1: %\documentclass[showpacs,preprint,amsmath,amssymb]{revtex4}
2: \documentclass[twocolumn,prb,showpacs]{revtex4}
3:
4:
5: %\documentclass[showpacs,preprint,amsmath,amssymb]{revtex4}
6: %\documentclass[showpacs,preprintnumbers,amsmath,amssymb,twocolumn,prb]{revtex4}
7: %\newcommand{\Sc}{Schr\"odinger }
8: % \newcommand{\al}{\alpha }
9: % \newcommand{\vfi}{\varphi }
10:
11: %\ifx\pdftexversion\undefined
12: % \usepackage[dvips]{graphics}
13: %\else
14: % \usepackage[pdftex]{graphics}
15: %\fi
16:
17: %\usepackage{graphicx}
18:
19:
20:
21: \usepackage{graphicx}
22: \usepackage{epsfig}
23: \usepackage{dcolumn}
24: \usepackage{bm}
25: \usepackage{amsmath}
26:
27: \begin{document}
28: \title{Hall Voltage with the Spin Hall Effect}
29: \author{Yuriy V. Pershin}
30: \email{pershin@physics.ucsd.edu}
31: \author{Massimiliano Di Ventra}
32: \email{diventra@physics.ucsd.edu} \affiliation{Department of
33: Physics, University of California, San Diego, La Jolla, California
34: 92093-0319}
35:
36: \begin{abstract}
37: The spin Hall effect does not generally result in a charge Hall
38: voltage. We predict that in systems with inhomogeneous electron
39: density in the direction perpendicular to main current flow, the
40: spin Hall effect is instead accompanied by a Hall voltage. Unlike
41: the ordinary Hall effect, we find that this Hall voltage is
42: quadratic in the longitudinal electric field for a wide range of
43: parameters accessible experimentally. We also predict spin
44: accumulation in the bulk and sharp peaks of spin-Hall induced charge
45: accumulation near the edges. Our results can be readily tested
46: experimentally, and would allow the electrical measurement of the
47: spin Hall effect in non-magnetic systems and without injection of
48: spin-polarized electrons.
49: \end{abstract}
50:
51: \pacs{72.25.Dc, 71.70.Ej}
52:
53: \maketitle
54:
55: Currently, much attention is given to studies of the spin Hall
56: effect, which allows to polarize electron spins without magnetic
57: fields and/or magnetic materials
58: \cite{natureSHC,r0,r1,r2,r3,Rashba,Rashba1,r4,r5,r6,r7,r8,r9,r10,Kato,Wunderlich,Valenzuela}.
59: In the spin Hall effect, electrically induced electron spin
60: polarization accumulates near the edges of a channel and is zero
61: in its central region. This effect is caused by deflection of
62: carriers moving along an applied electric field by
63: extrinsic\cite{r1} and/or intrinsic\cite{r3} mechanisms. In a
64: non-magnetic homogeneous system, spin accumulation is {\em not}
65: accompanied by a charge Hall voltage because two spin Hall
66: currents cancel each other \cite{natureSHC}. The absence of Hall
67: voltage leads to difficulties in probing the spin Hall effect,
68: since measuring a charge accumulation is much easier than
69: measuring a spin accumulation. Recently, the spin Hall effect has
70: been observed both optically \cite{Kato,Wunderlich} and
71: electrically \cite{Valenzuela}. In the latter case, a charge
72: accumulation has been created through injection of spin-polarized
73: electrons into the sample \cite{Valenzuela}.
74:
75: In the present Letter, we predict that in a system with an
76: inhomogeneous electron density profile in the direction
77: perpendicular to the direction of main current flow, the spin Hall
78: effect results in {\em both} spin and charge accumulations. The
79: pattern of charge accumulation is determined by the interplay of two
80: mechanisms. The first mechanism of charge accumulation is based on
81: the dependence of spin-up and spin-down currents on local spin-up
82: and spin-down densities. Spin currents, outgoing from regions with
83: higher densities, are not fully compensated by incoming currents,
84: therefore, a charge accumulation appears. This mechanism is
85: primarily responsible for the non-zero Hall voltage. The second
86: mechanism of charge accumulation is related to scattering of spin
87: currents on sample boundaries which act like obstacles. Like in the
88: case of Landauer resistivity dipoles \cite{Landauer}, this
89: scattering leads to formation of local charge accumulation, which is
90: also expected in traditionally-studied spin Hall systems. In
91: addition, we show that in systems with inhomogeneous electron
92: density the spin accumulation appears not only near the sample
93: boundaries, but also in the bulk. Our approach does not involve any
94: use of magnetic materials and fields, therefore, the spin Hall
95: effect can be measured electrically in completely non-magnetic
96: system and without injection of spin-polarized electrons.
97:
98: To illustrate this effect, let us begin by considering a system
99: having a step profile of electron density, as shown in Fig.
100: \ref{fig1}. There are several possible ways to fabricate such a
101: system including density depletion by an electrode, inhomogeneous
102: doping \cite{YP}, or variation of the sample height. What is
103: important for us is that the perpendicular (in $y$ direction) spin
104: currents are different in the regions with different electron
105: density. Then, if we consider currents passing through the boundary
106: separating regions with different charge densities ($n_1$ and
107: $n_2$), it is clear that the spin current from the region with
108: higher electron density has a larger magnitude than the current in
109: the reverse direction. The difference in currents implies charge
110: transfer through the boundary and formation of a dipole layer.
111:
112: \begin{figure}[b]
113: \includegraphics[angle=270,width=7.5cm]{fig1}
114: \caption{\label{fig1} (Color online) Spin Hall effect in a system
115: with a step profile of the electron density in $y$ direction,
116: $n_1>n_2$. Spin currents through the boundary between $n_1$ and
117: $n_2$ do not cancel each other, resulting in a Hall voltage.}
118: \end{figure}
119:
120: Let us now provide a quantitative analysis of this effect. We employ
121: a two-component drift-diffusion model \cite{flatte,PDV}, and in
122: order to find a self-consistent solution, we supplement the
123: drift-diffusion equations with the Poisson equation. In our
124: drift-diffusion calculation scheme, the inhomogeneous charge density
125: profile $n(y)$ is defined via an assigned positive background
126: density profile $N(y)$ (such as the one in Fig.~\ref{fig1}), which,
127: as discussed above, can be obtained in different ways. Assuming
128: homogeneous charge and current densities in $x$ direction and
129: homogeneous $x$-component of the electric field in both $x$ and $y$
130: directions, we can write a set of equations including only $y$ and
131: $t$ dependencies:
132:
133: \begin{equation}
134: e\frac{\partial n_{\uparrow (\downarrow)}}{\partial
135: t}=\textnormal{div} j_{y,\uparrow
136: (\downarrow)}+\frac{e}{2\tau_{sf}}\left(n_{\downarrow
137: (\uparrow)}-n_{\uparrow (\downarrow)} \right), \label{contEq}
138: \end{equation}
139: \begin{equation}
140: j_{y,\uparrow (\downarrow)}=\sigma_{\uparrow (\downarrow)}
141: E_y+eD\nabla n_{\uparrow (\downarrow)}\pm \gamma I_{x,\uparrow
142: (\downarrow)} , \label{currentEq}
143: \end{equation}
144: and
145: \begin{equation}
146: \textnormal{div}E_y=\frac{e}{\varepsilon\varepsilon_0}\left(
147: N(y)-n\right), \label{puaeq}
148: \end{equation}
149: where $-e$ is the electron charge, $n_{\uparrow (\downarrow)}$ is
150: the density of spin-up (spin-down) electrons, $j_{y,\uparrow
151: (\downarrow)}$ is the current density, $\tau_{sf}$ is the spin
152: relaxation time, $\sigma_{\uparrow (\downarrow)}=en_{\uparrow
153: (\downarrow)}\mu$ is the spin-up (spin-down) conductivity, $\mu$ is
154: the mobility, $D$ is the diffusion coefficient, $\epsilon$ is the
155: permittivity of the bulk, and $\gamma$ is the parameter describing
156: deflection of spin-up (+) and spin-down (-) electrons. The current
157: $I_{x,\uparrow (\downarrow)}$ in $x$-direction is coupled to the
158: homogeneous electric field $E_0$ in the same direction as
159: $I_{x,\uparrow (\downarrow)}=en_{\uparrow (\downarrow)}\mu E_0$. The
160: last term in Eq. (\ref{currentEq}) is responsible for the spin Hall
161: effect.
162:
163: Equation (\ref{contEq}) is the continuity relation that takes into
164: account spin relaxation, Eq. (\ref{currentEq}) is the expression for
165: the current in $y$ direction which includes drift, diffusion and
166: spin Hall effect components, and Eq. (\ref{puaeq}) is the Poisson
167: equation. It is assumed that $D$, $\mu$, $\tau_{sf}$ and $\gamma$
168: are equal for spin-up and spin-down electrons.~\cite{prec} In our
169: model, as it follows from Eq. (\ref{currentEq}), the spin Hall
170: correction to spin-up (spin-down) current (the last term in Eq.
171: \ref{currentEq}) is simply proportional to the local spin-up
172: (spin-down) density. All information about microscopic mechanisms
173: for the spin Hall effect is therefore lumped in the parameter
174: $\gamma$.
175:
176: Combining equations (\ref{contEq}) and (\ref{currentEq}) for
177: different spin components we can get the following equations for
178: electron density $n=n_{\uparrow}+n_{\downarrow}$ and spin density
179: imbalance $P=n_{\uparrow}-n_{\downarrow}$:
180: \begin{equation} \frac{\partial
181: n}{\partial t}=\frac{\partial}{\partial y} \left[ \mu n E_y + D
182: \frac{\partial n}{\partial y} +\gamma P \mu E_0 \right] \label{CC}
183: \end{equation}
184: and
185: \begin{equation} \frac{\partial P}{\partial t}=\frac{\partial}{\partial y}
186: \left[ \mu P E_y + D \frac{\partial P}{\partial y} +\gamma n \mu
187: E_0 \right]-\frac{P}{\tau_{sf}}. \label{Peq}
188: \end{equation}
189:
190: {\it Analytical solution --} Before solving Eqs.
191: (\ref{puaeq})-(\ref{Peq}) numerically, let us try to find analytical
192: solutions in specific cases. This will help us in the discussion of
193: the numerical results. An analytical steady-state solution of these
194: equations can indeed be found for the case of exponential density
195: profile in a system which is infinite in the $y$ direction.
196:
197: The structure of Eqs. (\ref{puaeq})-(\ref{Peq}) allows us to
198: select a solution in the form
199: \begin{eqnarray}
200: n=N(y)=Ae^{\alpha y} \label{s1}, \\ P=Ce^{\alpha y},
201: \\E_y=\textnormal{const},\label{s2}
202: \end{eqnarray}
203: where $A$, $C$ and $\alpha$ are constants ($A$ and $\alpha$ are
204: assigned). This solution corresponds to constant spin polarization
205: $p=P/n$. Substituting Eqs. (\ref{s1})-(\ref{s2}) into Eqs.
206: (\ref{CC}) and (\ref{Peq}) (note that the Poisson equation
207: (\ref{puaeq}) is automatically satisfied) we obtain
208: \begin{eqnarray}
209: \mu E_y A+D\alpha A+\gamma \mu E_0C=0,\label{eq123}
210: \\ \mu E_y \alpha
211: C+D\alpha^2C+\gamma \mu E_0 \alpha A-\frac{C}{\tau_{sf}}=0.
212: \end{eqnarray}
213: From these equations, eliminating $E_y$, we find
214: \begin{equation}
215: C=\frac{-1\pm\sqrt{1+\left( 2\tau_{sf}\gamma \mu E_0 \alpha
216: \right)^2}}{2\tau_{sf}\gamma \mu E_0 \alpha}A. \label{ccoeff}
217: \end{equation}
218: The physical solution corresponds to the + sign in Eq.
219: (\ref{ccoeff}). It can be easily verified that the solution given by
220: Eqs. (\ref{s1})-(\ref{s2}), (\ref{ccoeff}) corresponds to $j_y=0$.
221: Substituting Eq. (\ref{ccoeff}) into Eq. (\ref{eq123}) we finally
222: get
223: \begin{equation}
224: E_y=-\frac{D}{\mu}\alpha-\frac{-1+\sqrt{1+\left( 2\tau_{sf}\gamma
225: \mu E_0 \alpha \right)^2}}{2\tau_{sf} \mu \alpha}. \label{Ey}
226: \end{equation}
227: The second term on the RHS of Eq. (\ref{Ey}) is the electric field
228: needed to compensate the transverse current arising due to the spin
229: Hall effect. If we now assume the sample has a finite (but large)
230: width $L$, then, this term can be interpreted as due to charge
231: accumulation near the edges, as in the ordinary Hall effect.
232: Therefore, the Hall voltage can be written as
233:
234: \begin{eqnarray}
235: V_H\simeq L \frac{-1+\sqrt{1+\left( 2\tau_{sf}\gamma \mu E_0
236: \alpha \right)^2}}{2\tau_{sf} \mu \alpha}\approx \nonumber \\
237: \approx \left\{
238: \begin{array}{cc}
239: L\tau_{sf} \mu \alpha\gamma^2E_0^2,& 2\tau_{sf}\gamma \mu E_0
240: \alpha \ll 1 \\ \\ L\gamma E_0,& 2\tau_{sf}\gamma \mu E_0 \alpha
241: \gg 1
242: \end{array}
243: \right. . \label{VH}
244: \end{eqnarray}
245: From this equation we see that the Hall voltage is quadratic in
246: $E_0$ for small values of the parameter $2\tau_{sf}\gamma \mu E_0
247: \alpha$, and linear in $E_0$ for large values of this parameter. In
248: fact, the quadratic dependence is quite unusual, since in the
249: ordinary Hall effect the Hall voltage is linear in the longitudinal
250: current. The reason for this unusual dependence can be understood as
251: follows. The charge current in the $y$ direction, determined by the
252: difference of spin-up and spin-down currents, has a component
253: (related to the last term in Eq. (\ref{currentEq})) proportional to
254: the spin density imbalance $P$ times $\gamma E_0$. At small values
255: of $2\tau_{sf}\gamma \mu E_0 \alpha$, the spin density imbalance is
256: proportional to $\gamma E_0$ itself. Therefore, the charge current
257: and Hall voltage are quadratic in $E_0$. At large values of
258: $2\tau_{sf}\gamma \mu E_0 \alpha$, the spin density imbalance
259: saturates and the current dependence on $E_0$ becomes linear.
260: Another difference with respect to the ordinary Hall effect is that
261: the polarity of the Hall voltage in the spin Hall effect is fixed by
262: the geometry of the structure, and does not depend on the direction
263: of the longitudinal current.
264:
265: Let us now estimate the magnitude of $2\tau_{sf}\gamma \mu E_0
266: \alpha$. Taking parameters related to experiments on GaAs
267: ($\tau_{sf}=10$ns, $\gamma=10^{-3}$(Ref.~\onlinecite{Rashba}),
268: $\mu=8500$cm$^2$/(Vs), $E_0=100$V/cm, $\alpha=2/L$, $L=100\mu$m), we
269: find $2\tau_{sf}\gamma \mu E_0 \alpha=3.4\cdot10^{-3}$. Therefore,
270: in experiments with GaAs, most likely, a quadratic Hall voltage
271: dependence on the longitudinal electric field can be observed.
272:
273: \begin{figure}[t]
274: \includegraphics[angle=270,width=8.5cm]{fig2a}
275: \includegraphics[angle=270,width=8.5cm]{fig2b}
276: \caption{\label{fig2}(Color online) Distributions of the electron
277: density $n(y)$ and spin density imbalance
278: $P(y)=n_{\uparrow}-n_{\downarrow}$ for a step (a) and exponential
279: (b) background density profiles. The plots presented in the paper
280: were obtained using the parameter values
281: $\mu=8500$cm$^2$/(Vs), $D=55$cm$^2$/s, $\varepsilon=12.4$,
282: $\tau_{sf}=10$ns, $\gamma=10^{-3}$, $E_0=100$V/cm and the background density
283: profiles: (a) $N=10^{16}(1+\theta(y-L/2))$cm$^{-3}$ and (b)
284: $N=10^{16}\textnormal{exp}(2y/L)$cm$^{-3}$ , where $\theta (..)$ is the
285: step function, and $L=100\mu $m is the sample width.}
286: \end{figure}
287:
288: {\it Numerical solution --} Equations (\ref{puaeq})-(\ref{Peq}) can
289: be solved numerically for any reasonable form of $N(y)$. We choose
290: for their simplicity (and possibility to be realized in practice) a
291: step profile and an exponential profile. We solve these equations
292: iteratively, starting with the electron density $n(y)$ close to
293: $N(y)$ and $P(y)$ close to zero and recalculating $E_y(y)$ at each
294: time step.~\cite{prec1} Once the steady-state solution is obtained,
295: the Hall voltage as a function of $E_0$ is calculated as a change of
296: the electrostatic potential across the sample.
297:
298: \begin{figure}[t]
299: \includegraphics[angle=270,width=8.5cm]{fig3}
300: \caption{\label{fig3}(Color online) Variations in the transverse
301: charge density induced by the longitudinal current. Here, $\delta
302: n=n(E_0=100\textnormal{V/cm})-n(E_0=0)$. The curve for the
303: exponential profile has been shifted vertically by $3\cdot
304: 10^{10}$cm$^{-3}$ for clarity. The dashed lines corresponding to
305: $\delta n=0$ are there to guide the eye.}
306: \end{figure}
307:
308: Fig. \ref{fig2} shows distributions of the charge density and spin
309: density imbalance in systems with a step (panel (a) of Fig.
310: \ref{fig2}) and exponential (panel (b) of Fig. \ref{fig2})
311: background densities. The values of parameters used for these
312: particular simulations were selected to be close to experimental
313: conditions reported in Ref. \onlinecite{Kato}. However, we have
314: tested the robustness of our predictions by solving Eqs.
315: (\ref{puaeq})-(\ref{Peq}) for different values of parameters, and
316: found that the predicted Hall voltage should be measurable under a
317: wide range of experimental parameters. Quite generally, the
318: self-consistent charge density $n(y)$ is very close to the
319: background density $N(y)$. Small deviations of $n(y)$ from $N(y)$
320: can be observed in regions with strong gradients of $N(y)$. In
321: particular, we can notice that the step profile of electron
322: density in Fig. \ref{fig2}(a) is smoothed out. Such a charge
323: redistribution is related to the diffusion term in Eq. (\ref{CC}).
324: The charge diffusion leads to the formation of a built-in electric
325: field that equilibrates the charge diffusion.
326:
327: We also find that the induced spin density imbalance $P$ in systems
328: with inhomogeneous electron densities shows some new features, in
329: addition to the well-known spin accumulation near the edges. For
330: instance, in Fig. \ref{fig2}(a), $P$ has an additional peak around
331: $y=50\mu$m. In Fig. \ref{fig2}(b), $P$ is almost constant in the
332: central region of the sample. In both cases, the physics of non-zero
333: spin density imbalance is the same: the spin-current incoming from
334: the right is stronger than the spin-current incoming from the left.
335: We note that the integral spin density imbalance is always zero.
336:
337: At $E_0=0$, the system is spin-unpolarized and there is no Hall
338: voltage. When the longitudinal current is switched on, the electron
339: charge redistributes, and the associated Hall voltage appears. The
340: change of electron density due to the spin Hall effect is presented
341: in Fig. \ref{fig3}. The first interesting observation is that there
342: is a strong charge accumulation near the edges followed by a charge
343: depletion region. Another observation is that the total electron
344: density in the left region of the samples ($y<50\mu$m) has increased
345: and, correspondingly, the total electron density in the right region
346: has decreased. This change of the electron distribution can be seen
347: in Fig. \ref{fig3}. Therefore, the left part of the samples is
348: charged negatively and the right part is charged positively, as
349: schematically shown in Fig. \ref{fig1}.
350:
351: The mechanism of formation of sharp peaks of charge accumulation
352: near the edges is similar to the mechanism of formation of Landauer
353: resistivity dipoles \cite{Landauer}. From the point of view of spin
354: currents, the sample edges act as obstacles which block the current
355: flow, and lead to charge accumulation. The adjacent regions with the
356: depleted electron density can be interpreted as screening clouds. We
357: stress that this Landauer-type dipoles of charge accumulation are
358: quite general for spin Hall systems, and should thus be present also
359: in traditionally studied structures with a constant density profile.
360:
361: \begin{figure}[t]
362: \includegraphics[angle=270,width=8.5cm]{fig4}
363: \caption{\label{fig4}(Color online) Hall voltage as a function of
364: the longitudinal electric field $E_0$.}
365: \end{figure}
366:
367: We finally plot in Fig. \ref{fig4} the change of the electrostatic
368: potential across the sample as a function of longitudinal electric
369: field. The Hall voltage, for both density profiles, has a
370: dependence on $E_0$ which is very close to the exponential
371: dependence we have predicted analytically in Eq.~\ref{VH} for
372: small values of $2\tau_{sf}\gamma \mu E_0 \alpha$. The fact that
373: this exponential dependence appears also in the step profile,
374: hints at a possible ``general'' property of the Hall voltage in
375: spin Hall systems with inhomogeneous densities. We emphasize that
376: a Hall voltage should also appear in spin Hall systems with a
377: homogeneous electron density, but inhomogeneous $\gamma$. This
378: corresponds to the case in which the spin-orbit coupling is
379: dependent on space\cite{serra,nitta}.
380:
381: In conclusion, we have shown that a Hall voltage would appear in
382: spin Hall systems with inhomogeneous electron density in the
383: direction perpendicular to main current flow. The striking result is
384: that this Hall voltage is generally quadratic in the longitudinal
385: electric field, unlike the ordinary Hall voltage which is linear in
386: the same field. These results can be easily verified experimentally,
387: and would simplify tremendously the measurement of the spin Hall
388: effect by allowing an electrical measurement of the latter in
389: non-magnetic systems, and without injection of spin-polarized
390: electrons.
391:
392: This work is partly supported by the NSF Grant No. DMR-0133075.
393:
394: \begin{thebibliography}{}
395:
396: \bibitem{natureSHC} J. Inoue and H. Ohno, Science {\bf 309}, 2004
397: (2005).
398:
399: \bibitem{r0} M. I. D'yakonov and V. I. Perel, Phys. Lett. A {\bf 35}, 459
400: (1971).
401:
402: \bibitem{r1} J. E. Hirsch, Phys. Rev. Lett. {\bf 83}, 1834 (1999).
403:
404: \bibitem{r2} S. Murakami, N. Nagaosa, S.-C. Zhang, Science {\bf 301}, 1348
405: (2003).
406:
407: \bibitem{r3} J. Sinova, D. Culcer, Q. Niu, N. A. Sinitsyn, T. Jungwirth,
408: and A. H. MacDonald, Phys. Rev. Lett. {\bf 92}, 126603 (2004).
409:
410: \bibitem{Rashba} H.-A. Engel, B. I. Halperin, and E. I. Rashba
411: Phys. Rev. Lett. {\bf 95}, 166605 (2005).
412:
413: \bibitem{Rashba1} H.-A. Engel, E. I. Rashba, and B. I. Halperin
414: Phys. Rev. Lett. {\bf 98}, 036602 (2007).
415:
416: \bibitem{r4} J. Inoue, G. E. W. Bauer, and L. W. Molenkamp, Phys. Rev. B {\bf 70}, 041303(R) (2004).
417:
418: \bibitem{r5} B. A. Bernevig and S.-C. Zhang, Phys. Rev. Lett. {\bf 95}, 016801 (2005).
419:
420: \bibitem{r6} M. W. Wu and J. Zhou, Phys. Rev. B {\bf 72}, 115333
421: (2005).
422:
423: \bibitem{r7} E. M. Hankiewicz, G. Vignale, and M. E. Flatt\'e, Phys. Rev. Lett. {\bf 97}, 266601
424: (2006).
425:
426: \bibitem{r8} B. K. Nikoli\'c, S. Souma, L. P. Z\^arbo, and J.
427: Sinova, Phys. Rev. Lett. {\bf 95}, 046601 (2005).
428:
429: \bibitem{r9} J. I. Inoue, G. E. W. Bauer, L. W. Molenkamp, Phys. Rev. B {\bf 70}, 041303(R)
430: (2004).
431:
432: \bibitem{r10} A. G. Mal'shukov and C. S. Chu, Phys. Rev. Lett. {\bf 97}, 076601
433: (2006).
434:
435: \bibitem{Kato} Y. K. Kato, R. C. Myers, A. C. Gossard, D. D. Awschalom,
436: Science {\bf 306}, 1910 (2004).
437:
438: \bibitem{Wunderlich} J. Wunderlich, B. Kaestner, J. Sinova, and T.
439: Jungwirth, Phys. Rev. Lett. {\bf 94}, 047204 (2005).
440:
441: \bibitem{Valenzuela} S. O. Valenzuela and M. Tinkham, Nature {\bf 442}, 176
442: (2006).
443:
444: \bibitem{Landauer} R. Landauer, IBM J. Res. Develop. {\bf 1}, 223
445: (1957).
446:
447: \bibitem{YP} Y. V. Pershin and V. Privman, Phys. Rev. Lett. {\bf 90}, 256602
448: (2003).
449:
450: \bibitem{flatte} Z. G. Yu and M. E. Flatt\'e, Phys. Rev. B {\bf 66}, 201202 (2002).
451:
452: \bibitem{PDV} Y. V. Pershin and M. Di Ventra, cond-mat/0701678.
453:
454: \bibitem{prec} This is a good approximation for the range of
455: parameters considered in this work.
456:
457: \bibitem{prec1} We have employed the Scharfetter-Gummel discretization scheme
458: [D. L. Scharfetter and H. K. Gummel, IEEE. Trans. Electron.
459: Devices, {\bf ED-16}, 64 (1969)] to solve both Eqs. (\ref{CC}) and
460: (\ref{Peq}) numerically.
461:
462: \bibitem{serra} M. Val\'{\i}n-Rodr\'{\i}guez, A. Puente, and L. Serra
463: Phys. Rev. B {\bf 69}, 153308 (2004).
464:
465: \bibitem{nitta} J. Nitta, T. Akazaki, and H. Takayanagi, Phys. Rev. Lett. {\bf 78}, 1335 (1997).
466:
467: \end{thebibliography}
468:
469: \end{document}
470: