cond-mat0304222/so.tex
1: \documentclass[aps,prb,twocolumn,draft,groupedaddress,showpacs]{revtex4}
2: 
3: \usepackage{amssymb}
4: \usepackage{amsfonts}
5: \usepackage{amsmath}
6: \usepackage{epsf}
7: 
8: \bibliographystyle{apsrev}
9: 
10: \begin{document}
11: 
12: \title{Weak localization and conductance fluctuations in a quantum dot with
13: parallel magnetic field and spin-orbit scattering}
14: \author{Jan-Hein Cremers}
15: \affiliation{Lyman Laboratory of Physics, Harvard University, 
16: Cambridge MA 02138}
17: 
18: \author{Piet W.\ Brouwer}
19: \affiliation{Laboratory of Atomic and Solid State Physics, Cornell 
20: University, Ithaca, NY 14853-2501}
21: 
22: \author{Vladimir I.\ Fal'ko}
23: \affiliation{Physics Department, Lancaster University, Lancaster L41
24:   4YB, United Kingdom}
25: 
26: \date{\today}
27: 
28: \begin{abstract}
29: In the presence of both spin-orbit scattering and a magnetic field the
30: conductance of a chaotic GaAs quantum dot displays quite a rich behavior.
31: Using a Hamiltonian derived by Aleiner and Fal'ko [Phys. Rev. Lett. 
32: \textbf{87}, 256801 (2001)] we calculate the weak localization
33: correction and the covariance of the conductance, as a function of
34: parallel and perpendicular magnetic field and spin-orbit coupling
35: strength. We also show how the combination of an in-plane 
36: magnetic field and spin-orbit scattering gives rise to a component
37: to the magnetoconductance that is anti-symmetric with respect to 
38: reversal of the perpendicular component of the magnetic field and
39: how spin-orbit scattering leads to a ``magnetic-field echo'' in the
40: conductance autocorrelation function. Our results can be used for
41: a measurement of the Dresselhaus and Bychkov-Rashba spin-orbit 
42: scattering lengths in a GaAs/GaAlAs heterostructure.
43: 
44: 
45: \end{abstract}
46: 
47: \pacs{73.23.-b, 05.45.Mt, 72.20.My, 73.63.Kv}
48: \maketitle
49: 
50: \section{Introduction}
51: 
52: A two-dimensional (2D) electron gas
53: offers ample opportunity for manipulation of the electron's spin and orbital
54: state. In the absence of spin-orbit coupling, such manipulation can
55: be performed separately on the electron spin and orbital degree of freedom,
56: by means of magnetic fields of various orientations and electrostatic gates.
57: On one hand, a magnetic field parallel to the plane of the 2D gas lifts the
58: spin degeneracy and allows for a measurement that distinguishes between the
59: transport properties of \textquotedblleft up\textquotedblright\ and
60: \textquotedblleft down\textquotedblright\ spins: As long as the confining
61: potential that determines the 2D electron gas is sharp, it does not
62: significantly affect the electron's orbital degrees of freedom. On the other
63: hand, gate voltages and a weak magnetic field perpendicular to the quantum
64: well act on the orbital degrees of freedom and do not couple significantly
65: to the electron's spin.\cite{mesoreview1}
66: 
67: Spin-orbit scattering couples the spin and orbital degrees of freedom.
68: Conventionally, the main signature of spin-orbit coupling in the 2D
69: electron gas in GaAs heterostructures is the observation of weak
70: anti-localization, a small positive quantum correction to the conductivity
71: that is suppressed by a magnetic field.\cite{Hikami} Weak anti-localization
72: is the counterpart of weak localization, a negative correction to
73: the conductivity at zero magnetic field caused by the constructive
74: interference of back-scattered electron waves in a phase-coherent disordered
75: conductor in the absence of spin-orbit scattering.\cite{Bergmann} Only
76: recently, spin-dependent phase coherent transport was considered for a
77: quantum dot, a small island of the electron gas confined by gates 
78: \cite{Khaetskii}. Motivated by an experiment by Folk 
79: \emph{et al.},\cite{Folk}
80: theoretical works by Halperin \emph{et al.}\cite{Halperin} and by two of the
81: authors,\cite{AF} have shown that the combination of spin-orbit scattering
82: and a spin degeneracy lifting parallel magnetic field leads to a remarkably
83: rich structure of quantum interference phenomena in III-V semiconductor 
84: quantum dots, that far surpasses a simple
85: interpolation between weak localization and weak antilocalization physics.
86: 
87: Not only does the crossover regime between weak localization and
88: anti-localization represent an interesting issue of its own,
89: understanding features imposed upon quantum transport by spin-orbit
90: coupling is needed for a correct extraction of the decoherence time
91: from experimental data. For electrons in a small dot with weak but
92: finite spin-orbit coupling, the the crossover of weak localization to
93: anti-localization manifests itself as a suppression of the
94: zero-temperature value of localization correction to the dot
95: conductance. This suppression of the weak localization correction may
96: be misinterpreted as a saturation of the decoherence time
97: $\tau_{\phi}$ at low temperatures. Without a priori knowledge of the
98: spin-orbit coupling constants for a given semiconductor dot, one must
99: use the features of spin-orbit coupling induced interference effects
100: in order to identify the origin of what can be mistaken for a suppression
101: of the interference part of conductance, when the observed weak
102: localization correction is less than the
103: prediction of the quantum transport theory for to the crossover
104: between orthogonal and unitary symmetry classes.
105: 
106: In this publication, we present a detailed quantitative analysis of
107: the influence of the coupling between electron spin and orbital motion
108: in a two-dimensional semiconductor on the interference corrections to
109: the conductance of a quantum dot. We consider the effects of
110: enhanced/suppressed back-scattering and the variance and correlation
111: properties of universal conductance fluctuations. The calculations,
112: which are performed in the framework of the random scattering matrix
113: approach of random matrix theory, are described in Sections
114: \ref{sec:2} and
115: \ref{sec:3}. In Section \ref{sec:2}, 
116: we present the analysis of the average conductance
117: of a chaotic dot, Section \ref{sec:3} is devoted to conductance
118: fluctuations. For technical reasons, we limit our attention to the
119: case of a large number of channels, $N_{1}$ and $N_{2}$ in the point
120: contact connecting the dot to the bulk, though all qualitative
121: features we discover for $N\gg 1$ would persist in the case of $N\sim
122: 1$. These two technical sections are preceded by a qualitative
123: discussion (Sec.\ \ref{sec:1}) and followed by an analysis of the 
124: effect of a
125: spatially non-uniform spin-orbit coupling strength (Sec.\ 
126: \ref{sec:nonuniform}).
127: 
128: \section{Interplay between SO coupling and Zeeman splitting in quantum
129:   dots} \label{sec:1}
130: 
131: In GaAs/AlGaAs heterostructures, spin-orbit coupling owes its
132: existence to the asymmetry of the potential confining the 2DEG and the lack
133: of inversion symmetry in the crystal structure, leading to the
134: Bychkov-Rashba and Dresselhaus terms in the Hamiltonian. In the theory
135: we present below, we consider electrons in a heterostructure or quantum well
136: lying in the (001) crystallographic plane of a zinc-blend type semiconductor
137: and choose coordinates $x_{1},x_{2}$ along crystallographic directions 
138: $\mathbf{\hat{e}}_{1}=[110]$ and $\mathbf{\hat{e}}_{2}=[1\bar{1}0]$
139: (coordinate $x_{3}$ is perpendicular to the plane of the two-dimensional
140: electron gas). The effective two-dimensional Hamiltonian of
141: electrons in a quantum dot takes the form 
142: \begin{eqnarray}
143:   \mathcal{H} &=&\frac{1}{2m}\left[ \left(
144:   p_{1}-eA_{1}-\frac{\sigma_{2}}{2\lambda_{1}}\right)^{2}+
145:   \left( p_{2}-eA_{2}+\frac{\sigma_{1}}{2\lambda_{2}}\right)^{2}
146:   \right]  \notag \\
147:   && \mbox{}+V(\mathbf{r})+\frac{1}{2}\mu_{B}g \mathbf{B} 
148:   \cdot \mbox{\boldmath $\sigma$}. \label{eq:FullHam}
149: \end{eqnarray}%
150: where $\mathbf{B}$ is the magnetic field, 
151: $\mathbf{A}=B_{3}(\mathbf{\hat{e}}_{3}\times \mathbf{r})/2c$ 
152: is the vector potential corresponding to the
153: component of the magnetic field perpendicular to the plane of the
154: two-dimensional electron gas, and $\mbox{\boldmath $\sigma$}=(\sigma_{1},\sigma
155: _{2},\sigma_{3})$ the vector of Pauli matrices. The potential $V$ both
156: confines the electrons to the quantum dot and describes elastic scattering
157: from non-magnetic impurities in the dot. The two length scales, 
158: $\lambda_{1}$ and $\lambda_{2}$ are associated with
159: spin-orbit coupling for an electron moving along the principal
160: crystallographic directions $\mathbf{\hat{e}}_{1}$ and 
161: $\mathbf{\hat{e}}_{2}$,\cite{RashbaDresselhaus} and characterize
162: the length at which spin of an initially polarized electron would
163: precess with an angle $2\pi $.
164: 
165: For a dot with homogeneous electron density and, therefore, parameters of
166: confining potential, spin-orbit coupling parameters $\lambda_{1}$ and 
167: $\lambda_{2}$ are independent of the position. In that case, the spin-orbit
168: coupling takes the form of a spin-dependent \textquotedblright vector
169: potential\textquotedblright , which is non-abelian since the Pauli matrices
170: do not commute with each 
171: other.\cite{Meir,MathurStone,Oreg} If $\lambda_{1}$ and 
172: $\lambda_{2}$ are large compared to the dot size $L$, the non-commutativity
173: of this spin-dependent vector potential involves higher powers of $L/\lambda
174: _{1,2}$, so that the spin-orbit coupling can be \textquotedblleft gauged
175: out\textquotedblright\ to leading order in $L/\lambda_{1,2}$. This
176: may be achieved through the unitary transformation $\psi (\mathbf{r})=
177: U(\mathbf{r})\tilde{\psi}(\mathbf{r})$, 
178: \begin{eqnarray}
179:   U &=&\exp \left( \frac{ix_{1}\sigma_{2}}{2\lambda_{1}}-\frac{ix_{2}\sigma
180: _{1}}{2\lambda_{2}}\right) 
181:   \nonumber \\ &=& \cos R+\left( \frac{ix_{1}\sigma_{2}}{2\lambda
182: _{1}}-\frac{ix_{2}\sigma_{1}}{2\lambda_{2}}\right) \dfrac{\sin R}{R},
183: \label{eq:U} 
184: \\
185:   R &=&\sqrt{\left( \frac{x_{1}}{2\lambda_{1}}\right) ^{2}+
186:   \left( \frac{x_{2}}{2\lambda_{2}}\right) ^{2}},
187: \end{eqnarray}%
188: which performs a position-dependent rotation of the spin out of the plane of
189: the two-dimensional electron gas to the locally adjusted
190: frame.\cite{AF} 
191: Below, we assume that $R\ll 1$, and use $R$ as a small
192: parameter to derive the form of the Hamiltonian $\tilde{\mathcal{H}}
193: =U^{\dagger }\mathcal{H}U$ in the locally rotated spin-frame, 
194: \begin{eqnarray}
195: \tilde{\mathcal{H}}&=&\frac{1}{2m}\left( -i\hbar \nabla \mathbf{-}e\mathbf{A}-
196: \mathbf{a}_{\bot }-\mathbf{a}_{\Vert }\right) ^{2}+h^{\mathrm{Z}}+h_{\bot }^{
197: \mathrm{Z}}+V(\mathbf{r}).
198:   \nonumber \\
199:   \label{eq:Hamtilde} \label{eq:hamiltonian}
200: \end{eqnarray}%
201: Here we introduced the spin-dependent vector potential
202: $$
203: \mathbf{a}_{\bot }=\frac{\sigma_{3}}{4\lambda_{1}\lambda_{2}}[\mathbf{
204: \hat{e}}_{3}\times \mathbf{r}]
205: $$
206: that has the same form as a magnetic field $\pm e c \mathbf{\hat{e}_3}
207: /2 \lambda_{1}\lambda_{2}$ with opposite directions for electrons with
208: spins ``up'' and ``down'' in the new
209: spin-frame, and we also abbreviated
210: \begin{eqnarray}
211:   \mathbf{a}_{\Vert } &=&\frac{1}{6\lambda_{1}\lambda_{2}}
212:   \left( \frac{x_{1}\sigma_{1}}{\lambda_{1}}
213:   + \frac{x_{2}\sigma_{2}}{\lambda_{2}}\right)
214:   [\mathbf{\hat{e}}_{3}\times \mathbf{r}]  \notag \\
215:   h^{\mathrm{Z}} &=&\frac{1}{2}\mu_{B}g\mathbf{B}\cdot 
216:   \mbox{\boldmath $\sigma$}, 
217:   \notag \\
218:   h_{\bot }^{\mathrm{Z}} &=&-\mu_{B}g
219:   \left( \frac{B_{1}x_{1}}{2\lambda_{1}}
220:   + \frac{B_{2}x_{2}}{2\lambda_{2}}\right) \sigma_{3}.
221: \end{eqnarray}%
222: 
223: The Hamiltonian (\ref{eq:Hamtilde})
224: describes electrons in a rotated spin frame. It
225: contains all relevant terms that lift the high degree of symmetry of a
226: system with uncoupled orbital and spin degrees of freedom, to leading
227: order in the small parameter $L/\lambda_{1,2}$. In these equations,
228: we omitted sub-leading terms of higher order in $L/\lambda_{1,2}$
229: that do not affect the symmetry of the Hamiltonian.
230: 
231: As long as the rate at which electrons escape from the quantum dot into the
232: leads is much smaller than the Thouless energy $E_{\mathrm{Th}}$, transport
233: properties of the quantum dot can be calculated using random matrix 
234: theory.\cite{Beenakker97}
235:  The use of random matrix theory requires an
236: analysis of the symmetries of the scattering matrix, which are set by the
237: relative magnitudes of characteristic energy scales for each of the terms in
238: the Hamiltonian (\ref{eq:hamiltonian}) and the escape rate. These energy
239: scales are 
240: \begin{eqnarray}
241: \varepsilon_{B} &=&\kappa E_{\mathrm{Th}}\left( {eB_{3}L^{2}}/2{\hbar }
242: \right) ^{2},  \notag \\
243: \varepsilon_{\bot }^{\mathrm{so}} &=&\kappa E_{\mathrm{Th}}\left(
244: 	    {L^{2}}/{4\lambda ^{2}}\right) ^{2},  \notag \\
245: \varepsilon_{\Vert }^{\mathrm{so}} &=&\kappa ^{\prime }\left( 
246: {L^{2}/4\lambda ^{2}}\right) \varepsilon_{\bot }^{\mathrm{so}},
247:   \label{eq:E_so} \label{eq:kappa}  \\
248: \varepsilon ^{\mathrm{Z}} &=&\mu_{B}gB,  \notag \\
249: \varepsilon_{\bot }^{\mathrm{Z}} &=&\frac{\kappa^{\prime \prime}
250: (\varepsilon^{\mathrm{Z}})^{2}}{E_{\mathrm{Th}}}\left( L^{2}/4\lambda
251: ^{2}\right) ,  \notag
252: \end{eqnarray}%
253: for the orbital contribution of the magnetic field, the spin-orbit
254: terms $a_{\Vert }$ and $a_{\bot }$, and the Zeeman coupling terms 
255: $h^{\mathrm{Z}}$ and $h_{\bot }^{\mathrm{Z}}$, respectively. 
256: For weak uniform spin-orbit coupling in a small dot such that 
257: $L/\lambda_{1,2}\ll 1$, one has the strong inequalities 
258: \begin{equation}
259: \varepsilon_{\Vert }^{\mathrm{so}}\ll \varepsilon_{\bot }^{\mathrm{so}},\
260: \ \varepsilon_{\bot }^{\mathrm{Z}}\ll \varepsilon ^{\mathrm{Z}}.
261: \label{eq:ineq}
262: \end{equation}
263: Here $E_{\rm Th}$ is the Thouless energy, which is the largest
264: energy scale in the problem, $\lambda ^{2}=\lambda
265: _{1}\lambda_{2}$, and $\kappa $, $\kappa ^{\prime }$, and $\kappa ^{\prime
266: \prime }$ are coefficients of order unity. The coefficients $\kappa ^{\prime
267: }$ and $\kappa ^{\prime \prime }$ depend on the sample geometry and on the
268: ratio $\lambda_{1}/\lambda_{2}$ of the spin-orbit lengths. The coefficient 
269: $\kappa ^{\prime \prime }$ also depends on the direction of the magnetic
270: field. The coefficient $\kappa $ depends on the sample geometry only; it
271: appears both for the orbital contribution of the magnetic field and for the
272: spin-orbit term $a_{\bot }$ because both terms have the same spatial
273: dependence. For a circular quantum dot with radius $L$, mean free path 
274: $l\ll L$, Fermi velocity $v_{F}$, and $\lambda_{1}=\lambda_{2}=\lambda $,
275: one has $E_{\mathrm{Th}}=\hbar v_{F}l/2L^{2}$, $\kappa =2$, $\kappa ^{\prime
276: }=1/3$ and $\kappa ^{\prime \prime }\approx 0.292$, see App.\ \ref{app:1}.
277: 
278: If the spin-orbit coupling is non-uniform, i.e., when the spin-orbit lengths 
279: $\lambda_{1}$ and $\lambda_{2}$ depend on position, the \textquotedblleft
280: spin-dependent vector potential\textquotedblright\ in Eq.\
281: (\ref{eq:FullHam}) can no longer be gauged away to leading order 
282: in $L/\lambda_{1,2}$. A
283: spatially non-uniform spin-orbit coupling can be created intentionally, with
284: the help of a metal gate parallel to the two-dimensional electron gas that
285: changes the asymmetry of the quantum well.\cite{Shayegan,Marcus,Zumbuehl},
286: or arise as a by-product of the confining gates. Generically, the
287: non-uniformity gives rise to a term in the Hamiltonian of the same symmetry
288: as $\mathbf{a}_{\Vert }$ in Eq.\ (\ref{eq:hamiltonian}).\cite{BCH} Hence,
289: the main effect of non-uniformities is to increase the corresponding energy
290: scale $\varepsilon_{\Vert }^{\mathrm{so}}$ by an amount of order 
291: $E_{\mathrm{Th}}(L/\lambda_{\mathrm{fl}})^{2}$, where 
292: $1/\lambda_{\mathrm{f}}$
293: is a measure of the fluctuations of the spin-orbit coupling $\lambda
294: _{1,2}^{-1}$.\cite{BCH} For very small quantum dots, this increase of 
295: $\varepsilon_{\Vert }^{\mathrm{so}}$ may eventually reverse the first
296: inequality in Eq.\ (\ref{eq:ineq}). 
297: 
298: It is the existence of two energy scales each to describe the strength of
299: the spin-orbit and Zeeman terms in the Hamiltonian that leads to the rich
300: parameter dependence of the conductance distribution of a chaotic quantum
301: dot. When all relevant energy scales are either much larger or much smaller
302: than the escape rate, the scattering matrix has a well defined symmetry, and
303: the conductance distribution can be found using symmetry considerations
304: alone.\cite{AF} 
305: It is the goal of this paper to address the general parameter regime, where
306: an interpolation between the various symmetry classes is called for. Our
307: results are important for a quantitative analysis of recent experiments by
308: Zumb\"{u}hl \emph{et al.},\cite{Zumbuehl} and for the development of
309: techniques for experimental determination of spin-orbit coupling 
310: parameters. Looking
311: into various regimes, we show how to extract such information from (a) weak
312: localization measurements in the presence of an in-plane magnetic field; (b)
313: an analysis of how the in-plane magnetic field gives rise to a
314: component of the magnetoconductance that is anti-symmetric
315: in the perpendicular magnetic field; and (c) from the possible 
316: observation of a \textquotedblleft
317: magnetic field echo\textquotedblright\ in the auto-correlation function of
318: conductance fluctuations.
319: 
320: The latter prediction is made on the basis of the following
321: theoretical observation. After the unitary transformation (\ref{eq:U}), the
322: leading spin-orbit term $\mathbf{a}_{\bot }^{\mathrm{so}}$ has precisely the
323: same spatial dependence as the vector potential due to the orbital
324: magnetic field (in a symmetric gauge). Hence, 
325: $\mathbf{a}_{\bot }^{\mathrm{so}}$ can be regarded as an effective 
326: magnetic field perpendicular
327: to the plane of the quantum well and of opposite sign for the two spin
328: directions. A suitably chosen orbital magnetic field can balance this
329: effective field, or change its sign for one spin direction, leading to a
330: partial re-appearance of weak localization and to a \textquotedblleft
331: magnetic field echo\textquotedblright in the auto-correlation
332: function of conductance fluctuations. The magnetic field echo
333: appears in the magnetoconductance autocorrelation function for
334: a magnetic field difference
335: \begin{equation}
336:   \Delta B_3 = \frac{\hbar}{e \lambda_1 \lambda_2};
337: \end{equation}
338: The shift of weak localization peaks is by half this field
339: strength.
340: 
341: According to the Onsager's relations, the two-terminal conductance of a
342: non-magnetic system is symmetric with respect to reversal of the
343: magnetic field $\mathbf{B}$. In the absence of spin-orbit coupling, 
344: the symmetry with respect to the reversal of the perpendicular
345: magnetic field component $B_{3}$ would persist even in the presence 
346: of Zeeman splitting caused in a 2D
347: electron gas by the in-plane magnetic field $\mathbf{B}_{\Vert}$
348: (excluding effects arising from the finite width of the quantum
349: well.\cite{FJ}) However,
350: in the presence of even a weak SO coupling, the exact symmetry exists
351: only if both $B_{3}$ and $B_{\Vert }$ are reversed simultaneously, 
352: whereas the reversal of the perpendicular component only generates an
353: antisymmetric contribution to the magnetoconductance with variance   
354: \begin{equation}
355: \mathrm{var}\left[ G(B_{3},B_{\Vert })-G(-B_{3},B_{\Vert })\right] \sim 
356: \dfrac{\left( \varepsilon ^{\mathrm{Z}}\right) ^{2}
357: \varepsilon_{\Vert }^{\mathrm{SO}}}{\left( N\Delta /2\pi \right)^{3}} 
358: \left( \dfrac{e^{2}}{h}\right) ^{2},  
359:   \label{eq:AsymmetricPart}
360: \end{equation}%
361: where $N$ is the number of channels in the leads, $\Delta $ is the
362: mean level spacing in the dot, and $\langle G \rangle$ is the
363: ensemble-averaged conductance of the quantum dot. Since the
364: fluctuating antisymmetric contribution the conductance,
365: $G(B_{3},B_{\Vert })- G(-B_{z},B_{\Vert })$ is linear in the in-plane
366: magnetic field, $\varepsilon ^{\mathrm{Z}}\approx g\mu_{B}B_{\Vert
367: }$, it dominates over the conductance asymmetry caused by
368: inter-subband mixing due to the in-plane magnetic field for
369: intermediate magnetic fields strengths,\cite{FJ} which is proportional
370: to $B_{\Vert }^{3}$.
371: 
372: \section{Average conductance}
373: 
374: \label{sec:2} 
375: 
376: The starting point of our calculation is the Landauer formula for the
377: two-terminal conductance $G$ of the quantum dot at zero temperature, 
378: \begin{equation}
379: G = {\frac{2e^{2}}{h}}{\frac{N_{1}N_{2}}{N}} - {\frac{e^{2}}{h}}
380: \mbox{tr}\,S\Lambda S^{\dagger }\Lambda,  \label{eq:Landauer}
381: \end{equation}
382: where the diagonal matrix $\Lambda $ has elements 
383: \begin{equation}
384: \Lambda_{jj}=\left\{ 
385: \begin{array}{ll}
386: N_{2}/N, & j=1,\ldots ,N_{1}, \\ 
387: -N_{1}/N, & j=N_{1}+1,\ldots ,N.
388: \end{array}
389: \right.
390: \end{equation}
391: As a matrix of complex numbers, the scattering matrix $S$ has dimension $2N
392: = 2(N_1 + N_2)$; it may also be seen as an $N \times N$ matrix of
393: quaternions, which are $2 \times 2$ matrices with special rules for complex
394: conjugation and transposition.\cite{Mehta}
395: 
396: \begin{figure}[t]
397: \epsfxsize=0.9\hsize 
398: \epsffile{fig1.eps} 
399: \caption{Illustration of the quantum dot with stub. Energy dependent
400: scattering from the quantum dot in the presence of spin-orbit coupling and a
401: magnetic field is modeled as scattering from a chaotic quantum dot with a
402: parameter independent scattering matrix $U$, with a stub described by a
403: reflection matrix $R$ which depends on energy, spin-orbit coupling and the
404: magnetic field. }
405: \label{fig:1a}
406: \end{figure}
407: 
408: In order to calculate the average of the conductance at Fermi energy 
409: $\varepsilon $ and magnetic field $\mathbf{B}$, or the covariance at energies 
410: $\varepsilon $ and $\varepsilon ^{\prime }$ and magnetic fields $\mathbf{B}$
411: and $\mathbf{B}^{\prime }$, it is sufficient to compute the average 
412: \begin{equation}
413: \langle S_{kl;\mu \nu }(\varepsilon ,\mathbf{B})S_{k^{\prime }l^{\prime
414: };\mu ^{\prime }\nu ^{\prime }}(\varepsilon ^{\prime },\mathbf{B}^{\prime
415: })^{\ast }\rangle ,  \label{eq:corr}
416: \end{equation}%
417: where roman indices refer to the propagating channels in the leads, and
418: Greek indices refer to spin. Hereto, we need a statistical description of
419: the scattering matrix $S$ in the presence of spin-orbit scattering and for
420: arbitrary values of the magnetic field $\mathbf{B}$ and Fermi energy 
421: $\varepsilon $. Within the \textquotedblleft random scattering matrix
422: approach\textquotedblright\ of random-matrix theory, this is provided by the
423: \textquotedblleft stub model\textquotedblright .\cite{WavesRM} In this
424: model, a fictitious \textquotedblleft stub\textquotedblright\ is attached to
425: the quantum dot, see Fig.\ \ref{fig:1a}. If the number of channels in the
426: stub is $M-N$, the scattering matrix $S$ can be written 
427: \begin{equation}
428: S=PU(1-Q^{\dagger }RQU)^{-1}P^{\dagger },  \label{eq:SU}
429: \end{equation}%
430: where $U$ is the $M\times M$ scattering matrix of the quantum dot (without
431: stub), $R$ is the $(M-N)$-dimensional reflection matrix of the stub, and $P$
432: and $Q$ are $N\times M$ and $(M-N)\times M$ projection matrices,
433: respectively, with $P_{ij}=\delta_{i,j}\openone$, and $Q_{ij}=\delta
434: _{i+N,j}\openone$. (Here $\openone$ is the $2\times 2$ unit matrix in spin
435: space.) The area and width of the stub are chosen such that (i) the dwell
436: time in the stub is much larger than the dwell time in the quantum dot and
437: (ii) the time for ergodic exploration of the dot plus stub system is much
438: shorter than the time for escape into one of the two leads. The first
439: condition ensures that all dependence on Fermi energy, magnetic field, or
440: spin-orbit scattering rate will be through the reflection matrix $R$, so
441: that the matrix $U$ can be taken at a fixed reference energy $\varepsilon =0$
442: and at zero magnetic field and spin-orbit scattering rate. Since scattering
443: from the quantum dot is chaotic, this implies that elements of $U$ are
444: proportional to the $2\times 2$ unit matrix in spin space $\openone$ and
445: that $U$ can be chosen from Dyson's circular orthogonal ensemble of random
446: matrix theory.\cite{Beenakker97} The second condition, which requires $M\gg
447: N $, ensures that electrons explore the phase space of the combined dot plus
448: stub system ergodically before they exit to the leads, so that the
449: distribution of the scattering matrix $S$ remains universal and is
450: unaffected by the addition of the stub. Indeed, the parametric dependence of
451: the scattering matrix described by the stub model has been shown to be the
452: same as that described by the Hamiltonian approach of random-matrix 
453: theory.\cite{WavesRM}
454: 
455: We take the reflection matrix $R$ of the form 
456: \begin{equation}
457: R=\mbox{exp}\left[ \frac{2\pi i}{M\Delta }(\varepsilon -
458: \mathcal{{H^{\prime }})}\right] ,  \label{eq:R}
459: \end{equation}%
460: where $\Delta $ is the mean level spacing of the closed dot and 
461: $\mathcal{{H^{\prime }}}$ is an $(M-N)$ dimensional matrix which 
462: describes the effect
463: of the magnetic field and spin-orbit scattering, 
464: \begin{eqnarray}
465: {\mathcal{H}^{\prime }} &=&\frac{\Delta }{2\pi }
466:   \left[i \mathcal{X}(x\openone+a_{\bot }\sigma_{3})+
467:   ia\left( \mathcal{A}_{1}\sigma_{1}+\mathcal{A}_{2}\sigma_{2}\right) 
468:   \right.  \notag \\
469: &&\left. \mbox{}-\mathbf{b}\cdot \mbox{\boldmath $\sigma$}+b_{\bot }
470:   \mathcal{B}_{h}\sigma_{3})\right] .
471: \end{eqnarray}%
472: Here $\mathcal{X}$, $\mathcal{A}_{1}$ and $\mathcal{A}_{2}$ are real
473: antisymmetric matrices, with $\left\langle \mbox{tr}\,
474: \mathcal{XX}^{T}\right\rangle =M^{2}$ and $\left\langle \mbox{tr}\,
475: \mathcal{A}_{i}
476: \mathcal{A}_{j}^{T}\right\rangle =\delta_{ij}M^{2}$, while $\mathcal{B}_{h}$
477: is a real symmetric matrix with $\left\langle \mbox{tr}\,\mathcal{B}
478: _{h}^{2}\right\rangle =M^{2}$. The dimensionless parameters $x$, $a_{\bot }$
479: , $a$, $b$, and $b_{\bot }$ correspond to the energy scales defined in
480: Eq.\ (\ref{eq:E_so}) as 
481: \begin{eqnarray}
482: x^{2} &=&\pi \varepsilon_{B}/\Delta ,  \notag \\
483: a_{\bot }^{2} &=&\pi \varepsilon_{\bot }^{\mbox{so}}/\Delta ,  \notag \\
484: a^{2} &=&\pi \varepsilon_{\Vert }^{\mbox{so}}/\Delta ,  \label{eq:defs} \\
485: b &=&\pi \varepsilon ^{\mathrm{Z}}/\Delta ,  \notag \\
486: b_{\bot }^{2} &=&\pi \varepsilon_{\bot }^{\mathrm{Z}}/\Delta .  \notag
487: \end{eqnarray}
488: We define the corresponding quantities $x^{\prime }$, $b^{\prime }$,
489: and $b_{\bot }^{\prime }$ when calculating a correlator between 
490: scattering matrix
491: elements at different values $\mathbf{B}$ and $\mathbf{B}^{\prime }$ of the
492: magnetic field. For weak uniform spin-orbit scattering, one has the
493: strong inequalities $a \ll a_{\bot}$, $b_{\bot} \ll b$.
494: 
495: To calculate the correlator (\ref{eq:corr}), we expand $S$ in powers of $U$
496: using Eq.~(\ref{eq:SU}) and integrate $U$ over the space of unitary
497: symmetric matrices using the diagrammatic technique of Ref.\
498: \onlinecite{diagrams}. To leading order in $1/M$ and $1/N$ we can take the
499: elements of $U$ to be Gaussian random variables with mean zero and with
500: variance $\left\langle U_{ij}U_{kl}^{\ast }\right\rangle =M^{-1}\left(
501: \delta_{ik}\delta_{jl}+\delta_{il}\delta_{jk}\right) $. Then only ladder
502: diagrams and maximally crossed diagrams contribute to the correlator
503: Eq.\ 
504: (\ref{eq:corr}). Summing the contributions from these two classes of
505: diagrams yields 
506: \begin{eqnarray}
507: \langle S_{kl;\mu \nu }^{\vphantom{*}}(\varepsilon ,\mathbf{B})S_{k^{\prime
508: }l^{\prime };\mu ^{\prime }\nu ^{\prime }}^{\ast }(\varepsilon ^{\prime },
509: \mathbf{B}^{\prime })\rangle \! &=&\delta_{kk^{\prime }}\delta_{ll^{\prime
510: }}D_{\mu \nu ;\nu ^{\prime }\mu ^{\prime }} \\
511: &&\!\mbox{}+\delta_{kl^{\prime }}\delta_{lk^{\prime }}(\mathcal{T}
512: C \mathcal{T})_{\mu \nu ;\mu ^{\prime }\nu ^{\prime }},  \notag
513: \end{eqnarray}%
514: where 
515: \begin{eqnarray}
516: D &=&\left( M\openone\otimes \openone-\mbox{tr}\,R\otimes R^{\prime
517: }{}^{\dagger }\right) ^{-1},  \notag \\
518: C &=&\left( M\openone\otimes \openone-\mbox{tr}\,R\otimes R^{\prime
519: }{}^{\ast }\right) ^{-1}.  \label{eq:CD_defs}
520: \end{eqnarray}%
521: Here $R^{\prime }$ is given by Eq.\ (\ref{eq:R}), with $x$, $b$, $b_{\bot }$
522: and $\varepsilon $ replaced by $x^{\prime }$, $b^{\prime }$, $b_{\bot
523: }^{\prime }$ and $\varepsilon ^{\prime }$, respectively. The superscript $
524: \ast $ denotes the quaternion complex conjugate of $R^{\prime }$, 
525: $\mathcal{T}=\openone\otimes \sigma_{2}$, and the trace is taken 
526: over the channel
527: indices only. The tensor multiplication in Eq.\ (\ref{eq:CD_defs}) has a
528: reverse-order multiplication for the Pauli matrices in second place, 
529: \begin{equation}
530: (\sigma_{i}\otimes \sigma_{j})(\sigma_{i^{\prime }}\otimes \sigma
531: _{j^{\prime }})=(\sigma_{i}\sigma_{i^{\prime }})\otimes (\sigma
532: _{j^{\prime }}\sigma_{j}).  \label{eq:mult}
533: \end{equation}%
534: The two contributions $C$ and $D$ are the equivalents of cooperon and
535: diffuson in the conventional diagrammatic perturbation 
536: theory.\cite{SimonsAltshuler} Taking the limit $M\rightarrow \infty $ 
537: and using Eq.\ (\ref{eq:R}) we find 
538: \begin{eqnarray}
539: D^{-1} &=&N\openone\otimes \openone+\frac{2\pi i(\varepsilon -\varepsilon
540: ^{\prime })}{\Delta }\openone\otimes \openone  \notag  \label{eq:D} \\
541: &&\mbox{}+\frac{1}{2}[(x-x^{\prime })\openone\otimes \openone+a_{\bot
542: }(\sigma_{3}\otimes \openone-\openone\otimes \sigma_{3})]^{2}  \notag \\
543: &&\mbox{}+\frac{1}{2}a^{2}(\sigma_{1}\otimes \openone-\openone\otimes
544: \sigma_{1})^{2}  \notag \\
545: &&\mbox{}+\frac{1}{2}a^{2}(\sigma_{2}\otimes \openone-\openone\otimes
546: \sigma_{2})^{2}  \notag \\
547: &&\mbox{}+i\mathbf{b}\cdot (\mbox{\boldmath $\sigma$}\otimes \openone-
548: \openone\otimes \mbox{\boldmath $\sigma$})  \notag \\
549: &&\mbox{}+\frac{1}{2}(h\sigma_{3}\otimes \openone-h^{\prime }\openone
550: \otimes \sigma_{3})^{2}.
551: \end{eqnarray}%
552: \begin{eqnarray}
553: C^{-1} &=&N\openone\otimes \openone+\frac{2\pi i(\varepsilon -\varepsilon
554: ^{\prime })}{\Delta }\openone\otimes \openone  \notag  \label{eq:C} \\
555: &&\mbox{}+\frac{1}{2}[(x+x^{\prime })\openone\otimes \openone+a_{\bot
556: }(\sigma_{3}\otimes \openone-\openone\otimes \sigma_{3})]^{2}  \notag \\
557: &&\mbox{}+\frac{1}{2}a^{2}(\sigma_{1}\otimes \openone-\openone\otimes
558: \sigma_{1})^{2}  \notag \\
559: &&\mbox{}+\frac{1}{2}a^{2}(\sigma_{2}\otimes \openone-\openone\otimes
560: \sigma_{2})^{2}  \notag \\
561: &&\mbox{}+i\mathbf{b}\cdot (\mbox{\boldmath $\sigma$}\otimes \openone+
562: \openone\otimes \mbox{\boldmath $\sigma$})  \notag \\
563: &&\mbox{}+\frac{1}{2}(h\sigma_{3}\otimes \openone+h^{\prime }\openone
564: \otimes \sigma_{3})^{2}.
565: \end{eqnarray}
566: 
567: \begin{figure}[t]
568: \epsfxsize=0.9\hsize 
569: \epsffile{fig2.eps} 
570: \caption{Dependence of the quantum-interference correction to the
571: conductance, $\protect\delta G$, on the perpendicular magnetic field $x$.
572: The conductance is measured in units of $(2e^2/h)N_{1}N_{2}/(N_1+N_2)^2$.
573: Panel (a) shows $\protect\delta G$ for $a_{\bot}=3$, $a=0$, $b_{\bot}=0$ and
574: two values of the parallel magnetic field, $b=0$ and $b=3$, and for the
575: limit $b \to \infty$. Panel (b) shows $\protect\delta G$ for
576: $a_{\bot}=3$, 
577: $b=b_{\bot}=0$, and three values of the spin-orbit parameter $a$:
578: $a=0$, 
579: $a=0.5$ and $a\to \infty$. Panel (c) shows $\protect\delta G$ in the
580: crossover regime, with $a_{\bot}=3.0$, $a=0.5$, and for the same values of
581: the parallel magnetic field as in (a). In all three plots, we have set 
582: $b_{\bot} = 0$. Choosing $b_{\bot} \ll b^2$ is appropriate for small quantum
583: dots with $L \ll \protect\lambda_{1,2}$. A small nonzero value of $b_{\bot}$
584: has only a slight effect on the shape of the curves, its general effect
585: being to further reduce weak antilocalization. }
586: \label{fig:1}
587: \end{figure}
588: 
589: Setting $x^{\prime}=x$, $\mathbf{b}=\mathbf{b}^{\prime}=b \hat {e_1}$
590: and 
591: $b_{\bot}^{\prime}=b_{\bot}$, we find the average conductance, 
592: \begin{eqnarray}
593: \left\langle G\right\rangle &=&\frac{2e^{2}}{h}\frac{N_{1}N_{2}}{N_1+N_2} +
594: \delta G, \\
595: \delta G &=& -\frac{e^2}{h} \frac{N_1 N_2}{N_1+N_2} \sum_{\mu ,\nu }\left( 
596: \mathcal{T}C\mathcal{T}\right)_{\mu \nu ;\mu \nu}  \notag \\
597: &=& -\frac{e^2}{h} \frac{N_1 N_2}{N_1+N_2} \left( \frac{1}{4 a^2 + F_C}
598: \right.  \notag \\
599: && \left. \mbox{} + \frac{4b^{2}+G_C^2+2G_{C}F_{C}-16 a_{\bot}^2 x^2}{
600: 4G_{C}b^{2}+G_{C}^{2}F_{C}-16a_{\bot}^{2}F_{C}x^{2}} \right),
601: \label{eq:weak}
602: \end{eqnarray}
603: where we abbreviated 
604: \begin{eqnarray}
605: F_{C} &=&N+2\left(x^{2}+b_{\bot}^{2}\right)  \label{eq:fc} \\
606: G_{C} &=&N+2\left(x^{2}+ a^{2}+a_{\bot}^{2}\right).  \notag
607: \end{eqnarray}
608: The expression for the average conductance simplifies considerably when
609: either the perpendicular magnetic field or the parallel magnetic field is
610: zero. In the first case, $b=b_{\bot}=0$, one finds 
611: \begin{eqnarray}
612: \delta G &=& \frac{e^2}{h} \frac{N_1 N_2}{N_1+N_2} \left[\frac{1}{N + 2x^2}
613: - \frac{1}{4a^2 + N + 2x^2}\right.  \notag \\
614: && \left.\mbox{} - \sum_{\pm} \frac{1}{2a^2 + N + 2(a_{\bot} \pm x)^2}
615: \right].
616: \end{eqnarray}
617: This simple result can be understood recalling that the spin-orbit coupling
618: term $a_{\bot}$ acts as an effective perpendicular field with opposite sign
619: for up and down spins. When the applied perpendicular field $x$ exactly
620: cancels this effective field, i.e., when $x=\pm a_{\bot}$, time-reversal
621: symmetry is ``restored'' for one spin direction, leading to a weak
622: localization-like correction to the conductance centered at $x = \pm
623: a_{\bot} $. A non-zero value of the spin-orbit coupling term $a$ leads to a
624: weak antilocalization peak at $x=0$, while a finite parallel magnetic fields
625: causes a weak-localization peak at zero perpendicular field, as is
626: illustrated in Fig.~\ref{fig:1}b and a, respectively. Both a parallel
627: magnetic field and the spin-orbit coupling term $a$ suppress the features at 
628: $x = \pm a_{\bot}$. (The case $a\gg a_{\bot}$ is applicable when the
629: spin-orbit coupling is not uniform, as discussed in the introduction.)
630: Figure \ref{fig:1}c shows an example for the dependence of the
631: conductance $G $ on the perpendicular field in the crossover regime
632: when all parameters 
633: ($a_{\bot}$, $a$, and $b$) are important.
634: 
635: In the special case when the perpendicular magnetic field, $x$, is zero, we
636: find 
637: \begin{eqnarray}
638: \delta G &=& -\frac{e^2}{h} \frac{N_1 N_2}{N_1+N_2} \left[ \frac{1}{N + 2
639: a^2 + 2a_{\bot}^2} \right.  \notag \\
640: && \left. \mbox{} - \frac{2 a^2 + 2a_{\bot}^2 - b_{\bot}^2} {4 b^2 + (N + 2
641: b_{\bot}^2)(N + 2 a^2 + 2a_{\bot}^2)} \right.  \notag \\
642: && \left. \mbox{} + \frac{1}{N + 4 a^2 + 2 b_{\bot}^2} \right].
643: \end{eqnarray}
644: In the case of uniform spin-orbit scattering, this may be further simplified
645: using $a^{2}\ll a_{\bot}^2$ and $b_{\bot}^2\ll b$,\cite{AF} 
646: \begin{eqnarray}
647: \delta G &=& -\frac{e^2}{h} \frac{N_1 N_2}{N_1+N_2} \left[ \frac{1}{2
648: a_{\bot}^2 + N}\right.  \label{eq:G_x0} \\
649: &&\left. \mbox{} - \frac{1}{4 a^2 + 2 b_{\bot}^2 + N} +\frac{a_{\bot}^2}{4
650: b^2 + N^2 + 2 N a_{\bot}^2}\right].  \notag
651: \end{eqnarray}
652: 
653: At finite temperatures, dephasing will lead to a further suppression of weak
654: localization. Dephasing can be included in the theory presented here by the
655: introduction of a fictitious voltage probe.\cite{BuettikerVoltage,BM,BB}
656: This amounts to the replacement $N\rightarrow N+\hbar /\tau_{\phi }\Delta $
657: in the final results (\ref{eq:weak})--(\ref{eq:G_x0}), where $\tau_{\phi }$
658: is the dephasing time and $\Delta $ is the mean level spacing in the closed
659: quantum dot.\cite{BCH}
660: 
661: \section{Conductance fluctuations}
662: 
663: \label{sec:3}
664: 
665: Conductance fluctuations are described by the covariance of the conductance
666: at two different values $\varepsilon $ and $\varepsilon ^{\prime }$ of the
667: Fermi energy and at two different magnetic fields $\mathbf{B}$ and 
668: $\mathbf{B}^{\prime }$, 
669: \begin{eqnarray}
670:   \mbox{cov}\,[G(\varepsilon ,\mathbf{B}),
671:   G(\varepsilon ^{\prime },\mathbf{B}^{\prime })] &=&
672:   \langle G(\varepsilon ,\mathbf{B}),G(\varepsilon ^{\prime },
673:   \mathbf{B}^{\prime })\rangle  \notag  \label{eq:cov_def} \\
674:   &&\mbox{}-\langle G(\varepsilon ,\mathbf{B})\rangle \langle 
675:   G(\varepsilon^{\prime },\mathbf{B}^{\prime })\rangle .  \notag \\
676: &&
677: \end{eqnarray}%
678: For this calculation, we need to know the average of a product of four
679: scattering matrix elements. However, if the conductance is expressed in
680: terms of the scattering matrix $S$ using Eq.\ (\ref{eq:Landauer}), to
681: leading order in $1/N$, the scattering matrix elements may be considered
682: Gaussian random numbers, and the average of four scattering matrix elements
683: can be factorized into products of pair averages of the form of Eq.\ 
684: (\ref{eq:corr}).\cite{Argaman} One thus obtains 
685: \begin{eqnarray}
686: \mbox{cov}\,\left[ G\left( \varepsilon ,\mathbf{B}\right) ,G\left(
687: \varepsilon ^{\prime },\mathbf{B}^{\prime }\right) \right] &=&\left( 
688: \frac{e^{2}}{h}\right) ^{2}\left( \frac{N_{1}N_{2}}{N_{1}+N_{2}}\right)^{2} 
689: \notag \\
690: &&\mbox{}\times \left( V_{D}+V_{C}\right) ,  \label{eq:covresult}
691: \end{eqnarray}
692: where 
693: \begin{widetext}
694: \begin{eqnarray} 
695:   \label{eq:D_def}
696:   V_D &=& \sum_{\mu,\nu,\mu',\nu' =\pm}
697:   D_{\mu \nu ;\mu' \nu'}D_{\nu' \mu';\nu \mu }
698:   \\
699:   &=& 
700:   \frac{\Xi_{D}}{\left| (b^{2}-b^{\prime
701: }{}^{2})^{2}+2K_{D}(-4a^{2}bb^{\prime }+(b^{2}+b^{\prime
702: }{}^{2})L_{D})+(L_{D}^{2}-4a^{4})(K_{D}^{2}-4a_{\bot}^{2}(x-x^{\prime
703: })^{2})\right| ^{2}}, \nonumber \\
704:   \label{eq:C_def}
705:   V_C &=& \sum_{\mu,\nu,\mu',\nu' =\pm}
706:   \left( {\cal T}C{\cal T}\right)_{\mu \nu ;\mu' \nu'}
707:   \left( {\cal T}C{\cal T}\right)_{\mu' \nu' ;\mu \nu}
708:   \\ &=& \nonumber
709:   \frac{\Xi_{C}}{\left| (b^{2}-b^{\prime
710: }{}^{2})^{2}+2K_{C}(4a^{2}bb^{\prime }+(b^{2}+b^{\prime
711: }{}^{2})L_{C})+(L_{C}^{2}-4a^{4})(K_{C}^{2}-4a_{\bot}^{2}(x+x^{\prime
712: })^{2})\right| ^{2}},
713: \end{eqnarray}
714: Here we abbreviated 
715: \begin{eqnarray*}
716: \Xi_{D} &=& 
717:   2\left| K_{D}(4a^{4}-L_{D}^{2})+
718:   4a^{2}bb^{\prime }-L_{D}(b^{2}+b^{\prime}{}^{2})\right| ^{2}
719: %  \nonumber \\ && \mbox{}
720:   + 2\left| K_{D}(b^{2}+b^{\prime}{}^{2})+K_{D}^{2}L_{D}-
721:   4a_{\bot}^{2}L_{D}(x-x^{\prime })^{2}\right|^{2}\\
722:   && \mbox{} + 
723:   8a_{\bot}^{2}(x-x^{\prime })^{2}\left(\left(b+b^{\prime }\right)^2+
724:   \left|L_{D}+2a^2\right|^2\right)\left(\left(b-b^{\prime }\right)^2+
725:   \left|L_{D}-2a^2\right|^2\right)
726: %  \nonumber \\ && \mbox{}
727:   +8\left| K_{D}bb^{\prime }+a^{2}(K_{D}^{2}-
728:   4a_{\bot}^{2}(x-x^{\prime})^{2})\right| ^{2}
729:   \\ && \mbox{} 
730:   + 4\left| b(b^{2}-b^{\prime }{}^{2})-K_{D}(2a^{2}b^{\prime}- 
731:   bL_{D})\right|^{2}+4\left| b^{\prime }(b^{\prime }{}^{2}-b^{2})-
732:   K_{D}(2a^{2}b-b^{\prime}L_{D})\right| ^{2} 
733: %  \nonumber \\ && \mbox{}
734:   + 8\left| a^{2}(b^{2}+b^{\prime }{}^{2})-
735:   bb^{\prime }L_{D}\right|^{2} \\
736:   \Xi_{C} &=& 2\left| K_{C}(4a^{4}-L_{C}^{2})-
737:   4a^{2}bb^{\prime }-L_{C}(b^{2}+b^{\prime}{}^{2})\right| ^{2}
738: %  \nonumber \\ && \mbox{}
739:   + 2\left| K_{C}(b^{2}+b^{\prime}{}^{2})+K_{C}^{2}L_{C}-
740:   4a_{\bot}^{2}L_{C}(x+x^{\prime })^{2}\right|^{2}\\
741:   && \mbox{} 
742:   + 8a_{\bot}^{2}(x+x^{\prime })^{2}\left(\left(b-b^{\prime }\right)^2+
743:   \left|L_{C}+2a^2\right|^2\right)\left(\left(b+b^{\prime }\right)^2+
744:   \left|L_{C}-2a^2\right|^2\right)
745: %  \nonumber \\ && \mbox{}
746:   + 8\left| K_{C}bb^{\prime }+a^{2}(-K_{C}^{2}+
747:   4a_{\bot}^{2}(x+x^{\prime})^{2})\right| ^{2}
748:   \\ && \mbox{} 
749:   + 4\left| b(b^{2}-b^{\prime }{}^{2})+
750:   K_{C}(2a^{2}b^{\prime} + bL_{C})\right|^{2}
751:   +4\left| b^{\prime }(b^{\prime }{}^{2}-b^{2})+
752:   K_{C}(2a^{2}b+b^{\prime}L_{C})\right| ^{2}
753: %  \nonumber \\ && \mbox{}
754:   + 8\left| a^{2}(b^{2}+b^{\prime }{}^{2})+
755: bb^{\prime }L_{C}\right|^{2}
756: \end{eqnarray*}
757: and 
758: \begin{eqnarray*}
759:   K_{D} &=& N + \frac{2\pi i ( \varepsilon -\varepsilon')}{\Delta} 
760:   + \frac{1}{2}\left(x-x^{\prime}\right)^2+
761: 2\left( a^{2}+a_{\bot}^{2}\right)+\frac{1}{2}\left(b_{\bot}+
762: b_{\bot}^{\prime}\right)^2, \\
763:   L_{D} &=& N + \frac{2\pi i ( \varepsilon -\varepsilon')}{\Delta} 
764:   +\frac{1}{2}\left(x-x^{\prime}\right)^2+
765: 2a^{2}+\frac{1}{2}\left(b_{\bot}-b_{\bot}^{\prime}\right)^2, \\
766:   K_{C} &=& N + \frac{2\pi i ( \varepsilon -\varepsilon')}{\Delta} 
767:   +\frac{1}{2}\left(x+x^{\prime}\right)^2+
768:   2\left( a^{2}+a_{\bot}^{2}\right)+\frac{1}{2}\left(b_{\bot}-
769:   b_{\bot}^{\prime}\right)^2, \\
770:   L_{C} &=& N + \frac{2\pi i ( \varepsilon -\varepsilon')}{\Delta} 
771:   +\frac{1}{2}\left(x+x^{\prime}\right)^2+
772:   2a^{2}+\frac{1}{2}\left(b_{\bot}+b_{\bot}^{\prime}\right)^2.
773: \end{eqnarray*}
774: 
775: 
776: \subsection{Variance}
777: 
778: The variance of the conductance at zero temperature is obtained from
779: Eq.\ (\ref{eq:covresult}) by setting $\varepsilon' = \varepsilon$ and
780: $\mathbf{B'} = \mathbf{B}$,
781: \begin{eqnarray}
782:   \mbox{var}\, G &=&
783:   \frac{e^{4}}{h^{2}} \frac{N_{1}^{2}N_{2}^{2}}{(N_1+N_2)^2}
784:   \left[\frac{1}{N^2} + \frac{1}{G_D^2} +
785:   \frac{8 b^2 + G_D^2 + (4 a^2 + N)^2}{(4 a^2 G_D + 4 b^2 + G_D N)^2}
786:   + \frac{1}{(4 a^2 + F_C)^2}
787:   \right. \nonumber \\ && \left. \mbox{}
788:   + \frac{2 F_C}{G_C (4 b^2 + G_C F_C) - 16 a_{\bot}^2 F_C x^2}
789: %  \right. \nonumber \\ && \left. \mbox{}
790:   + \frac{(4 b^2 + G_C^2 + 16 a_{\bot}^2 x^2)^2
791:     + 64 a_{\bot}^2 x^2 (F_C^2 - G_C^2)}
792:   {(G_C (4 b^2 + G_C F_C) - 16 a_{\bot}^2 F_C x^2)^2} \right].
793:   \label{eq:genvar}
794: \end{eqnarray}
795: where $F_{C}$ and $G_{C}$ are given by Eq.\ (\ref{eq:fc}), and
796: \begin{equation}  
797:   G_{D} =N+2\left( a^{2}+a_{\bot}^{2}+b_{\bot}^{2}\right).
798: \end{equation}
799: Simplifications occur in the limits of zero parallel or
800: perpendicular field, and for large perpendicular field. In the
801: absence of a parallel field, $b=b_{\bot}=0$, one finds the 
802: variance
803: \begin{eqnarray}
804: \mbox{var}\,  G &=& \frac{e^{4}}{h^{2}} \frac{N_{1}^{2}N_{2}^{2}}
805:   {(N_1 + N_2)^{2}}
806: \left[\frac{1}{(2 a^2 + N + 2 (a_{\bot} - x)^2)^2} + 
807: \frac{1}{(N + 2 x^2)^2} + \frac{1}{(4 a^2 + N + 2 x^2)^2}
808: \right.\nonumber \\ && \left. \mbox{}
809:   + \frac{1}{(2 a^2 + N + 2 (a_{\bot} + x)^2)^2}
810:   + \frac{1}{N^2} + \frac{1}{(4 a^2 + N)^2} + \frac{2}{(2 (a^2 +
811: a_{\bot}^2) + N)^2}\right],\label{eq:var_b0}
812: \end{eqnarray}
813: while in the absence of a perpendicular magnetic field, $x=0$, one has
814: \begin{eqnarray}
815: \mbox{var}\, G &=&\frac{e^{4}}{h^{2}} \frac{N_{1}^{2}N_{2}^{2}}
816:   {(N_1 + N_2)^{2}}\left[
817:   \frac{1}{N^2} + \frac{1}{(4 a^2 + 2 b_{\bot}^2 + N)^2} + \frac{2}{[2
818:   (a^2 + a_{\bot}^2 + b_{\bot}^2) + N]^2} 
819: %  \nonumber \right. \\ && \left. \mbox{}
820:   + \frac{2}{4 b^2 + (4 a^2 + N) [2 (a^2 + a_{\bot}^2) + N]}
821:   \nonumber \right. \\ && \left. \mbox{}
822:    + \frac{2}{4 b^2 + N (2 (a^2 + a_{\bot}^2) + N)} + 
823:   \frac{4 (a^2 + a_{\bot}^2)^2}
824:   {\left(4 b^2 + N [2 (a^2 + a_{\bot}^2) + N]\right)^2} 
825: %  \nonumber \right. \\ && \left. \mbox{}
826:   + \frac{4 (a^2 - a_{\bot}^2)^2}
827:   {\left(4 b^2
828: + (4 a^2 + N) [2 (a^2 + a_{\bot}^2) + N]\right)^2}\right].
829: \label{eq:var_x0}
830: \end{eqnarray}
831: In the last equality we also used $b_{\bot} \ll b^2$. 
832: In the limit of large perpendicular field, $x \gg 1$, one has
833: \begin{eqnarray}
834:   \mbox{var}\, G &=&\frac{e^{4}}{h^{2}} \frac{N_{1}^{2}N_{2}^{2}}
835:   {(N_1 + N_2)^{2}}\left[\frac{1}{N^2}+ 
836:   \frac{2}{4b^2 + (4a^2 + N)\left[2(a^2 + a_{\bot}^2 + b_{\bot}^2) + N\right]}
837:   \nonumber \right.\\
838:   && \left. \mbox{} +
839:   \frac{1}{\left[2(a^2 + a_{\bot}^2 + b_{\bot}^2) + N\right]^{2}}
840:   + \frac{4(-a^2 + a_{\bot}^2 + b_{\bot}^2)^2}
841:   {\left(4b^2 + (4a^2 + N)\left[2(a^2 + a_{\bot}^2 + b_{\bot}^2) +
842:   N\right]\right)^2}
843:   \right].
844: \end{eqnarray}
845: When
846: both the perpendicular and the parallel field are large, i.e. 
847: $x^{2},b\gg a_{\bot}^{2},a^{2},b_{\bot}^{2},N$ a particularly simple 
848: expression for the variance results 
849: \begin{eqnarray}
850: \nonumber
851: \mbox{var}\, G &=& \frac{e^{4}}{h^{2}}
852: \frac{N_{1}^{2}N_{2}^{2}}{(N_1+N_2)^{2}}\left[
853: \frac{1}{N^2} 
854:   + \frac{1}{\left(N + 2a^2 + 2a_{\bot}^2 
855:   + 2b_{\bot}^2\right)^{2}}\right].\label{eq:var_large}
856: \end{eqnarray}
857: \end{widetext}
858: 
859: \begin{figure}[t]
860: \epsfxsize=0.9\hsize 
861: \epsffile{fig3.eps} 
862: \caption{Conductance fluctuations $\mbox{var}\, G$, measured in units
863:   of 
864: $[(e^2/h)(N_1 N_2)/(N_1 + N_2)^2]^2$ as a function of the perpendicular
865: magnetic field $x$, for $a_{\bot}=3$, $a=0$, $b_{\bot}=0$. From bottom to
866: top, the curves are for $b=0$, $b=3$, and for the limit $b \to \infty$.}
867: \label{fig:2}
868: \end{figure}
869: 
870: The variance of the conductance in the presence of spin-orbit coupling is
871: shown in Fig.~\ref{fig:2}, for three different parallel field strengths. An
872: increase of the conductance fluctuations is observed around $x= \pm 
873: a_{\bot}$. This feature is explained noting that at $x=\pm a_{\bot}$ the
874: perpendicular magnetic field exactly cancels the effective magnetic field
875: due to the spin-orbit coupling term $a_{\bot}$ for one spin direction, so
876: that the effective Hamiltonian for that spin direction is real, not complex.
877: A parallel field and the spin-orbit coupling term $a$ suppress this feature
878: (data for $a$ not shown). Both effects also reduce the variance at zero
879: perpendicular field by a factor of two.
880: 
881: At finite temperatures, both the effects of thermal smearing and dephasing
882: need to be taken into account for a calculation of the variance of the
883: conductance. Dephasing (with dephasing time $\tau_{\phi }$) is incorporated
884: in the results (\ref{eq:covresult})--(\ref{eq:var_large}) by the
885: substitution $N\rightarrow N+\hbar /\tau_{\phi }\Delta $. The effect of
886: thermal smearing requires an integration over the energies $\varepsilon $
887: and $\varepsilon ^{\prime }$, 
888: \begin{equation}
889: \mbox{var}\,G\left( T\right) =
890:   \int d\varepsilon d\varepsilon ^{\prime }
891:   \frac{df}{d\varepsilon }\frac{df}{d\varepsilon ^{\prime }}
892:   \mbox{cov}\left[G(\varepsilon ),G(\varepsilon ^{\prime })\right] ,
893:   \label{eq:T}
894: \end{equation}%
895: where $f$ is the Fermi distribution function. The integrations over energy
896: needs to be done by numerical methods in all but a few special cases. We
897: refer to Ref.\ \onlinecite{thesis} for details. Equation (\ref{eq:T}) was
898: used for a quantitative comparison of theory and experiment by Zumb\"{u}hl 
899: \emph{et al}.\cite{Zumbuehl}
900: 
901: \subsection{Symmetry with respect to perpendicular magnetic field inversion}
902: 
903: \label{sec:4}
904: 
905: The presence of spin-orbit coupling has interesting consequences for the
906: symmetry of the conductance under reversal of the perpendicular field.
907: Although, in general, the conductance does not change when the total
908: magnetic field is reversed, $b\rightarrow -b$, $b_{\bot }\rightarrow
909: -b_{\bot }$, and $x\rightarrow -x$, the conductance need not be symmetric
910: under reversal of the perpendicular component $x$ only. In order to study
911: the symmetry of the conductance under reversal of the perpendicular
912: component $x$ of the magnetic field, we consider the correlator 
913: \begin{equation}
914: Q(x)=\frac{\left\langle G(x)G(-x)\right\rangle -\langle G(x)\rangle
915:   ^{2}}
916: {\mbox{var}\,G(x)}
917: \end{equation}%
918: for $x^{2}\gg N$. Perfect symmetry under reversal of the perpendicular
919: component of the magnetic field only implies $Q(x)=1$, while the absence
920: of any correlations between $G(x)$ and $G(-x)$ implies $Q(x)=0$. 
921: 
922: Note that the invariance of the two-terminal conductance conductance 
923: under a complete reversal of the magnetic field implies that
924: $$
925:   \left\langle G(x,b)G(-x,b)\right\rangle =
926:   \left\langle G(x,b)G(x,-b)\right\rangle,
927: $$
928: so that $Q$ can also be defined as a correlator for conductances
929: at opposite values of the parallel magnetic field, keeping the
930: perpendicular field constant.
931: 
932: In the absence of an in-plane magnetic field, reversal of the
933: perpendicular magnetic field component is the same as reversal of the total
934: magnetic field, so that $G$ is trivially symmetric in $x$. In the absence
935: of spin-orbit scattering, the sole effect of the parallel magnetic field is
936: to lift spin degeneracy, and symmetry of $G$ follows from the fact that the
937: conductance for each spin species is symmetric in $x$. In the presence of
938: both spin-orbit scattering and an in-plane field, there is no general
939: symmetry that requires that $G$ is symmetric under reversal 
940: $B_{3}\rightarrow -B_{3}$, however the measure of observable
941: asymmetry depends upon to what cross-over symmetry class the generated
942: random matrix theory ensemble belongs.
943: 
944: Indeed, calculation of the correlator $Q(x)$ using our general result 
945: (\ref{eq:covresult}) for the covariance of the conductance shows that the
946: symmetry of the conductance with respect to reversal of the perpendicular
947: component of the magnetic field is destroyed when either $b\neq 0$ and 
948: $a\neq 0$, $b_{\bot }\neq 0$ and $a_{\bot }\neq 0$, or $b_{\bot }\neq
949: 0$ and 
950: $a\neq 0$. In the case $a=0$ and $b_{\bot }=0$, the conductance remains
951: symmetric in $x$, even if $a_{\bot }\neq 0$ and $b\neq 0$. In that case, the
952: effective Hamiltonian (\ref{eq:hamiltonian}) still has the extra
953: symmetry 
954: $\sigma_{1}\tilde{\mathcal{H}}\sigma_{1}=\tilde{\mathcal{H}}$, which
955: ensures the symmetry with respect to the perpendicular magnetic
956: field reversal.\cite{AF} Symmetry in the perpendicular component of the
957: magnetic field is also preserved if the parallel magnetic field is so large
958: that both $b$, $b_{\bot }\neq 0$, but $a_{\bot }^{2}\ll N$ and $a^{2}\ll N$,
959: see Ref.\ \onlinecite{AF} for a symmetry argument. Figure \ref{fig:3} shows
960: plots of $Q(x)$ for the two minimal cases $a=0$ or $b_{\bot }=0$. 
961: 
962: \begin{figure}[tbp]
963: \epsfxsize=0.9\hsize 
964: \epsffile{fig4.eps} 
965: \caption{The correlator $Q(x)=(\left\langle G(x)G(-x)\right\rangle -\langle
966: G(x)\rangle ^{2})/\mbox{var}\,G(x)$ for $x^{2}\gg N$. In panel (a), $Q$ is
967: shown as a function of parallel field $b$, with $a_{\bot }=b_{\bot }=0$. In
968: panel (b), $Q$ is shown as a function of spin-orbit coupling $a_{\bot }$,
969: with $b=5$ and $a=0$.}
970: \label{fig:3}
971: \end{figure}
972: 
973: The general parameter dependence of $Q(x)$ for $x^{2}\gg N$ can be obtained
974: from Eq.\ (\ref{eq:covresult}). Simple expressions are found in the
975: cases 
976: $b\gg N$ and $a_{\bot }^{2}\gg N$, for which we have 
977: \begin{equation}
978: Q=\frac{N^{2}}{\left( 4a^{2}+2b_{\bot }^{2}+N\right) ^{2}},
979: \label{eq:theta_large}
980: \end{equation}%
981: and for $b\gg N$, while still $b_{\bot }=0$, 
982: \begin{eqnarray}
983: Q &=&1-\left[ 1-\frac{N^{2}}{(4a^{2}+N)^{2}}\right]  \\
984: &&\mbox{}\times \left[ 1-\frac{N^{2}}{2\left[ (a^{2}+a_{\bot
985: }^{2})^{2}+(a^{2}+a_{\bot }^{2}+N)^{2}\right] }\right] .  \notag
986: \end{eqnarray}%
987: On the other hand, if $b\gg N$ and $b_{\perp }\gg N$, the correlation
988: function $Q$ is given by 
989: \begin{equation}
990: Q=\frac{N^{2}}{\left( 2a^{2}+2a_{\bot }^{2}+N\right) ^{2}},
991: \label{eq:theta_bblarge}
992: \end{equation}%
993: confirming that $G$ remains a symmetric function of $x$ if $b$ and $b_{\perp
994: }$ are large but $a$ and $a_{\perp }$ are not.
995: 
996: Experimentally, the component of the conductance fluctuations 
997: that is antisymmetric with respect to the reversal
998: $B_{3}\rightarrow -B_{3}$ in the presence of in-plane magnetic 
999: field can be used as a test to the strength of spin-orbit coupling, 
1000: even if $b\ll N$. This is
1001: because an alternative mechanism of time-reversal symmetry breaking by an
1002: in-plane magnetic field via the inter-subband mixing\cite{FJ} 
1003: generates anti-symmetric conductance fluctuations in the third order
1004: in the field only, thus 
1005: $1-Q(x\gg N)\sim B_{\Vert }^{6}$. In contrast, the spin-orbit-coupling 
1006: induced asymmetry, which can be analyzed by expanding $Q(x)$
1007: with respect to $b^{2}$ and $b_{\bot }^{2}$, has
1008: \begin{widetext}
1009: \begin{eqnarray}
1010:   Q &=& 1 - 4 [4 a^2 b^2 (32 a^4 a_{\perp}^4 + 24 a^2 a_{\perp}^4 N
1011:   + 6 a_{\perp}^4 N^2 + 4 a_{\perp}^2 N^3 + N^4) 
1012:   \nonumber \\ && \mbox{}
1013:   + a_{\perp}^2 b_{\perp}^2 (128 a^6 a_{\perp}^4 + 96 a^4 a_{\perp}^4 N
1014:   + 24 a^2 a_{\perp}^4 N^2 + 4 a_{\perp}^4 N^3 + 6 a_{\perp}^2 N^4 +
1015:   3 N^5)]  \nonumber \\ && \mbox{} \times
1016:   [N ( 4 a^2 + N ) (2 a_{\perp}^2 + N)
1017:   (16 a^4 a_{\perp}^4 + 8 a^2 a_{\perp}^4 N + 2 a_{\perp}^4 N^2 + 
1018:   2 a_{\perp}^2 N^3 + N^4)]^{-1},
1019: \end{eqnarray}
1020: \end{widetext}
1021: for the case of uniform spin-orbit coupling. In the limit of a dot
1022: with a large number of channels $N \gg a_{\perp}^2$ but still
1023: in the universal regime $N \ll E_{\rm Th}/\Delta$, this simplifies
1024: to
1025: \begin{equation}
1026:   Q = 1 - 16 a^2 b^2 N^{-3}.
1027: \end{equation}
1028: Replacing the dimensionless spin-orbit coupling and magnetic
1029: field $a_{\perp}$ and $b_{\perp}$ by their dimensionful
1030: counterparts, one arrives at the result (\ref{eq:AsymmetricPart}) 
1031: advertised in the introduction.
1032: 
1033: \subsection{Autocorrelation function and the ``correlation echo''}
1034: 
1035: Correlations between the conductance $G$ at different values of the
1036: perpendicular magnetic field $x$ are studied through the conductance
1037: autocorrelation function 
1038: \begin{equation}
1039: \chi(x,x^{\prime}) = \left<G(x)G(x^{\prime})\right> - \langle G(x) \rangle
1040: \langle G(x^{\prime}) \rangle.
1041: \end{equation}
1042: For large perpendicular fields, $x^2, x^{\prime}{}^2 \gg N$, the
1043: autocorrelator $\chi(x,x^{\prime})$ depends on the difference 
1044: $(x-x^{\prime}) $ only. Since time-reversal symmetry is fully broken for such
1045: large magnetic fields, any features in the autocorrelator are signatures of
1046: the spin-orbit coupling.
1047: 
1048: For zero parallel field, $b=b_{\bot}=0$, one finds from Eq.\ 
1049: (\ref{eq:covresult}), 
1050: \begin{eqnarray}
1051: \chi(\xi) &=& \left(\frac{2 e^2}{h} \frac{N_1 N_2}{N_1 + N_2} \right)^2 
1052: \left[\frac{1}{(2N + 8a^2 + \xi^2)^{2}}\right.  \notag \\
1053: && \mbox{} + \frac{1}{(2N + \xi^2)^{2}} + \frac{1}{(2N + 4a^2 + (-2a_{\bot}
1054: + \xi)^2)^{2}}  \notag \\
1055: && \left. \mbox{} + \frac{1}{(2N + 4a^2 + (2a_{\bot} + \xi)^2)^{2}}\right],
1056: \label{eq:chi_b0}
1057: \end{eqnarray}
1058: where $\xi=x-x^{\prime}$.
1059: 
1060: \begin{figure}[tbp]
1061: \epsfxsize=0.8\hsize 
1062: \hspace{0.1\hsize} 
1063: \epsffile{fig5.eps} 
1064: \caption{The correlator $\protect\chi(x-x^{\prime})$ as a function of the
1065: dimensionless perpendicular magnetic field difference $(x-x^{\prime})$, for
1066: three values of the spin-orbit term $a_{\bot}$: $a_{\bot}=0$, $a_{\bot}=1$,
1067: and $a_{\bot}=3$. Curves in panel (a) are for $a=0$ and $b=b_{\bot}=0$.
1068: Curves in panel (b) are for $a=0.5$ and $b=b_{\bot}=0$. Panel (c) shows the
1069: effect of the parallel field, $b=2$, with $a=b_{\bot}=0$. In all three
1070: panels, $\protect\chi(x-x^{\prime})$ is measured in units of $[(e^2/h)
1071: N_{1}^{2}N_{2}^{2}/(N_1+N_2)^{2}]^2$. }
1072: \label{fig:4}
1073: \end{figure}
1074: 
1075: A plot of $\chi (x-x^{\prime })$ for different values of $a_{\bot }$ is
1076: shown in Fig.~\ref{fig:4}. Remarkably, the conductance autocorrelator shows
1077: an \textquotedblleft echo\textquotedblright\ at $|x-x^{\prime }|=
1078: 2a_{\bot }$. This echo appears even for large values of $a_{\bot }$, 
1079: for which the
1080: conductance correlations have decayed for intermediate values of the
1081: magnetic-field difference, see Fig.\ \ref{fig:4}a. The magnetic-field echo
1082: has the same origin as the reappearance of the weak localization and
1083: conductance fluctuations at finite values of the perpendicular magnetic
1084: field $x$: The spin-orbit term $a_{\bot }$ acts as a homogeneous
1085: perpendicular field with a different direction for the up and down spins
1086: (defined with respect to the axis $\mathbf{\hat{e}}_{3}$ perpendicular to
1087: the plane of the quantum well). In the presence of an external perpendicular
1088: magnetic field $x$, Up spins and down spins experience effective
1089: perpendicular magnetic fields $x+a_{\bot }$ and $x-a_{\bot }$, respectively.
1090: If the magnetic field difference $x-x^{\prime }$ equals $\pm 2a_{\bot }$,
1091: one spin species at field $x$ moves in the same effective perpendicular
1092: field as the other spin species at field $x^{\prime }$. Hence, in the
1093: absence of a parallel field and a spin-orbit term $a$, the conductance has
1094: identical fluctuating contributions at $x$ and $x^{\prime }$. However, only
1095: one spin species is involved in the \textquotedblleft
1096: echo\textquotedblright\ --- the autocorrelation function reaches only one
1097: half of the total variance of the conductance; the other spin species have
1098: uncorrelated contributions to the conductance at the external fields $x$ and 
1099: $x^{\prime }$. A magnetic field in the plane of the quantum well and the
1100: spin-orbit coupling term $a$ suppress the echo, since they cause scattering
1101: between the $\pm \mathbf{\hat{e}}_{3}$ spin directions, see Figs.\ 
1102: \ref{fig:4}b and c.
1103: 
1104: \section{Effect of a non-uniform SO coupling.}
1105: \label{sec:nonuniform}
1106: 
1107: The expressions of the previous two sections are valid for arbitrary
1108: magnitudes of the dimensionless spin-orbit rates $a_{\perp}$ and 
1109: $a$. In the case of uniform spin-orbit coupling, one has the
1110: strong inequality $a \ll a_{\perp}$. In general, the spin-orbit 
1111: coupling may be non-uniform throughout the quantum dot, {\em e.g.}, 
1112: as a result of a spatial modulation of the asymmetry of the quantum 
1113: well or as a result of a modulation of the electron density 
1114: (and therefore confining potential width) across the dot. 
1115: The main effect of a non-uniformity of the spin-orbit coupling 
1116: is an increase of $a$,\cite{BCH} thus promoting the crossover into 
1117: the standard symplectic ensemble. In this section, we'll give a
1118: quantitative estimate of this effect for a special choice of the
1119: non-uniformity. Another effect of a non-uniform spin-orbit coupling 
1120: is a rescaling of $a_{\perp }$, and will not be considered here.
1121: 
1122: Assuming that the non-uniformity of spin-orbit coupling affects only
1123: its strength without changing the generic form of the Hamiltonian, we
1124: describe its effect via coordinate dependent spin-orbit coupling
1125: lengths $\lambda_{1,2}^{-1}(\mathbf{r})$,
1126: \begin{eqnarray}
1127: \mathcal{H}\! &=&\!\frac{1}{2m}\left( \mathbf{p}-e\mathbf{A}-
1128:   \mathbf{a}
1129:   \right) ^{2}, \nonumber \\
1130:   \mathbf{a} &=& \frac{\sigma_{2}}{2\lambda_{1}(\mathbf{r})}
1131:   \mathbf{\hat{e_1}} -
1132:   \frac{\sigma_{1}}{2\lambda_{2}(\mathbf{r})}
1133:   \mathbf{\hat{e_2}}.
1134: \end{eqnarray}%
1135: In the case that 
1136: $$
1137:   \beta =\nabla \times \mathbf{a} =
1138:   \frac{\sigma_{2}}{2}\partial_{x_{2}}\lambda_{1}^{-1}(\mathbf{r})
1139:   + \frac{\sigma_{1}}{2}\partial_{x_{1}}\lambda_{2}^{-1}(\mathbf{r})
1140:   \neq 0,
1141: $$
1142: it is not possible to eliminate the spin-orbit coupling from the
1143: Hamiltonian in the first order in spin-orbit coupling constant by a unitary
1144: transformation. Qualitatively, this explains why the spatial variation
1145: of the spin-orbit coupling length gives rise of a term with the
1146: symmetry of $\mathbf{a_{\parallel}}$. For a quantitative estimate of
1147: how much the spatial variations of $\lambda_{1,2}$ contribute to the
1148: energy scale $\varepsilon_{\parallel}$,
1149: we note that it is still possible to find a transformation
1150: $U$ such that the resulting non-Abelian vector potential in a locally
1151: rotated spin frame,
1152: \begin{equation}
1153: \mathbf{\tilde{a}=}-iU^{\dagger }\nabla U-U^{\dagger }\mathbf{a}U,
1154: \end{equation}%
1155: would satisfy the conditions
1156: \begin{equation}
1157:   \nabla \cdot \mathbf{\tilde{a}} = 0
1158: \end{equation}
1159: in the interior of the quantum dot and $ \mathbf{\hat{n}}\cdot
1160: \mathbf{\tilde{a}}=0$, on the dot boundary, where $\mathbf{\hat{n}}$
1161: is the unit vector normal to the dot's boundary, at least, in the
1162: lowest order in SO coupling. Such conditions would enable one to
1163: satisfy boundary conditions for Cooperon and diffuson propagators from
1164: diagrammatic perturbation theory, which is needed in order to make the
1165: zero-dimensional approximation necessary for the use of random matrix
1166: theory. This program is carried out in the appendix for a circular
1167: quantum dot with a special choice of the coordinate dependence of the
1168: spin-orbit coupling lengths $\lambda_{1,2}$.
1169: 
1170: Here, as an example, we consider here the case where the spin-orbit
1171: coupling is stronger in the center of a quantum dot (where 
1172: density is higher and the well is narrower) than
1173: at the edges (where density is lower and the well is broader),
1174: \begin{equation}
1175: \frac{1}{\lambda_{1,2}}=\dfrac{1}{\lambda ^{\mathrm{c}}}-\frac{\left(
1176: r/L\right) ^{2}}{\lambda ^{\mathrm{f}}},  \label{eq:nonunif}
1177: \end{equation}%
1178: where $L$ is the typical size of the dot, without making a 
1179: restriction of the size of the dot. We perform a unitary
1180: transformation
1181: \begin{equation}
1182: U = \exp \left( i\frac{x_{1}\sigma_{2}-x_{2}\sigma_{1}}
1183:   {2\lambda^{\mathrm{c}}}\left[ 1-
1184:   \dfrac{\lambda^{\mathrm{c}}}{3\lambda ^{\mathrm{f}}}
1185: \left( \dfrac{r}{L}\right) ^{2}\right] \right),
1186: \end{equation}
1187: which rotates to a local spin reference axis.
1188: With this transformation, the matrix vector $\mathbf{\tilde{a}}$
1189: becomes, to lowest order in the inverse spin-orbit coupling length,
1190: \begin{eqnarray}
1191: \mathbf{\tilde{a}} &=& -i U^{\dagger} \nabla U - 
1192:   \mathbf{a} + {\cal O}(\lambda^{-2}) \nonumber \\
1193: &=&-\dfrac{x_{1}\sigma_{1}+x_{2}\sigma_{2}}{3\lambda ^{\mathrm{f}}L^2}\left[ 
1194: \mathbf{\hat{e}}_{3}\times \mathbf{r}\right].
1195: \end{eqnarray}%
1196: We conclude that (a) the spin relaxation and the crossover into 
1197: symplectic ensemble start, indeed, in the lowest order in the
1198: non-uniform part of the spin-orbit coupling and (b) for the
1199: non-uniform coupling (\ref{eq:nonunif}), the functional form of
1200: the perturbation after transformation
1201: is equal to that of the term $\mathbf{a_{\parallel}}$
1202: in Eq.\ (\ref{eq:hamiltonian}). The latter observation allows us to
1203: express the contribution of the non-uniformity to the energy scale
1204: $\varepsilon_{\parallel}$ as
1205: \begin{equation}
1206:   \varepsilon_{\parallel} =
1207:   \frac{1}{4} \kappa' \kappa E_{\rm Th}
1208:   \left( \frac{L^2}{4 \lambda^{{\rm f}2}} \right),
1209:   \label{eq:kappageneral}
1210: \end{equation}
1211: where the geometry-dependent coefficients $\kappa$ and $\kappa'$ are
1212: the same as those defined for the spatially homogeneous spin-orbit
1213: coupling in Eq.\ (\ref{eq:kappa}).
1214: 
1215: \section{Conclusions}
1216: 
1217: \label{sec:5}
1218: 
1219: In this paper we calculated the average and the fluctuations of the
1220: conductance of a chaotic quantum dot in the presence of Bychkov-Rashba
1221: and (linear) Dresselhaus spin-orbit coupling and a magnetic field and
1222: studied the symmetry of the conductance under reversal of the 
1223: perpendicular component of the magnetic
1224: field. The calculations were done to leading order in $1/N$, where
1225: $N=N_1+N_2$ is the total number of open channels in the leads
1226: connecting the quantum dot to the reservoirs.  After a unitary
1227: transformation, the spin-orbit term in the Hamiltonian and the Zeeman
1228: coupling to the external magnetic field are transformed into four
1229: terms that all have different symmetries.\cite{AF}
1230: The corresponding energy scales
1231: are the conventional Zeeman energy $\varepsilon^{\rm Z} = \mu_B g B$ and
1232: three energy scales that depend on the spin-orbit coupling: two
1233: spin-orbit scales $\varepsilon^{\rm so}_{\perp}$ and $\varepsilon^{\rm
1234: so}_{\parallel}$, and the second Zeeman energy scale $\varepsilon^{\rm
1235: Z}_{\perp}$. With the help of the theory presented here, the energy
1236: scales $\varepsilon^{\rm so}_{\perp}$, $\varepsilon^{\rm
1237: so}_{\parallel}$, and $\varepsilon^{\rm Z}_{\perp}$ can be obtained
1238: from a measurement of the conductance autocorrelation function of the
1239: quantum dot.
1240: 
1241: Our results provide several routes to a direct measurement of the
1242: product of the spin-orbit coupling lengths $\lambda_1$ and 
1243: $\lambda_2$. The product $\lambda_1 \lambda_2$ can be found from
1244: a ``magnetic-field echo'' in the conductance fluctuations as a 
1245: function of the perpendicular magnetic field. The magnetic-field
1246: echo appears at the magnetic-field difference
1247: \begin{equation}
1248:   \Delta B = \frac{\hbar}{e \lambda_1 \lambda_1}.
1249: \end{equation}
1250: The same magnetic-field scale splits the weak-localization peak in the
1251: absence of a parallel magnetic field. The ratio $\lambda_1/\lambda_2$
1252: can be found from a measurement of the energy scale $\varepsilon^{\rm
1253: Z}_{\perp}$ in a roughly circular quantum dot. In this case,
1254: $(\lambda_2/\lambda_1)^2$ is equal to the ratio of the energy scales
1255: $\varepsilon^{\rm Z}_{\perp}$ for the in-plane magnetic field in the
1256: crystallographic directions $\hat e_1 = [110]$ and $\hat e_2 = [1\bar
1257: 1 0]$.  Measurement of $\varepsilon^{\rm so}_{\parallel}$ is less
1258: suitable to determine $\lambda_1/\lambda_2$, as it may be
1259: affected by non-uniformities in the spin-orbit scattering rate.
1260: 
1261: The theory presented here has been used for the interpretation of
1262: measurements of the conductance of quantum dots in GaAs/GaAlAs 
1263: heterostructures with spin-orbit coupling as a function of 
1264: parallel and perpendicular components of the magnetic field
1265: by Zumb\"{u}hl {\em et al.}\cite{Zumbuehl} 
1266: The quantum dots used in the experiments of Ref.\
1267: \onlinecite{Zumbuehl}
1268: were built of the same material, but they varied in size, ranging 
1269: from $L/\lambda=0.27$ to $0.64$.
1270: Here $\lambda=(\lambda_{1}\lambda_{2})^{1/2}$ is the
1271: geometric average of the spin-orbit lengths $\lambda_1$ and
1272: $\lambda_2$. Whereas the conductance of smaller dots had a
1273: minimum at zero perpendicular magnetic field, the conductance
1274: of larger dots had a (local) maximum at zero magnetic field. 
1275: {}From a fit to our expression for the average
1276: conductance at zero parallel field, Eq. (\ref{eq:weak}), values for the
1277: experimentally unknown parameters $\lambda =
1278: 4.4\mu \mathrm{m}$, the decoherence
1279: time $\tau_{\phi }$ and a geometrical factor, see Eqs. (\ref{eq:E_so}), were
1280: extracted. 
1281: Measurement of the average conductance in
1282: the presence of a magnetic field parallel to the plane of the
1283: two-dimensional electron gas was in good agreement with our Eq.\
1284: (\ref{eq:weak}), without additional fit parameters and
1285: up to a parallel field of about 0.3 T.\cite{Zumbuehl} 
1286: Upon inclusion of the effects of time-reversal symmetry breaking by a strong parallel field
1287: due to the finite extent of the electron wavefunction across the
1288: heterostructure,\cite{FJ} the authors of Ref.\ \onlinecite{Zumbuehl}
1289: could extend the agreement between theory and experiment is 
1290: extended to even higher magnetic fields.
1291: 
1292: \begin{acknowledgments}
1293: We thank Igor Aleiner, Boris Altshuler,
1294: Bertrand Halperin, Charles Marcus, Jeffrey Miller,
1295: Yuli Nazarov, and Dominik Zumb\"uhl
1296: for important discussions.
1297: This work was supported by NSF under grant no.\ PHY 0117795
1298: (JC), by the NSF under grant no.\ DMR
1299: 0086509 and by the Packard foundation (PWB), and
1300: by EPSRC and NATO CLG (VF).
1301: \end{acknowledgments}
1302: 
1303: \appendix
1304: 
1305: \section{Geometry-dependent coefficients for a circular quantum dot}
1306: 
1307: \label{app:1}
1308: 
1309: In this appendix, we calculate 
1310: \begin{eqnarray}
1311: \mathcal{D}_{\mu_{1}\nu_{1};\mu_{2}\nu_{2}}
1312:   (\mathbf{r}_{1},\mathbf{r}_{2}) &=&
1313:   \langle G_{\mu_{1}\nu_{1}}^{\mathrm{R}}
1314:   (\mathbf{r}_{1},\mathbf{r}_{2})(E,B)  \notag \\
1315:   &&\mbox{}\times G_{\mu_{2}\nu_{2}}^{\mathrm{A}}
1316:   (\mathbf{r}_{1},\mathbf{r}_{2})(E^{\prime },B^{\prime })\rangle
1317: \end{eqnarray}%
1318: for a disordered circular quantum dot with spin-orbit scattering. Here 
1319: $G^{\mathrm{R}}$ and $G^{\mathrm{A}}$ are the retarded and advanced Green
1320: functions (inverses of $\varepsilon -\mathcal{H}$ for $\varepsilon $ just
1321: above and just below the real axis, respectively), respectively, and the
1322: indices $\mu_{1}$, $\mu_{2}$, $\nu_{1}$, and $\nu_{2}$ refer to the
1323: electron spin. Comparison of our result with the random matrix theory of
1324: Sec.\ \ref{sec:2} allows us to find the geometry-dependent
1325: coefficients 
1326: $\kappa $, $\kappa ^{\prime }$, and $\kappa ^{\prime \prime }$ of
1327: Eqs.\ (\ref{eq:E_so}).
1328: 
1329: In a closed dot of volume $V$ and with diffusive dynamics, the
1330: diffuson 
1331: $\mathcal{D}(\mathbf{r}{_{1},\mathbf{r}_{2}})$ obeys the equation 
1332: \begin{equation}
1333: (i\hbar \omega +\Pi )\mathcal{D}(\mathbf{r}{_{1},\mathbf{r}_{2}})=\frac{2\pi 
1334: }{V\Delta }\delta (\mathbf{r}_{1}-\mathbf{r}_{2})\openone\otimes
1335:   \openone,
1336:   \label{eq:diff}
1337: \end{equation}%
1338: where $\hbar \omega =\varepsilon ^{\prime }-\varepsilon $ and the
1339: differential operator $\Pi $ is given by 
1340: \begin{eqnarray}
1341: \Pi &=&(D/\hbar )(-i\hbar \nabla_{\mathbf{r}_{1}}+\mathbf{a}\otimes
1342: \openone
1343: -\openone\otimes \mathbf{a}^{\prime })^{2}  \notag \\
1344: &&\mbox{}+ih_{\mathrm{Z}}\otimes \openone-i\openone\otimes 
1345: h_{\mathrm{Z}}^{\prime }.  \label{eq:Pidef}
1346: \end{eqnarray}%
1347: Here $D=v_{F}l/2$ is the diffusion coefficient in the quantum dot, the
1348: matrix vector potentials $\mathbf{a}$ and $\mathbf{a}^{\prime }$ are the sum of a
1349: spin-independent contribution from the perpendicular magnetic field and a
1350: spin-dependent contribution from the Bychkov-Rashba and Dresselhaus 
1351: spin-orbit
1352: coupling, 
1353: \begin{eqnarray}
1354: \mathbf{a} &=&\left( \frac{eB_{3}y}{2c}-\frac{\hbar }{2\lambda_{1}}\sigma
1355: _{2}\right) \mathbf{\hat{e}}_{1}-\left(
1356: \frac{eB_{3}x}{2c}-\frac{\hbar}
1357: {2\lambda_{2}}\sigma_{1}\right) \mathbf{\hat{e}}_{2},  \notag \\
1358: \mathbf{a}^{\prime } &=&\left( \frac{eB_{3}^{\prime }y}{2c}-
1359: \frac{\hbar }{2\lambda_{1}}\sigma_{2}\right) \mathbf{\hat{e}}_{1}-
1360: \left( \frac{eB_{3}^{\prime }x}{2c}-\frac{\hbar }{2\lambda_{2}}
1361: \sigma_{1}\right) 
1362: \mathbf{\hat{e}}_{2},
1363:   \nonumber \\
1364: \end{eqnarray}%
1365: while $h_{\mathrm{Z}}$ and $h_{\mathrm{Z}}^{\prime }$ represent the Zeeman
1366: coupling to the magnetic field, 
1367: $$
1368: h_{\mathrm{Z}}=\frac{1}{2}\mu_{B}g\mathbf{B}\cdot \mbox{\boldmath $\sigma$},
1369: \ \ 
1370: h_{\mathrm{Z}}^{\prime }=\frac{1}{2}\mu_{B}g\mathbf{B}^{\prime }\cdot 
1371: \mbox{\boldmath $\sigma$}.
1372: $$
1373: Equation (\ref{eq:diff}) is supplemented with the boundary condition 
1374: \begin{equation}
1375: \hat{n}\cdot (-i\hbar \nabla_{\mathbf{r}_{1}}+\mathbf{a}\otimes \openone-
1376: \openone\otimes \mathbf{a}^{\prime })\mathcal{D}(\mathbf{r}{_{1},
1377: \mathbf{r}_{2}})=0  \label{eq:boundary}
1378: \end{equation}%
1379: at the sample boundary, where $\hat{n}$ is the unit vector normal to the
1380: boundary.
1381: 
1382: To find $\mathcal{D}$, we shift to the basis of eigenfunctions of the
1383: diffusion equation (i.e., eigenfunctions of $\Pi $ in the absence of
1384: spin-orbit scattering and a magnetic field) and calculate the matrix
1385: elements of $\Pi $ in that basis. For this procedure to work, it is
1386: important that the matrix vector potentials $\mathbf{a}$ and $\mathbf{a}^{\prime }$
1387: do not change the Von Neumann boundary conditions (\ref{eq:boundary}), i.e.,
1388: that 
1389: \begin{equation}
1390: \mathbf{\hat{n}}\cdot \mathbf{a}=\mathbf{\hat{n}}\cdot \mathbf{a}^{\prime }=0
1391: \label{eq:r1}
1392: \end{equation}%
1393: at the boundary of the quantum dot. This requirement ensures that the
1394: eigenfunctions of the diffusion equation with Von Neumann boundary
1395: conditions can be used to construct the diffuson $\mathcal{D}$. In order to
1396: fix the gauge, we also require that 
1397: \begin{equation}
1398: \nabla \cdot \mathbf{a}=\nabla \cdot \mathbf{a}^{\prime }=0  \label{eq:r2}
1399: \end{equation}%
1400: in the interior of the dot (London gauge).
1401: 
1402: As discussed in Ref.\ \onlinecite{AF} and in the introduction of this paper,
1403: the requirements (\ref{eq:r1}) and (\ref{eq:r2}) are realized by performing
1404: a suitable unitary transformation $U$ to the Hamiltonian, $\mathcal{H} \to
1405: U^{\dagger} \mathcal{H} U$. For Eq.\ (\ref{eq:diff}), such a unitary
1406: transformation implies 
1407: \begin{eqnarray}
1408: \mathcal{D} &\to& (U^{\dagger} \otimes U) \mathcal{D} (U \otimes
1409: U^{\dagger}), \\
1410: \Pi &\to& (U^{\dagger} \otimes U) \Pi (U \otimes U^{\dagger}).
1411: \end{eqnarray}
1412: 
1413: We'll now perform an explicit calculation for the case of a circular quantum
1414: dot of radius $L$, the origin of our coordinate system being chosen at the
1415: center of the dot. In polar coordinates $x_{1}=r\cos \phi $, $x_{2}=r\sin
1416: \phi $, the tensor eigenfunctions of the diffusion equation are of the form 
1417: \begin{eqnarray}
1418: f_{n0}^{ij}(r,\phi ) &=&c_{n0}J_{0}(rx_{n0}/L)\,\sigma_{i}\otimes \sigma
1419: _{j},  \notag \\
1420: f_{nm}^{ij}(r,\phi ) &=&c_{nm}J_{m}(rx_{nm}/L)\cos (m\phi )\,\sigma
1421: _{i}\otimes \sigma_{j},  \notag \\
1422: g_{nm}^{ij}(r,\phi ) &=&c_{nm}J_{m}(rx_{nm}/L)\sin (m\phi )\,\sigma
1423: _{i}\otimes \sigma_{j},
1424: \end{eqnarray}%
1425: where $i,j=0,1,2,3$, $m=1,2,\ldots $, $x_{nm}$, $n=1,2,\ldots $, is
1426: the 
1427: $n$th positive root of $J_{m}^{\prime }(x)=0$, and we used the 
1428: notation $\sigma
1429: _{0}=\openone$. The corresponding eigenvalues are 
1430: \begin{equation}
1431: \lambda_{nm}=D\hbar (x_{nm}/L)^{2}
1432: \end{equation}%
1433: and the normalization constants are 
1434: \begin{equation}
1435: c_{nm}^{-2}=\frac{1}{2}\pi
1436: L^{2}(J_{m}(x_{nm})^{2}-J_{m+1}(x_{nm})^{2})(1+\delta_{m0}).
1437: \end{equation}%
1438: The smallest eigenvalue is found for $n=m=0$, $\lambda_{00}=0$; all other
1439: eigenvalues are larger than the Thouless energy $E_{\mathrm{Th}}=D\hbar
1440: /L^{2}$.
1441: 
1442: For a circular quantum dot, the magnetic-field contribution to the matrix
1443: vector potential of Eq.\ (\ref{eq:Pidef}) already satisfies the requirements
1444: (\ref{eq:r1}) and (\ref{eq:r2}). For the spin-orbit contribution we use the
1445: transformation 
1446: \begin{eqnarray}
1447: U &=& \exp \left[i \frac{x_1 \sigma_2}{2 \lambda_1} - i \frac{x_2
1448:     \sigma_1}{2 \lambda_2} \right. \\
1449: && \left. \mbox{} - i \frac{(x_1^2 + x_2^2 - 3 L^2)}{48 \lambda_1 \lambda_2}
1450: \left( \frac{x_2 \sigma_1}{\lambda_1} - \frac{x_1 \sigma_2}{\lambda_2}
1451: \right) \right] ,  \notag
1452: \end{eqnarray}
1453: which fulfills the requirements (\ref{eq:r1}) and (\ref{eq:r2}) to leading
1454: order in $L/\lambda$. [Note: although the transformation of 
1455: Eq.\ (\ref{eq:U}) is sufficient to satisfy the requirements to order 
1456: $L/\lambda$, the
1457: remaining term $a_{\Vert}$ of Eq.\ (\ref{eq:hamiltonian}) does not obey Eq.\
1458: (\ref{eq:r2}).]
1459: 
1460: With this transformation, the matrix vector potential $\mathbf{a}$ and the
1461: Zeeman term $h_{\rm Z}$ change to 
1462: \begin{eqnarray}
1463: \mathbf{a} &\rightarrow &\mathbf{\hat{e}}_{1}\left[ \frac{eB_{3}x_{2}}{2c}+
1464: \frac{\hbar x_{2}\sigma_{3}}{4\lambda_{1}\lambda_{2}}\right.  \notag \\
1465: &&\left. \mbox{}+\frac{\hbar }{16\lambda_{1}\lambda_{2}}\left( 
1466: \frac{2x_{1}x_{2}\sigma_{1}}{\lambda
1467:  _{1}}+\frac{(3x_{2}^{2}+x_{1}^{2}-L^{2})
1468: \sigma_{2}}{\lambda_{2}}\right) \right]  \notag \\
1469: &&\mbox{}-\hat{e}_{2}\left[ \frac{eB_{3}x_{1}}{2c}+\frac{\hbar x_{1}\sigma
1470: _{3}}{4\lambda_{1}\lambda_{2}}\right.  \notag \\
1471: &&\left. \mbox{}+\frac{\hbar }{16\lambda_{1}\lambda_{2}}\left( 
1472: \frac{2x_{1}x_{2}\sigma_{2}}{\lambda_{2}}+\frac{(3x_{1}^{2}+x_{2}^{2}-L^{2})
1473: \sigma_{1}}{\lambda_{1}}\right) \right] ,  \notag \\
1474: h_{\mathrm{Z}} &\rightarrow &\frac{1}{2}\mu_{B}g\mathbf{B}\cdot 
1475: \mbox{\boldmath $\sigma$}-\frac{1}{2}\sigma_{3}\mu_{B}g\left(
1476: \frac{B_{1}x_{1}}
1477: {\lambda_{1}
1478: }+\frac{B_{2}x_{2}}{\lambda_{2}}\right) ,  \label{eq:anew}
1479: \end{eqnarray}%
1480: with similar changes for $\mathbf{a}^{\prime }$ and $h^{\prime}_{\rm Z}$.
1481: 
1482: As long as all relevant energy scales are much smaller than 
1483: $E_{\mathrm{Th}}$, it is sufficient to know the action of $\Pi $ on
1484: the 16 eigenfunctions at $f_{00}^{ij}$ at $n=m=0$. In tensor notation, 
1485: this action is 
1486: \begin{widetext}
1487: \begin{eqnarray}
1488:   \Pi &=&
1489:   \left. \frac{D \hbar}{2 L^2}
1490:   \left(\frac{e}{c h} (\Phi - \Phi') \openone \otimes \openone
1491:     + \frac{L^2}{4 \lambda^2} (\sigma_3 \otimes \openone -
1492:       \openone \otimes \sigma_3) \right)^2 
1493:   \right. \ \ \nonumber \\ && \left. \mbox{} +
1494:   \frac{D \hbar L^6}{96 \lambda^4}
1495:   \left( \frac{\sigma_1 \otimes \openone -
1496:   \openone \otimes \sigma_1}{2 \lambda_1} \right)^2
1497:   + \frac{D \hbar L^6}{96 \lambda^4}
1498:   \left( \frac{\sigma_2 \otimes \openone -
1499:   \openone \otimes \sigma_2}{2 \lambda_2} \right)^2
1500:   \right. \ \ \nonumber \\ && \left. \mbox{} +
1501:   i \frac{\mu_B g \mathbf{B}}{2} (\mbox{\boldmath $\sigma$} \otimes \openone
1502:   - \openone \otimes \mbox{\boldmath $\sigma$}) 
1503:   + \frac{\mu_B^2 g^2 L^4}{D \hbar } \kappa_h
1504:     \left(\frac{B_1^2}{\lambda_1^2} + \frac{B_2^2}{\lambda_2^2}
1505:     \right)
1506:     (\openone \otimes \sigma_3 - \sigma_3 \otimes \openone)^2
1507:     \right..
1508: \end{eqnarray}
1509: \end{widetext}where $\Phi =B_{3}\pi L^{2}$ is the magnetic flux through the
1510: quantum dot and 
1511: $$
1512: \kappa_{h}=\frac{1}{2}\sum_{n=1}^{\infty}
1513: \frac{1}{x_{1n}^{4}(x_{1n}^{2}-1)}
1514: \approx 0.0182
1515: $$
1516: is a numerical constant. 
1517: Confining ourselves to the case $\lambda =\lambda_{1}=\lambda_{2}$ and
1518: comparing this result to Eq.\ (\ref{eq:D}), we can make the identifications 
1519: \begin{eqnarray}
1520:   x^{2} &=&\left( \frac{e\Phi }{ch}\right) ^{2}\frac{\hbar v_{F}l\pi }
1521:   {L^{2}\Delta }, \\
1522:   a_{\bot }^{2} &=&
1523:   \left( \frac{L^{2}}{4\lambda^{2}}\right)^{2}\frac{\hbar
1524:   v_{F}l\pi }{L^{2}\Delta }, \\
1525:   a^{2} &=&\left( \frac{L^{2}}{4\lambda ^{2}}\right)^{3} \frac{v_{F}l\pi}
1526:   {3L^{2}\Delta }, \\
1527:   b &=&\frac{\mu_{B}g\pi B}{\Delta }, \\
1528:   b_{\bot } &\approx &
1529:   1.83\frac{L^{2}}{\hbar v_{F}l\Delta }(\mu_{B}Bg)^{2}
1530:   \left( \frac{L^{2}}{4\lambda ^{2}}\right) .
1531: \end{eqnarray}%
1532: For the geometry-dependent constants $\kappa $, $\kappa ^{\prime }$, 
1533: and $\kappa ^{\prime \prime }$ in Eq.\ (\ref{eq:E_so}), this implies 
1534: $\kappa =2$, $\kappa ^{\prime }=1/3$ and $\kappa ^{\prime \prime }
1535: \approx 0.292$.
1536: 
1537: If the spin-orbit coupling is not uniform throughout the quantum
1538: dot, see Sec.\ \ref{sec:nonuniform}, the above estimates 
1539: for $a_{\perp}$ and $a$ and the corresponding
1540: energy scales $\varepsilon_{\bot}^{\rm so}$ and
1541: $\varepsilon_{\Vert}^{\rm so}$ need to be revisited. Here we 
1542: consider the example of Eq.\ (\ref{eq:nonunif}),
1543: $\lambda^{-1}_{1,2}(\mathbf{r}) =
1544: \lambda_{\rm c}^{-1} + (r^2/L^2 \lambda^{\rm f})$, and
1545: construct the unitary transformation $U$ that ensures that the
1546: conditions (\ref{eq:r1}) and (\ref{eq:r2}) are satisfied to 
1547: lowest order in $1/\lambda_{f}$,
1548: \begin{eqnarray}
1549:   U &=& \exp \left[i \frac{x_1 \sigma_2}{2 \lambda_1^{\rm c}}
1550:   - i \frac{x_2 \sigma_1}{2 \lambda_2^{\rm c}}
1551:   \right. \nonumber \\ && \left. \mbox{}
1552:   + i x_1 \sigma_2 \frac{(x_1^2 + x_2^2 + L^2)}
1553:   {8 L^2 \lambda^{\rm f}_{ 1}}
1554:   \right. \\ && \left. \mbox{}
1555:   - i x_2 \sigma_1 \frac{(x_1^2 + x_2^2 + L^2)}
1556:   {8 L^2 \lambda^{\rm f}_{ 2}}
1557:   \right] \nonumber .
1558: \end{eqnarray}
1559: With this transformation, the matrix vector $\mathbf a$ is mapped
1560: to
1561: \begin{eqnarray}
1562:   \mathbf a &\to&  \mathbf{\hat e_1} \left[
1563:   \frac{e B_3 x_2}{2 c} 
1564:   + \frac{\hbar x_2 \sigma_3}{4 \lambda_1^{\rm c} \lambda_2^{\rm c}} 
1565:   \right. \nonumber \\ && \left. \mbox{} 
1566:   - \frac{\hbar}{8 L^2}
1567:   \left(\frac{2 x_1 x_2 \sigma_1}{\lambda^{\rm f}_{ 2}} +
1568:   \frac{(3 x_2^2 + x_1^2 - L^2) \sigma_2}{\lambda^{\rm f}_{ 1}}
1569:   \right)
1570:   \right]
1571:   \nonumber \\
1572:   && \mbox{} -  \mathbf{\hat e_2} \left[\frac{e B_3 x_1}{2 c} 
1573:   + \frac{\hbar x_1 \sigma_3}{4 \lambda_1^{\rm c} \lambda_2^{\rm c}}  
1574:   \right. \nonumber \\ && \left. \mbox{}
1575:   - \frac{\hbar}{8 L^2}
1576:   \left(\frac{2 x_1 x_2 \sigma_2}{\lambda^{\rm f}_{ 1}} +
1577:   \frac{(3 x_1^2 + x_2^2 - L^2) \sigma_1}{\lambda^{\rm f}_{ 2}} \right)
1578:   \right].
1579:   \nonumber \\
1580:   \label{eq:anonunif}
1581: \end{eqnarray}
1582: Comparing Eqs.\ (\ref{eq:anonunif}) and (\ref{eq:anew}), 
1583: we immediately conclude that with a non-uniform 
1584: spin-orbit scattering strength of the form (\ref{eq:nonunif})
1585: with $\lambda^{\rm f} =
1586: \lambda^{\rm f}_{ 1} = \lambda^{\rm f}_{ 2}$
1587: one should make the identification
1588: \begin{equation}
1589:   a^2 =  
1590:   \left( \frac{L^2}{4 (\lambda^{\rm f})^2} \right)
1591:   \frac{v_F l \pi}{12 L^2 \Delta}.
1592: \end{equation}
1593: In terms of the energy scale $\varepsilon_{\Vert}^{\rm so}$, this
1594: implies
1595: \begin{equation}
1596:   \varepsilon_{\Vert}^{\rm so}
1597:   = \frac{1}{6} E_{\rm Th} \left( \frac{L^2}{4(\lambda^{\rm f})^2} 
1598: \right),
1599: \end{equation}
1600: in agreement with the general relation (\ref{eq:kappageneral}).
1601: 
1602: \begin{thebibliography}{99}
1603: \bibitem{mesoreview1} C. W. J. Beenakker and H. van Houten, Solid State
1604: Phys. \textbf{44}, 1 (1991).
1605: 
1606: \bibitem{Hikami} S. Hikami, A. I. Larkin, and Y. Nagaoka, Prog. Theor. Phys. 
1607: \textbf{63}, 707 (1980).
1608: 
1609: \bibitem{Bergmann} G. Bergmann, Phys. Rep. \textbf{107}, 1 (1984).
1610: 
1611: \bibitem{Khaetskii} A. V. Khaetskii and Yu. V. Nazarov, Phys. Rev. B 
1612: \textbf{61}, 012639 (2000).
1613: 
1614: \bibitem{Folk} J. A. Folk, S. R. Patel, K. M. Birnbaum, C. M. Marcus, C. I.
1615: Duru\"oz, and J. S. Harris, Jr., Phys. Rev. Lett. \textbf{86}, 2102 (2001).
1616: 
1617: \bibitem{Halperin} B. I. Halperin, A. Stern, Y. Oreg, J. N. H. J. Cremers,
1618: J. A. Folk, and C. M. Marcus, Phys. Rev. Lett. \textbf{86}, 2106 (2001).
1619: 
1620: \bibitem{AF} I. L. Aleiner and V. I. Fal'ko, Phys. Rev. Lett. \textbf{87},
1621: 256801 (2001); {\bf 89}, 079902(E) (2002).
1622: 
1623: \bibitem{RashbaDresselhaus} In conventional notation, the Hamiltonian is
1624: specified in coordinates with respect to the $\hat x = [100]$ and $\hat y =
1625: [010]$ directions. Then, the spin-orbit part of the Hamiltonian takes the
1626: form 
1627: $$
1628:   H_{\mathrm{so}} = \frac{\gamma}{m}(p_x \sigma_y - p_y \sigma_x) + 
1629:   \frac{\eta}{m} (p_x \sigma_x - p_y \sigma_y).
1630: $$
1631: Here $\gamma$ and $\eta$ are coupling constants for the Rashba and
1632: Dresselhaus terms, respectively. In terms of the rates $\gamma$ and $\eta$,
1633: one has $\hbar/2\lambda_1 = -\eta + \gamma$, $\hbar/2\lambda_2 = -\eta -
1634: \gamma$.
1635: 
1636: 
1637: \bibitem{Meir} Y. Meir, Y. Gefen, and O. Entin-Wohlman, 
1638: Phys. Rev. Lett. {\bf 63}, 798 (1989).
1639: 
1640: \bibitem{MathurStone} H. Mathur and A. D. Stone,
1641:   Phys. Rev. Lett. \textbf{68}, 
1642: 2964 (1992).
1643: 
1644: \bibitem{Oreg} Y. Oreg and O. Entin-Wohlman Phys. Rev. B \textbf{46}, 2393
1645: (1992).
1646: 
1647: \bibitem{Beenakker97} For reviews, see: C. W. J. Beenakker, Rev. Mod. Phys. 
1648: \textbf{69}, 731 (1997) and Y. Alhassid, Rev. Mod. Phys. \textbf{72}, 895
1649: (2000).
1650: 
1651: \bibitem{Shayegan} J. P. Lu, J. B. Yau, S. P. Shukla, M. Shayegan, L.
1652: Wissinger, U. R\"ossler, and R. Winkler, Phys. Rev. Lett. \textbf{81}, 1282
1653: (1998).
1654: 
1655: \bibitem{Marcus} J. B. Miller, D. M. Zumb\"uhl, C. M. Marcus, Y. B.
1656: Lyanda-Geller, D. Goldhaber-Gordon, K. Campman, A. C. Gossard,
1657: Phys. Rev. Lett. {\bf 90}, 076807 (2003).
1658: 
1659: \bibitem{Zumbuehl} D. M. Zumb\"uhl, J. B. Miller, C. M. Marcus, K. Campman,
1660: and A. C. Gossard, Phys. Rev. Lett. {\bf 89}, 276803 (2002).
1661: 
1662: \bibitem{BCH} P. W. Brouwer, J. N. H. J. Cremers, and B. I. Halperin, Phys.
1663: Rev. B \textbf{65}, 081302 (2002).
1664: 
1665: \bibitem{FJ} V. I. Fal'ko and T. Jungwirth, Phys. Rev. B \textbf{65}, 81306
1666: (2002).
1667: 
1668: \bibitem{Mehta} M.L. Mehta, \emph{Random Matrices}, (Academic, New York,
1669: 1991).
1670: 
1671: \bibitem{WavesRM} P. W. Brouwer, K. M. Frahm, and C. W. J. Beenakker, Waves
1672: in Random Media 9, 91 (1999) [cond-mat/9809022].
1673: 
1674: \bibitem{diagrams} P. W. Brouwer and C. W. J. Beenakker, J. Math. Phys. 
1675: \textbf{37}, 4904 (1996).
1676: 
1677: \bibitem{SimonsAltshuler} B. L. Altshuler and B. D. Simons in 
1678: \emph{Mesoscopic Quantum Physics}, E. Akkermans, G. Montambaux,
1679: J.-L. Pichard, and
1680: J. Zinn-Justin, eds, (North Holland, Amsterdam, 1995).
1681: 
1682: \bibitem{BuettikerVoltage} M. B\"uttiker, Phys. Rev. B \textbf{33}, 3020
1683: (1986); IBM J. Res. Dev. \textbf{32}, 63 (1988).
1684: 
1685: \bibitem{BM} H. U. Baranger and P. A. Mello, Phys. Rev. B \textbf{51}, 4703
1686: (1995).
1687: 
1688: \bibitem{BB} P. W. Brouwer and C. W. J. Beenakker, Phys. Rev. B \textbf{55},
1689: 4695 (1997).
1690: 
1691: \bibitem{Argaman} N. Argaman, Phys. Rev. Lett. \textbf{75}, 2750 (1995).
1692: 
1693: \bibitem{thesis} J. N. H. J. Cremers, Ph. D. thesis (Harvard, 2002).
1694: 
1695: \end{thebibliography}
1696: 
1697: \end{document}
1698: