0807.3067/ddt.tex
1: %% LyX 1.5.6 created this file.  For more info, see http://www.lyx.org/.
2: %% Do not edit unless you really know what you are doing.
3: \documentclass[english,twocolumn,showpacs,superscriptaddress,prl]{revtex4}
4: \usepackage[T1]{fontenc}
5: \usepackage[latin9]{inputenc}
6: \usepackage{amsmath}
7: \usepackage{graphicx}
8: \usepackage{amssymb}
9: \usepackage{babel}
10: 
11: \begin{document}
12: 
13: \title{Spin field effect transistors with ultracold atoms}
14: 
15: 
16: \date{\today{}}
17: 
18: 
19: \author{J. Y. Vaishnav}
20: 
21: 
22: \affiliation{Joint Quantum Institute, National Institute of Standards and Technology,
23: Gaithersburg MD 20899 USA}
24: 
25: 
26: \author{Julius Ruseckas}
27: 
28: 
29: \affiliation{Institute of Theoretical Physics and Astronomy of Vilnius University,
30: A. Go\v{s}tauto 12, Vilnius 01108, Lithuania}
31: 
32: 
33: \author{Charles W. Clark }
34: 
35: 
36: \affiliation{Joint Quantum Institute, National Institute of Standards and Technology,
37: Gaithersburg MD 20899 USA}
38: 
39: 
40: \author{Gediminas Juzel\=unas}
41: 
42: 
43: \affiliation{Institute of Theoretical Physics and Astronomy of Vilnius University,
44: A. Go\v{s}tauto 12, Vilnius 01108, Lithuania}
45: 
46: \begin{abstract}
47: We propose a method of constructing cold atom analogs of the spintronic
48: device known as the Datta-Das transistor (DDT), which despite its
49: seminal conceptual role in spintronics, has never been successfully
50: realized with electrons. We propose two alternative schemes for an
51: atomic DDT, both of which are based on the experimental setup for
52: tripod stimulated Raman adiabatic passage. Both setups involve atomic
53: beams incident on a series of laser fields mimicking the relativistic
54: spin orbit coupling for electrons that is the operating mechanism
55: of the DDT.
56: \end{abstract}
57: 
58: \pacs{37.10.Vz, 37.10.Jk,85.75.Hh}
59: 
60: \maketitle
61: The emerging technology of semiconductor spintronics exploits the
62: electron's spin degree of freedom, as well as its charge state. The
63: first scheme for a semiconductor spintronic device was a spin field-effect
64: transistor known as the Datta-Das transistor (DDT) (Fig.~\ref{fig:ddtschematic}a)
65: \citep{dattadas}. The eighteen years since the theoretical proposal
66: have seen numerous experimental efforts to construct the DDT. Various
67: experimental obstacles, such as difficulties in spin injection, stray
68: electric fields and insufficient quality of spin-orbit coupling, have
69: prevented successful implementation of the DDT \citep{zutic2004sfa}.
70: 
71: Cold atom systems, in contrast with their electronic counterparts, are
72: highly controllable and tunable. This suggests the possibility of
73: designing precise atomic analogs of electronic systems which, due
74: either to fundamental physical limits or technological difficulties,
75: are experimentally inaccessible in their original manifestations.
76: The idea grows out of recent interest in {}``atomtronics,'' or building
77: cold atom analogs of ordinary electronic materials, devices and circuits
78: \citep{Atom-Diode,seaman:023615,experimentaldiode}. In particular,
79: an atom diode has been proposed \citep{Atom-Diode} and realized \citep{experimentaldiode}. 
80: 
81: In this Letter, we identify a method for constructing a cold atom
82: analog of a Datta-Das transistor. The setup is based on a four level
83: {}``tripod'' scheme of atom-light coupling \citep{unanyan1998rca,unanyan99pra,theuer1999nlc,stirap,ruseckas-2005-95,lietfiz-2007}
84: involving three atomic ground states and one excited state (see Fig.~\ref{fig:ddtschematic}b).
85: Such tripod schemes are an extension of the usual three-level $\Lambda$-type
86: setup for stimulated Raman adiabatic passage (STIRAP) \citep{Atom-Diode,spielman-2008},
87: and are experimentally accessible in metastable Ne, $^{87}\mbox{Rb}$
88: and a number of other gases \citep{theuer1999nlc,stirap}. The proposed
89: device provides a robust method for atomic state manipulation that
90: is immune to the inhomogeneities intrinsic to programmed Rabi pulses. 
91: 
92: %
93: \begin{figure}
94: \begin{centering}
95: \includegraphics[width=0.8\columnwidth]{ddtschematic}
96: \par\end{centering}
97: 
98: \caption{(a) Schematic of a DDT. {}``S'' and {}``D'' are ferromagnetic
99: source and drain electrodes. In between is a semiconducting gate region,
100: where the spin precesses by an amount which depends periodically on
101: the tunable gate voltage $V_{g}.$ This precession results in a controllable
102: current modulation at $D.$ (b) A tripod scheme of atomic energy levels,
103: coupled by laser fields with Rabi frequencies $\Omega_{i}$. (c,d)
104: Two alternative setups for an atomic version of the DDT. Here, the
105: source is a state-polarized atomic beam (blue), the gate is the intersection
106: region of a configuration of laser beams (red), and the drain is an
107: atomic state analyzer (green). }
108: 
109: 
110: \label{fig:ddtschematic}
111: \end{figure}
112: 
113: 
114: The source terminal of an electronic DDT (Fig.~\ref{fig:ddtschematic}a)
115: is a ferromagnetic electrode that emits spin-polarized electrons.
116: The DDT drain terminal, a ferromagnetic analyzer, acts as a spin filter.
117: Between source and drain is a semiconducting gate region, in which
118: the gate-induced electric field produces a Rashba spin-orbit coupling
119: \citep{Rashba60} for electrons. While passing through the gate region,
120: the electron's spin precesses; the electron emerges at the drain having
121: undergone a spin rotation which is tunable via the gate voltage. Since
122: the drain passes only a certain spin direction, the drain current
123: is an oscillating function of the gate voltage.
124: 
125: Our atomic analog of the DDT (Figs.~\ref{fig:ddtschematic}c,d) uses
126: a beam of atoms in place of electrons. The two dark states in the
127: tripod setup play the role of the electron's spin states, and the
128: {}``source'' is a dilute atomic beam. The {}``gate'' region consists
129: of crossed laser beams engineered to mimic Rashba or Rashba-like spin
130: orbit couplings \citep{galitski,Jacob07,jayzb,prar08,juz08-neg-refl};
131: the analog of the gate voltage can be tuned by varying the relative
132: strengths of the lasers. The drain is a state-selective atomic filter,
133: such as a Stern-Gerlach device or radio-frequency or Raman outcoupler
134: \citep{Edwards}. While the goal of this paper is to explore the possibility
135: of constructing the atomic analog of spintronic devices, the two dark
136: states of the tripod atom can be considered qubit states \citep{qi2,qi0,unanyan04PRA,qi3};
137: in this context the atomic DDT represents a single-qubit phase gate
138: for a dilute atomic beam. In contrast to typical single qubit gates,
139: this setup does not involve time-dependent pulses, and the amount
140: of the qubit rotation within the gate region is independent of the
141: atom's velocity, due to the geometric nature of the process.
142: 
143: 
144: \paragraph{Tripod scheme}
145: 
146: The proposed DDT implementations exploit the tripod scheme (Fig.~\ref{fig:ddtschematic}b,c)
147: \citep{unanyan1998rca,unanyan99pra,theuer1999nlc,stirap,ruseckas-2005-95,lietfiz-2007},
148: in which a four level atom feels two counterpropagating stationary
149: laser beams and a third orthogonal beam \citep{Jacob07,prar08,juz08-neg-refl}.
150: The lasers induce transitions between the ground states $|j\rangle$
151: ($j=1,2,3$) and an excited state $|0\rangle$ with spatially dependent
152: Rabi frequencies $\Omega_{1}=|\Omega_{1}|e^{-i\kappa_{0}x}$ , $\Omega_{2}=|\Omega_{2}|e^{i\kappa_{0}x}$
153: and $\Omega_{3}=|\Omega_{3}|e^{i\kappa_{0}z}$ , $\kappa_{0}$ being
154: a wave-number.
155: 
156: The electronic Hamiltonian of a tripod atom is, in the interaction
157: representation and rotating wave approximation, $\hat{H}_{e}=-\hbar\Omega|B\rangle\langle0|+\mathrm{H.c.}$,
158: where $|B\rangle=\left(|1\rangle\Omega_{1}^{*}+|2\rangle\Omega_{2}^{*}+|3\rangle\Omega_{3}^{*}\right)/\Omega$
159: and $\Omega^{2}=|\Omega_{1}|^{2}+|\Omega_{2}|^{2}+|\Omega_{3}|^{2}$.
160: $\hat{H}_{e}$ has two degenerate dark states $|D_{j}\rangle$ containing
161: no excited state contribution: $\hat{H}_{e}|D_{j}\rangle=0$, $j=1,2.$
162: An additional pair of bright eigenstates $|\pm\rangle=\left(|B\rangle\pm|0\rangle\right)/\sqrt{2}$
163: is separated from the dark states by $\pm\hbar\Omega$. For the light
164: fields of interest, the dark states can be chosen as:
165: 
166: \begin{eqnarray}
167: |D_{1}\rangle & = & \left(\sin\varphi|1\rangle^{\prime}-\cos\varphi|2\rangle^{\prime}\right)\,,\label{eq:D1}\\
168: |D_{2}\rangle & = & \varepsilon\left(\cos\varphi|1\rangle^{\prime}+\sin\varphi|2\rangle^{\prime}\right)-\sqrt{1-\varepsilon^{2}}|3\rangle,\label{eq:D2}\end{eqnarray}
169: with $|1\rangle^{\prime}=|1\rangle e^{i\kappa_{0}(z+x)}$ and $|2\rangle^{\prime}=|2\rangle e^{i\kappa_{0}(z-x)}$,
170: where\begin{equation}
171: \varepsilon=|\Omega_{3}|/\Omega\,,\quad\varphi=\arctan(|\Omega_{1}|/|\Omega_{2}|)\label{eq:epsilon-phi}\end{equation}
172: characterize the relative intensities of the laser beams. The dark
173: states $|D_{j}\rangle\equiv|D_{j}(\mathbf{r})\rangle$ are position-dependent
174: due to the spatial variation of the Rabi frequencies $\Omega_{j}(\mathbf{r})$.
175: 
176: Let us adiabatically eliminate the bright states, so that the atom
177: evolves within the dark-state manifold.  The full atomic state vector
178: can then be expanded as $|\Psi(\mathbf{r},t)\rangle=\sum_{n=1}^{2}\chi_{n}(\mathbf{r},t)|D_{n}(\mathbf{r})\rangle$,
179: where $\chi_{n}(\mathbf{r},t)$ describes the motion of an atom in
180: the dark state $|D_{n}(\mathbf{r})\rangle$. The atomic center of
181: mass motion is thus represented by a two-component wavefunction $\chi=(\chi_{1},\chi_{2})^{T}$
182: obeying \citep{ruseckas-2005-95}\textbf{\begin{equation}
183: i\hbar\frac{\partial}{\partial t}\chi=\left[\frac{1}{2M}(-i\hbar\nabla-\mathbf{A})^{2}+U\right]\chi\,,\label{eq:general-eq}\end{equation}
184: }where $\mathbf{A}$ is the effective vector potential \citep{ruseckas-2005-95,berry84,wilczek84,mead91}
185: representing a $2\times2$ matrix whose elements are vectors, $\mathbf{A}_{n,m}=i\hbar\langle D_{n}(\mathbf{r})|\nabla D_{m}(\mathbf{r})\rangle$.
186: The particular light field configuration we have chosen yields\begin{eqnarray}
187: \mathbf{A}_{11} & = & -\hbar\kappa_{0}(\mathbf{e}_{z}-\cos(2\varphi)\mathbf{e}_{x}),\label{eq:A-11}\\
188: \mathbf{A}_{12} & = & -\hbar\varepsilon(\kappa_{0}\sin(2\varphi)\mathbf{e}_{x}+i\nabla\varphi),\label{eq:A-12}\\
189: \mathbf{A}_{22} & = & -\hbar\kappa_{0}\varepsilon^{2}(\mathbf{e}_{z}+\cos(2\varphi)\mathbf{e}_{x}),\label{eq:A-22}\end{eqnarray}
190: with $\mathbf{e}_{x}$ and $\mathbf{e}_{z}$ the unit Cartesian vectors.
191: The $2\times2$ matrix $U$ with elements $U_{nm}=(\hbar^{2}/2M)\langle D_{n}(\mathbf{r})|\nabla B(\mathbf{r})\rangle\langle B(\mathbf{r})|\nabla D_{m}(\mathbf{r})\rangle$
192: is an effective scalar potential; both $\mathbf{A}$ and $U$ arise
193: due to the spatial dependence of the atomic dark states. 
194: 
195: Suppose the incident atom has a velocity $\mathbf{v}$ much greater
196: than the recoil velocity $v_{\mathrm{rec}}=\hbar\kappa_{0}/M\approx0.5\mbox{cm/s}$
197: for $^{87}$Rb. In this limit, the laser beams do not significantly
198: change the atom's velocity, permitting a simplified semiclassical
199: approach with no reflected waves. We apply a gauge transformation
200: $\chi(\mathbf{r},t)=e^{iM\mathbf{v}\cdot\mathbf{r}/\hbar-iM\mathbf{v}^{2}t/2\hbar}\tilde{\chi}(\mathbf{r},t)$,
201: implying transition to a reference frame moving with velocity $\mathbf{v}$,
202: where the two-component envelope function $\tilde{\chi}$ varies slowly
203: with $\mathbf{r}$ over the atom's wavelength $\lambda=h/(Mv)$. Keeping
204: only terms containing $\mathbf{v}$ (or its time derivatives), we
205: arrive at the following approximate equation for $\tilde{\chi}$:
206: 
207: \begin{equation}
208: i\hbar\left(\partial/\partial t+\mathbf{v}\cdot\nabla\right)\tilde{\chi}(\mathbf{r},t)=-\mathbf{v}\cdot\mathbf{A}(\mathbf{r})\tilde{\chi}(\mathbf{r},t).\label{eq:specific-eq-envelope}\end{equation}
209: As the omitted scalar potential $U$ and the $A^{2}$ term are of
210: the order of the recoil energy $\hbar\omega_{\mathrm{rec}}=\hbar^{2}\kappa_{0}^{2}/2M\ll Mv^{2}/2$,
211: the fast moving atoms will not feel these potentials. For incident
212: velocities $v$ of the order of $v_{\mathrm{rec}}$ or smaller, the
213: atomic motion will undergo a \emph{Zitterbewegung} \citep{jayzb,patrikzb}
214: which is beyond the scope of the present study. While the atoms must
215: move much faster than the recoil velocity, they should also be slow
216: enough to avoid coupling to the bright states. We provide a quantitative
217: analysis of these limitations near the end of the Letter.
218: 
219: In both of the DDT schemes to be presented, the operator \textbf{$\mathbf{v}\cdot\mathbf{A}$}
220: commutes with itself at different times. Going to a moving frame of
221: reference $\mathbf{r}^{\prime}=\mathbf{r}-\mathbf{v}t$, we can thus
222: relate the wavefunction $\tilde{\chi}$ at time $t=t_{f}$ to the
223: wavefunction at a previous time $t=t_{i}$ through\begin{equation}
224: \tilde{\chi}(\mathbf{r}^{\prime},t_{f})=\exp(i\Theta)\tilde{\chi}(\mathbf{r}^{\prime},t_{i})\,.\label{eq:envelope-solution}\end{equation}
225: The $2\times2$ Hermitian matrix $\Theta=-\hbar^{-1}\int_{t_{i}}^{t_{f}}\mathbf{A}(\mathbf{r}^{\prime}+\mathbf{v}t)\cdot\mathbf{v}dt$
226: describes the evolution of the internal state of the atom as it traverses
227: the path from $\mathbf{r}_{i}=\mathbf{r}^{\prime}+\mathbf{v}t_{i}$
228: to $\mathbf{r}_{f}=\mathbf{r}^{\prime}+\mathbf{v}t_{f}$, \begin{equation}
229: \Theta=-\frac{1}{\hbar}\int_{\mathbf{r}_{i}}^{\mathbf{r}_{f}}\mathbf{A}(\mathbf{r})\cdot d\mathbf{r}\,.\label{eq:theta}\end{equation}
230: Our subsequent analysis of the atomic dynamics will center on Eqs.
231: (\ref{eq:envelope-solution})-(\ref{eq:theta}) and (\ref{eq:A-11})-(\ref{eq:A-22}). 
232: 
233: 
234: \paragraph{Atomic analogs of the DDT}
235: 
236: We first consider the setup depicted in Figs.~\ref{fig:ddtschematic}c
237: and \ref{fig:razmik}a. The atoms are incident along the $y$ axis,
238: along which laser beams $1$ and $2$ are relatively shifted \citep{unanyan1998rca,theuer1999nlc,unanyan99pra,lietfiz-2007},
239: so that \begin{equation}
240: A_{y}=\hbar\sigma_{y}\varepsilon(y)\partial\varphi(y)/\partial y\,.\label{eq:A-y}\end{equation}
241:  Equations (\ref{eq:A-y}) and (\ref{eq:theta}) yield \begin{equation}
242: \Theta=\alpha\sigma_{y}\,,\qquad\alpha=-\int_{y_{i}}^{y_{f}}\varepsilon(y)\frac{\partial}{\partial y}\varphi(y)dy\,,\label{eq:alpha-Razmik}\end{equation}
243: where $\alpha$ is the mixing angle, $\sigma_{y}$ (or $\sigma_{x}$)
244: being the usual Pauli matrix. By taking the initial and final times
245: sufficiently large, we have $y_{i}\rightarrow-\infty$ and $y_{f}\rightarrow+\infty$. 
246: 
247: %
248: \begin{figure}
249: \begin{centering}
250: \includegraphics[width=0.75\columnwidth]{fig2}
251: \par\end{centering}
252: 
253: \caption{Schematics of the first (a) and second (b) setups for an atomic transistor:
254: The atom, along its trajectories (shown in Figs.~\ref{fig:ddtschematic}c,d)
255: sees the above profile of laser fields. }
256: 
257: 
258: \label{fig:razmik}
259: \end{figure}
260: 
261: 
262: As Figs.~\ref{fig:ddtschematic}c and \ref{fig:razmik}a show, the
263: first laser beam dominates as the atom enters the gate region, while
264: the second dominates as it exits the region. In between, the atom
265: also feels the third beam. This configuration results in a gate-induced
266: rotation of the atom's internal state by a mixing angle \textbf{$\alpha$}.
267: Specifically, suppose the atom enters the gate region in the internal
268: state $|3\rangle=-|D_{2}(\mathbf{r}^{\prime},t_{i})\rangle$, with
269: center of mass wave-function \textbf{$\Phi(\mathbf{r}^{\prime})$}.
270: The atom then exits the gate region in the rotated state \begin{equation}
271: \tilde{\chi}(\mathbf{r}^{\prime},t_{f})=-\Phi(\mathbf{r}^{\prime})\left(\begin{array}{c}
272: \sin\alpha\\
273: \cos\alpha\end{array}\right).\label{eq:Psi-fin}\end{equation}
274: Thus, the probability for the atom to emerge in the second dark state
275: is \textbf{$\cos^{2}\alpha$}. Note that the second dark state coincides
276: with the third internal ground state upon exit:  \textbf{$|D_{2}(\mathbf{r}^{\prime},t_{f})\rangle=-|3\rangle$}.  This gate-controlled state rotation is an atomic analog of the action
277: of the DDT. Define $\eta=|\Omega_{3}|/|\Omega_{1}|$ as the relative amplitude of the third laser
278: at the central point. The specific relation between
279: $\alpha$ and $\eta$ depends on the particular choice of light field
280: configuration and is readily derived from Eqs. (\ref{eq:epsilon-phi})
281: and (\ref{eq:alpha-Razmik}).  For arbitrary light field configurations, $\alpha$ is a
282: complicated space-dependent function.  However for the particular
283: laser configuration we examine here, $\alpha$ simplifies to a function solely depending on $\eta$, and  $\eta$ controls $\alpha$.  Fig. \ref{fig:fig3}\textbf{ }shows the dependence of $\alpha$ on $\eta$ for Gaussian laser beams. As in the electronic DDT, the transmission coefficient $\cos\alpha$
284: is independent of the velocity of the incident atoms, so that the
285: transistor properties are robust to a spread in atomic velocities.
286: We estimate the regime of validity of this independence near the end
287: of the Letter.
288: \begin{figure}
289: \begin{centering}
290: \includegraphics[width=0.55\columnwidth]{fig3}
291: \par\end{centering}
292: 
293: \caption{The mixing angle $\alpha$ vs. the relative amplitude of the third
294: field $\eta$ for the first (solid line) and the second (dashed line)
295: setups. The amplitudes of the beams are Gaussian: \textbf{$|\Omega_{1}|=a\exp(-(u+\delta)^{2}/w_{1}^{2})$},
296: \textbf{$|\Omega_{2}|=a\exp(-(u-\delta)^{2}/w_{2}^{2})$}, and \textbf{$|\Omega_{3}|=a\eta\exp(-u^{2}/w_{3}^{2}-\delta^{2}/w_{1}^{2})$},
297: with $u=y$ (first setup) or $u=x$ (second setup). In the first setup,
298: $w_{1}=w_{2}=w_{3}=\delta=2\lambda$, with $\lambda=600\,\mathrm{nm}$
299: being the laser wave length. For the second setup, all the beams are
300: centered at the same point ($\delta=0$) and have the widths $w_{1}=w_{2}=10w_{3}=20\lambda$.
301: \textbf{\label{fig:fig3}}}
302: 
303: \end{figure}
304: 
305: 
306: Since $\varepsilon(y)\leq1$, the mixing angle given by Eq.~(\ref{eq:alpha-Razmik})
307: ranges from $0$ to $\pi/2$, and the sensitivity $|\Delta\alpha|/|\Delta\eta|$
308: of the DDT is on the order of unity. Small changes in the relative
309: Rabi frequency $\eta$ will thus lead to small changes in the mixing
310: angle: $|\Delta\alpha|\sim|\Delta\eta|$. We next analyze an alternative
311: setup which enables us to create a more sensitive DDT.
312: 
313: Now suppose that the first two light beams counterpropagate along
314: the $x$ axis with equal intensities (Fig.~\ref{fig:ddtschematic}d),
315: i.e., $\varphi=\pi/4$ in Eqs.~(\ref{eq:A-11})-(\ref{eq:A-22})
316: for $\mathbf{A}$. After the trivial gauge transformation $\exp[i\hbar\kappa_{0}(1+\varepsilon^{2})z\mathbf{I}]$,
317: the light-induced vector potential resembles the Rashba spin orbit
318: coupling which is the spin rotation mechanism of the electronic DDT: 
319: 
320: \begin{eqnarray}
321: A_{z} & = & -\frac{\hbar\kappa_{0}}{2}(1-\varepsilon^{2})\sigma_{z}\label{eq:A-2-setup}\\
322: A_{x} & = & -\hbar\kappa_{0}\varepsilon\sigma_{x},\qquad A_{y}=0.\end{eqnarray}
323: 
324: 
325: The atomic beam crosses the lasers at an angle in the $x-y$ plane,
326: with initial velocity components $v_{x}\neq0$ and $v_{y}$. Although
327: the atomic motion in the $y$ direction does not affect the internal
328: state rotation \textbf{$(A_{y}=0)$} , sending the beam in at an angle
329: removes the experimental difficulty of having the atoms incident from
330: the same direction as the laser beams. Along its trajectory, the atom
331: feels the laser beam profile illustrated in Fig.~\ref{fig:razmik}b.
332: The evolution matrix of Eq.~(\ref{eq:theta}) is then\begin{equation}
333: \Theta=\alpha\sigma_{x}\,,\qquad\alpha=\kappa_{0}\int_{x_{i}}^{x_{f}}\varepsilon(x)dx\,.\label{eq:theta-our}\end{equation}
334: Initial and final times are taken sufficiently large that the spatial
335: integration runs from $x_{i}=-\infty$ to $x_{f}=+\infty$.
336: 
337: As in the previous scheme, the intensity of the third laser vanishes
338: ($\varepsilon\rightarrow+0$) outside the gate region (see Fig.~\ref{fig:razmik}b).
339: Only the third laser's intensity has significant spatial dependence
340: inside the gate region, the intensities of the first two lasers being
341: nearly constant there. In both setups, the controlled state rotation
342: arises from the spatial dependence of the beams in the gate region.
343: In the first setup the variation is in the lasers' relative intensities.
344: Contrastingly, in the second setup, the intensities of the first two
345: lasers are constant in the gate region, so the controlled state rotation
346: is driven by only the relative \emph{phases} of the counterpropagating
347: laser beams. 
348: 
349: As in the previous setup, the atom enters the gate region in the internal
350: state \textbf{$|3\rangle=-|D_{2}(\mathbf{r}^{\prime},t_{i})\rangle$}
351: and with center of mass wave-function $\Phi(\mathbf{r}^{\prime})$.
352: The atom exits in the rotated state\begin{equation}
353: \tilde{\chi}(\mathbf{r}^{\prime},t_{f})=-\Phi(\mathbf{r}^{\prime})\left(\begin{array}{c}
354: i\sin\alpha\\
355: \cos\alpha\end{array}\right),\label{eq:Phi-fin-altern}\end{equation}
356: where the mixing angle $\alpha$ is controlled by the variation of
357: the relative intensity of the third laser beam.
358: 
359: To estimate the mixing angle, suppose that $\Omega_{3}$, and hence
360: $\varepsilon$, do not change significantly in the gate region. Equation
361: (\ref{eq:theta-our}) then gives $\alpha=\kappa_{0}\bar{\varepsilon}L$,
362: where $L$ (see Fig.~\ref{fig:razmik}b) is the length of the area
363: in which the third laser has the strongest intensity. Note that the
364: mixing angle is now proportional to $L$, as well as to the average
365: strength $\kappa_{0}\bar{\varepsilon}$ of the spin-orbit coupling.
366: This behavior is in direct analogy to the electronic DDT \citep{dattadas}.
367: As in Eq.~(2) of \citep{dattadas}, the output power of the atoms
368: in the internal state $|3\rangle$ is $P=\cos^{2}\alpha=\cos^{2}(\kappa_{0}\bar{\varepsilon}L)$.
369: Using this atomic setup, $\alpha=\kappa_{0}\bar{\varepsilon}L$ can
370: be much larger than $\pi/2$, provided $L\gg(\kappa_{0}\bar{\varepsilon})^{-1}$,
371: as shown in Fig.~\ref{fig:fig3}. Small changes in the relative amplitude
372: of the third laser $\eta=|\Omega_{3}|/|\Omega_{1}|$ can therefore
373: yield substantial changes in the mixing angle: $|\Delta\alpha|\sim|\Delta\eta|\kappa_{0}L$.
374: The sensitivity of such a DDT, $|\Delta\alpha|/|\Delta\eta|\sim\kappa_{0}L$,
375: can far exceed unity if $L$ is much greater than the optical wave-length
376: $\lambda=2\pi/\kappa_{0}$. 
377: 
378: Let us estimate the range of atomic beam velocities for which our
379: approximations are valid. The atom crosses the gate region in a time
380: $\tau=L/v$. Due to nonadiabatic coupling to the bright states, the
381: dark state atoms have the finite lifetime $\tau_{D}=\Omega^{2}/\gamma\Delta\omega^{2}$
382: \citep{Juz06PRA}, where $\gamma$ is the excited state decay rate
383: and $\Delta\omega=v\partial\varphi/\partial y\sim v\pi/L$ (first
384: setup), or $\Delta\omega=v\kappa_{0}$ (second setup). The frequency
385: shift $\Delta\omega$ represents the two-photon detuning due to the
386: finite time of the atom-light interaction (first setup) or the two-photon
387: Doppler shift (second setup). To avoid decay, we require the beam
388: to be in the adiabatic limit, i.e. $\tau/\tau_{D}\ll1$. Taking $\Omega=2\pi\times10^{7}\,\mathrm{Hz}$
389: \citep{Hau99}, $\gamma=10^{7}\,\mathrm{s}^{-1}$, $\kappa_{0}=2\pi/\lambda$,
390: $\lambda=600\,\mathrm{nm}$ and $L=4\lambda$, we require atomic velocities
391: $v\ll100\,\mathrm{m/s}$ for the first setup and $v\ll1\,\mathrm{m/s}$
392: for the second setup. The increased sensitivity in the second scheme
393: thus comes at the expense of increased non-adiabatic losses.
394: 
395: Ultracold atoms are highly tunable and controllable, and can thus
396: serve as quantum simulators for a variety of other systems, including
397: systems which have yet to be experimentally accessed in their original
398: manifestations. In this Letter, we have identified an atomic analog
399: of one such system, the spin field-effect transistor. Our atomic transistors,
400: like their electronic counterpart, provide controllable state manipulation
401: that is relatively insensitive to the thermal spread of beam velocities.
402: The devices we have proposed are based on the familiar tripod STIRAP
403: configuration, and appear to be feasible within current experimental
404: procedures.
405: 
406: \bibliographystyle{prsty}
407: 
408: \begin{thebibliography}{10}
409: 
410: \bibitem{dattadas}
411: S. Datta and B. Das, Appl. Phys. Lett. {\bf 56},  665  (1990).
412: 
413: \bibitem{zutic2004sfa}
414: I. Zutic, J. Fabian, and S. Das~Sarma, Rev. Mod. Phys. {\bf 76},  323  (2004).
415: 
416: \bibitem{Atom-Diode}
417: A. Ruschhaupt, J.~G. Muga, and M.~G. Raizen, J. Phys. B: At. Mol. Opt. Phys.
418:   {\bf 39},  L133  (2006).
419: 
420: \bibitem{seaman:023615}
421: B.~T. Seaman, M. Kr\"{a}mer, D.~Z. Anderson, and M.~J. Holland, Phys. Rev. A
422:   {\bf 75},  023615  (2007).
423: 
424: \bibitem{experimentaldiode}
425: J.~J. Thorn, E.~A. Schoene, T. Li, and D.~A. Steck, Phys. Rev. Lett. {\bf 100},
426:    240407  (2008).
427: 
428: \bibitem{unanyan1998rca}
429: R.~G. Unanyan, M. Fleischhauer, B.~W. Shore, and K. Bergmann, Opt. Commun. {\bf
430:   155},  144  (1998).
431: 
432: \bibitem{unanyan99pra}
433: R.~G. Unanyan, B.~W. Shore, and K. Bergmann, Phys. Rev. A {\bf 59},  2910
434:   (1999).
435: 
436: \bibitem{theuer1999nlc}
437: H. Theuer {\it et~al.}, Opt. Express {\bf 4},  77  (1999).
438: 
439: \bibitem{stirap}
440: F. Vewinger {\it et~al.}, Phys. Rev. Lett. {\bf 91},  213001  (2003).
441: 
442: \bibitem{ruseckas-2005-95}
443: J. Ruseckas, G. Juzeli{\=u}nas, P. {\"O}hberg, and M. Fleischhauer, Phys. Rev.
444:   Lett. {\bf 95},  010404  (2005).
445: 
446: \bibitem{lietfiz-2007}
447: G. Juzeli{\=u}nas, J. Ruseckas, P. {\"O}hberg, and M. Fleischhauer, Lithuanian
448:   J. Phys {\bf 47},  351  (2007).
449: 
450: \bibitem{spielman-2008}
451: Y.-J. Lin {\it et~al.}, arXiv:0809.2976  (2008).
452: 
453: \bibitem{Rashba60}
454: E.~I. Rashba, Sov. Phys. Sol. St. {\bf 2},  1224  (1960).
455: 
456: \bibitem{galitski}
457: T.~D. Stanescu, C. Zhang, and V. Galitski, Phys. Rev. Lett. {\bf 99},  110403
458:   (2007).
459: 
460: \bibitem{Jacob07}
461: A. Jacob, P. {\"O}hberg, G. Juzeli{\=u}nas, and L. Santos, Appl. Phys. B {\bf
462:   89},  439  (2007).
463: 
464: \bibitem{jayzb}
465: J.~Y. Vaishnav and C.~W. Clark, Phys. Rev. Lett. {\bf 100},  153002  (2008).
466: 
467: \bibitem{prar08}
468: G. Juzeli{\=u}nas {\it et~al.}, Phys. Rev. A {\bf 77},  011802(R)  (2008).
469: 
470: \bibitem{juz08-neg-refl}
471: G. Juzeli{\=u}nas {\it et~al.}, Phys. Rev. Lett. {\bf 100},  200405  (2008).
472: 
473: \bibitem{Edwards}
474: M. Edwards {\it et~al.}, J. Phys. B: At. Mol. Opt. Phys. {\bf 32},  2935
475:   (1999).
476: 
477: \bibitem{qi2}
478: L.~M. Duan, J.~I. Cirac, and P. Zoller, Science {\bf 292},  1695  (2001).
479: 
480: \bibitem{qi0}
481: Z. Kis and F. Renzoni, Phys. Rev. A {\bf 65},  032318  (2002).
482: 
483: \bibitem{unanyan04PRA}
484: R.~G. Unanyan and M. Fleischhauer, Phys. Rev. A {\bf 69},  050302(R)  (2004).
485: 
486: \bibitem{qi3}
487: S. Rebi{\'c} {\it et~al.}, Phys. Rev. A {\bf 70},  032317  (2004).
488: 
489: \bibitem{berry84}
490: M.~V. Berry, Proc. R. Soc. A {\bf 392},  45  (1984).
491: 
492: \bibitem{wilczek84}
493: F. Wilczek and A. Zee, Phys. Rev. Lett. {\bf 52},  2111  (1984).
494: 
495: \bibitem{mead91}
496: C.~A. Mead, Rev. Mod. Phys. {\bf 64},  51  (1992).
497: 
498: \bibitem{patrikzb}
499: M. Merkl, F.~E. Zimmer, G. Juzeli{\=u}nas, and P. {\"O}hberg, Europhys. Lett.
500:   {\bf 83},  54002  (2008).
501: 
502: \bibitem{Juz06PRA}
503: G. Juzeli{\=u}nas, J. Ruseckas, P. \"Ohberg, and M. Fleischhauer, Phys. Rev. A {\bf 73},
504:   025602  (2006).
505: 
506: \bibitem{Hau99}
507: L. Hau, S.~E. Harris, Z. Dutton, and C. Behrooz, Nature {\bf 397},  594
508:   (1999).
509: 
510: \end{thebibliography}
511: 
512: 
513: \end{document}
514: