, Science {\bf 297}, 1656 (2002).
1840: 
1841: \bibitem{berry} M. V. Berry, Proc. R. Soc. London, A {\bf 392}, 45 (1984).
1842: 
1843: \bibitem{anandan}Y. Aharonov, J. Anandan, Phys. 
1844: Rev. Lett. {\bf 58}, 1593 (1987).
1845: 
1846: \bibitem{rashba}E.I. Rashba, Fiz. Tverd. Tela {\bf 2}, 1224 (1960) [Sov.Phys. -
1847: Solid State {\bf 2}, 1109 (1960),Y.A. Bychkov, E.I. Rashba, J.Phys.{\bf C17},
1848:  6039 (1984).
1849: 
1850: \bibitem{meijer} F. E. Meijer, A. F. Morpurgo, T. M. Klapwijk, T. Koga
1851: and J. Nitta, Phys. Rev. B {\bf 70} 201307 (2004).
1852: 
1853: \bibitem{miller} J. B. Miller, D. M. Zumb\"uhl, C. M. Marcus,
1854:                  Y. B. Lyanda-Geller, D. Goldhaber-Gordon, K. Campman
1855:                  and A. C. Gossard, Phys. Rev. Lett. {\bf 90} 076807
1856:                  (2003).
1857: 
1858: \bibitem{aronov} A. G. Aronov and Y. L. Lyanda-Geller, Phys. Rev. 
1859: Lett.{\bf 70}, 343  (1993).
1860: 
1861: \bibitem{ac}  Y.Ahronov and A. Casher Phys. Rev. Lett. {\bf 53}, 319
1862:   (1984).
1863: 
1864: \bibitem{nitta}J. Nitta, T.Koga, F. E. Meijer, Physica E {\bf 18}, 143
1865:                 (2003); F. E. Meijer, J. Nitta, T. Koga,
1866:                 A.F. Morpurgo, T. M. Klapwijk, Physica E {\bf 22},
1867:                 402, (2004); M.J. Yang, C.H. Yang, K.A. Cheng and
1868:                 Y.B. Lyanda-Geller, cond-mat/0208260.
1869: 
1870: \bibitem{yau} J. B. Yau, E. P. de Poortere, and M. Shayegan,
1871:               Phys. Rev. Lett.  {\bf 88}, 146801 (2002).
1872: 
1873: \bibitem{morpurgo} A. F. Morpurgo, J. P. Heida, T. M. Klapwijk,
1874:                    B. J. van Wees and G. Borghs, Phys. Rev. Lett. {\bf
1875:                    80}, 1050 (1998).
1876: 
1877: \bibitem{kato} Y.K. Kato, R.C. Meyers, A.C. Gossard, D.D. Awshalom
1878:                cond/mat 2005
1879: 
1880: \bibitem{molenkamp}M. K\"onig, A. Tschetschetkin, E.M. Henkiewicz,
1881:                    J. Sinova, V. Hock, V. Daumer, M. Sch\"afer,
1882:                    C.R. Becker, H. Buhmann and L.W. Molenkamp,
1883:                    Phys. Rev. Lett. {\bf 96} 0760804 (2006).
1884: 
1885: \bibitem{noiletter} R.Capozza, D.Giuliano, P. Lucignano, and A. Tagliacozzo,  
1886: Phys. Rev. Lett. {\bf 95}, 226803 (2005).
1887: 
1888: \bibitem{shayegan} B. Habib, E. Tutuc and M. Shayegan cond-mat/0612638.
1889: 
1890: \bibitem{nonloso} A review is given by C. W. J. Beenakker and H. van
1891: Houten, Solid State Phys., {\bf 44}, 1 (1991).
1892: 
1893: \bibitem{loss} D. Loss, P. Goldbart and A. V. Balatsky, Phys. Rev. Lett. 
1894: {\bf 65}, 1655 (1990); H. A. Engel and D. Loss, Phys. Rev. B {\bf 62},
1895: 10238 (2000).
1896: 
1897: \bibitem{feynman}R. P. Feynmann, \emph{Quantum Mechanics and Path Integral},
1898: Mc Graw-Hill, New York 1965.
1899: 
1900: \bibitem{tserkov} Y. Tserkovnyak and A. Brataas cond-mat/0611086
1901: 
1902: \bibitem{frustaglia} D. Frustaglia, K. Richter, Phys. Rev. B {\bf 69} 235310 
1903: (2004).
1904:     
1905: \bibitem{molnar} B. Moln\'ar, F. M. Peeters and P. Vasilopoulos,
1906:  Phys. Rev. B { \bf 69}, 155335 (2004); X.F.Wang and P. Vasilopoulos
1907:  Phys. Rev. B {\bf 72}, 165336 (2005).
1908: 
1909: \bibitem{shen} S.Q. Shen, Z.J. Li and  Z. Ma,  Appl. Phys. Lett. B
1910:                {\bf 84}, 996 (2004).
1911: \bibitem{dario} D.Becioux, D.Frustaglia and  M.Governale, Phys. Rev. B
1912:   {\bf 72}, 113310 (2005). 
1913: 
1914: \bibitem{citro} R. Citro, F. Romeo and M. Marinaro Phys. Rev. B {\bf 74},
1915: 115329 (2006).
1916: 
1917: \bibitem{souma}  S.Souma and B.K.Nikoli\'c, Phys.Rev.B{\bf 70},195346(2004).
1918: 
1919: \bibitem{lozano} G.S. Lozano, M.J. Sanchez, Phys. Rev. B {\bf 72},
1920:   205315 (2005).
1921: 
1922: \bibitem{wu} B.H. Wu and J.C. Chao, Phys. Rev. B {\bf 74}, 115313 (2006)
1923: 
1924: \bibitem{landauer} R. Landauer, IBM J. Res. Dev.{\bf 1} 223 (1957);
1925:                    M. Buttiker, IBM J. Res. Dev.{\bf 32} 317 (1988).
1926: 
1927: \bibitem{haldane}F. D. M. Haldane Phys. Rev. Lett. {\bf 50}, 1153
1928: (1983).
1929: 
1930: \bibitem{plet}M. Pletyukhov, Ch. Amann, M. Metha and M. Brack, Phys. Rev. Lett.
1931:         {\bf  89 }, 116601 (2002); O.Zaitsev, D. Frustaglia and
1932:         K. Richter, Phys. Rev. Lett. {\bf94}, 026809 (2005).
1933: 
1934: \bibitem{morette} C. Morette-de Witt, A. Maheshwari and B. Nelson,
1935:   Phys. Rep. {\bf 50}, 55 (1979).
1936: 
1937: \bibitem{nota} In the current notation $k_{SO}R=2m\alpha
1938:   R/\hbar=2\gamma$
1939: 
1940: \bibitem{moran} G. Morandi and E. Menossi, {Eur. J. Phys.} {\bf 5} 
1941: 49 (1984).
1942: 
1943: \bibitem{aurbach} A. Aurbach, F. Berruto and L. Capriotti, Field
1944:                   theory for low-dimensional systems Ed.s G. Morandi,
1945:                   P.Sodano, A.Tagliacozzo and V.Tognetti
1946:                   (Springer New York 2000).
1947: 
1948: \bibitem{datta} S. Datta and B. Das, Appl. Phys. Lett. {\bf 56}, 665
1949: (1990).
1950: \bibitem{koga} T.Koga, Y.Sekine and J.Nitta,  Phys.Rev.B{\bf 74}, 041302
1951: (2006).
1952: 
1953: \end{thebibliography}
1954: 
1955: \end{document}
1956: