cond-mat0504583/eod.tex
1: \documentclass[aps,prb,twocolumn,groupedaddress,showpacs]{revtex4}
2: 
3: \usepackage{graphicx}
4: 
5: \bibliographystyle{apsrev}
6: 
7: \begin{document}
8: 
9: \title{Generating spin-entangled electron pairs in 
10: normal conductors using voltage pulses}
11: 
12: \author{A.V.\ Lebedev$^{\, a}$, G.B.\ Lesovik$^{\, a}$,
13: and G.\ Blatter$^{\, b}$}
14: 
15: \affiliation{$^{a}$L.D.\ Landau Institute for Theoretical Physics,
16: RAS, 119334 Moscow, Russia}
17: 
18: \affiliation{$^{b}$Theoretische Physik, ETH-H\"onggerberg, CH-8093
19: Z\"urich, Switzerland}
20: 
21: \date{\today}
22: 
23: \begin{abstract}
24:    We suggest an operating scheme for the deliberate generation
25:    of spin-entangled electron pairs in a normal-metal mesoscopic
26:    structure with fork geometry. Voltage pulses with associated
27:    Faraday flux equal to one flux unit $\Phi_0=hc/e$ drive
28:    individual singlet-pairs of electrons towards the beam splitter.
29:    The spin-entangled pair is created through a post-selection
30:    in the two branches of the fork. We analyze the appearance of
31:    entanglement in a Bell inequality test formulated in terms of
32:    the number of transmitted electrons with a given spin polarization.
33: \end{abstract}
34: 
35: \pacs{03.67.Mn, 05.30.Fk, 05.60.Gg, 73.23.-b}
36: 
37: \maketitle
38: 
39: \section{Introduction}\label{sec:intro}
40: 
41: Quantum entanglement of electronic degrees of freedom 
42: in mesoscopic devices has attracted a lot of interest 
43: recently. Early proposals for structures generating 
44: streams of entangled particles exploit the interaction 
45: between electrons as a resource for producing entanglement, 
46: the pairing interaction in superconductors~\cite{lesovik_01} 
47: or the repulsive Coulomb interaction in confined 
48: geometries~\cite{ionicioiu_01,oliver_02}. Recently, 
49: another class of devices has been suggested which 
50: avoids direct interparticle interaction; instead, the 
51: entanglement originates from a proper post-selection of
52: orbital~\cite{beenakker_03,samuelsson_03,samuelsson_04} or
53: spin~\cite{fazio_04,lebedev_05,lebedev_04} degrees of 
54: freedom. The majority of these proposals deals with 
55: the situation where the entangled particles are emitted 
56: in a random and uncontrolled fashion, while entanglement
57: on demand is implicit in the scheme proposed in Ref.\
58: \onlinecite{ionicioiu_01}.
59: 
60: Current interest concentrates on setups which are 
61: capable to produce pairs of entangled electrons `on demand'.
62: Such controlled entanglement is an essential step
63: towards the realization of quantum computing devices
64: for which electronic orbital- or spin degrees of freedom
65: may serve as qubits \cite{loss_98}. In addition,
66: prospects to convert electronic entanglement into a
67: photonic one with high efficiency look promising
68: \cite{loss_04}; this may open new opportunities for the
69: manipulation of entangled photons with an enhanced
70: efficiency.
71: 
72: While entanglement on demand is implicit in the work of 
73: Ionicioiu {\it et al.}, a detailed discussion of the 
74: controlled production of entanglement in a mesoscopic
75: device has only been given recently by 
76: Samuelsson and B\"uttiker \cite{samuelsson_04b}; 
77: they proposed a scheme for the dynamical generation of 
78: orbitally entangled electron-hole pairs where a 
79: time-dependent harmonic electric potential is applied 
80: between two spatially separated regions of a Mach-Zehnder 
81: interferometer operating in the Quantum Hall regime.  
82: This (perturbative) analysis concentrated on the limit 
83: of a weak pumping potential generating only a small 
84: fraction of entangled electron-hole pairs per cycle.
85: Later on, several schemes have been suggested producing
86: entangled pairs on demand with a high efficiency: in
87: their setup, Beenakker {\it et al.} \cite{beenakker_05}
88: make use of a ballistic two-channel conductor driven with
89: a strong oscillating potential. In their non-perturbative
90: analysis they demonstrate that this device can pump up
91: to one (spin- or orbital) entangled Bell-pair per two
92: cycles. A different proposal based on spin resonance
93: techniques acting on electrons trapped in a double quantum
94: dot structure and subsequently released into two quantum
95: channels has been suggested by  Blaauboer and DiVincenzo
96: \cite{divincenzo_05}; their detailed analysis of the
97: manipulation and measurement schemes demonstrates
98: that the production and measurement of entangled pairs
99: via an optimal entanglement witness can be performed with
100: present days experimental technology.
101: %
102: \begin{figure} [h]
103:    \includegraphics[scale=0.35]{fandf.eps}
104:    \caption[]{Mesoscopic normal-metal structures in a fork geometry
105:    generating spin-correlated electrons in the two arms of the
106:    fork. The Bell type setups detect the number of transmitted
107:    particles $N_i$, $i=1,3$, with spin projected onto the directions
108:    $\pm{\bf a}$ in the upper arm and correlates them with the number
109:    $N_j$, $j=2,4$, of particles with a spin projected onto the
110:    directions $\pm{\bf b}$ in the lower arm. (a) Fork with a simple
111:    splitter with particles injected along the single source lead.
112:    (b) Fork in the geometry of a four-terminal beam splitter
113:    with particles injected along one of the incoming channels
114:    only. Quenching the transmission $T_\mathrm{ud}$ between
115:    the upper and lower leads allows to eliminate equilibrium
116:    fluctuations spoiling the entanglement.}
117:    \label{fig:fork}
118: \end{figure}
119: %
120: 
121: In the present paper, we discuss an alternative scheme 
122: generating pulsed spin-entangled electron pairs in a 
123: normal-metal mesoscopic structure arranged in a fork geometry, 
124: see Fig.\ \ref{fig:fork}. In this device, spin-entangled 
125: electron pairs are generated via the injection of spin-singlet
126: pairs into the source lead from the reservoir \cite{lebedev_05}. 
127: This entanglement is made accessible by splitting the pair 
128: into the two leads `u' and `d' and subsequent projection 
129: (through the Bell measurement) to that part of the wave 
130: function describing separated electrons travelling in 
131: different leads \cite{lebedev_05,lebedev_04}.
132: Rather then quantum pumping with a cyclic potential as
133: in Refs.\ \onlinecite{samuelsson_04b,beenakker_05}, 
134: our proposal makes use of definite voltage 
135: pulses generating spin-entangled electron pairs. A pulsed 
136: sequence of ballistic electrons is implicitly assumed in the
137: generation of orbital entanglement by Ionicioiu {\it et al.} 
138: \cite{ionicioiu_01}, however, no description has been given 
139: how such (single-electron) pulses are generated in practice. 
140: Below we discuss a scheme where voltage pulses of specific form 
141: accumulating one unit of flux $\Phi_0 = -c \int dt \, V(t)$ and 
142: applied to the source lead `s' generate pairs of spin-entangled
143: electrons which then are distributed between the two outgoing
144: leads of the fork, the upper and lower arms denoted as `u' and
145: `d'. These spin-entangled electron states are subsequently analyzed
146: in a Bell experiment \cite{bell} involving the measurement of
147: cross-correlations \cite{nike_02} between the number of electrons
148: transmitted through the corresponding spin filters in the two arms
149: of the fork, see Fig.\ \ref{fig:fork}. Using time resolved
150: correlators, we are in a position to analyze arbitrary forms of
151: voltage pulses and determine the resulting degree of violation in
152: the Bell setup. We find that Lorentzian shaped pulses generate
153: spin-entangled pairs with 50 \% probability, corresponding in
154: efficiency to the optimal performance of one entangled pair per
155: two cycles as found by Beenakker {\it et al.} \cite{beenakker_05}.
156: The reduction in efficiency to 50 \% is due to the competing
157: processes where the spin-entangled pair generated by the
158: voltage pulse propagates into only one of the two arms.
159: In order to make use of this structure as a deterministic
160: entangler, the Bell measurement setup has to be replaced through a
161: corresponding projection device (post-) selecting that part of the
162: wave function with the two electrons distributed between the two
163: arms; alternatively, this post-selection may be part of the
164: application device itself, as is the case in the Bell inequality
165: measurement.
166: 
167: In the following, we first derive (Sec.\ \ref{sec:BI}) an
168: expression for the Bell inequality involving the particle-number
169: cross-correlators appropriate for a pulse driven experiment.
170: We proceed with the calculation of the particle-number
171: correlators for a single voltage pulse associated with
172: an arbitrary Faraday flux (Sec.\ \ref{sec:onepulse}). The
173: results are presented in section \ref{sec:results}: we find
174: the Bell inequalities violated for single pulses carrying
175: one Faraday flux, corresponding to one pair of electrons with
176: opposite spin. Although the Bell inequality appears to be
177: violated for weak pulses (producing less than one pair) too,
178: we argue that this violation is unphysical and that its
179: appearance is due to a misconception in the original derivation
180: of the Bell inequality arising in the weak pumping limit. We also
181: generalize the discussion to the situation with more complex
182: drives (multi-pulse case and alternating pulse sequences)
183: and demonstrate that our Bell inequalities again are violated
184: only for single-pair pulses flowing in either direction
185: through the device. Our analysis of an alternating signal
186: produces an apparent violation of the Bell inequality,
187: which, however, again appears to be an artefact resulting
188: from an improper derivation of the Bell inequality for
189: the alternating signal. In both cases of failure, weak pulses
190: and alternating pulse sequences, we encounter backflow
191: phenomena which spoil the proper derivation of the Bell
192: inequality for our setup.
193: 
194: \section{Bell Inequality with Number Correlators}\label{sec:BI}
195: 
196: The Bell inequality we are going to use here has been introduced
197: by Clauser and Horne~\cite{clauser}; it is based on the Lemma
198: saying that, given a set of real numbers $x$, $\bar x$, $y$, $\bar
199: y$, $X$, $Y$ with $|x/X|$, $|\bar x/X|$, $|y/Y|$, and $|\bar y/Y|$
200: restricted to the interval $[0,1]$, the inequality $|xy - x\bar y
201: +\bar x y + \bar x \bar y|\leq 2|XY|$ holds true. We define the
202: operator of electric charge $\hat N_i(t_\mathrm{ac})$ transmitted
203: through the $i$-th spin detector during the time interval
204: $[0,t_\mathrm{ac}]$, where $t_\mathrm{ac}>0$ is the accumulation
205: time. The charge operator $\hat N_i(t_\mathrm{ac})$ can be
206: expressed via the electric current $\hat I_i(t)$ flowing through
207: the $i$-th detector, $\hat N_i(t_\mathrm{ac})= \int_0^{t_\mathrm{ac}}
208: dt^\prime\, \hat I_i(t^\prime)$. In the Bell test experiment, see
209: Fig.~\ref{fig:fork}, one measures the number of transmitted
210: electrons with a given spin polarization, $N_i$, $i=1,\dots,4$,
211: and defines the quantities $x=N_1-N_3$, $y=N_2-N_4$, $X=N_1+N_3$,
212: and $Y=N_2+N_4$ for fixed orientations ${\bf a}$ and ${\bf b}$ of
213: the polarizers (and similar for $\bar x$ and $\bar y$ for the
214: orientations $\bar {\bf a}$ and $\bar{\bf b}$), see
215: Ref.~\onlinecite{nike_02}. Our Bell setup measures the correlations
216: %
217: \begin{eqnarray}
218:       {\cal K}_{ij}({\bf a},{\bf b}) &=&
219:       \langle \hat N_i(t_\mathrm{ac}) \hat N_j(t_\mathrm{ac}) \rangle
220:       \nonumber\\
221:       &=& \int\limits_0^{t_\mathrm{ac}} dt_1 dt_2 \,
222:       \langle \hat I_i(t_1) \hat I_j(t_2) \rangle
223:       \label{ncor}
224: \end{eqnarray}
225: %
226: between the number of transmitted electrons $N_{i}$, $i=1,3$, in
227: the lead `u' with spin polarization along $\pm{\bf a}$ and
228: their partners $N_j$, $j=2,4$, in lead `d' with spin
229: polarization along $\pm{\bf b}$. Using the above definitions
230: for $x$, $y$, $X$, and $Y$, we obtain the normalized
231: particle-number difference correlator,
232: %
233: \begin{eqnarray}
234:       E({\bf a},{\bf b}) &=&
235:       \frac{\langle [\hat N_1 -\hat N_3][\hat N_2 -\hat N_4]\rangle}
236:       {\langle [\hat N_1 +\hat N_3][\hat N_2 +\hat N_4]\rangle}
237:       \nonumber\\
238:       &=& \frac{{\cal K}_{12}-{\cal K}_{14}-{\cal K}_{32}+{\cal K}_{34}}
239:       {{\cal K}_{12}+{\cal K}_{14}+{\cal K}_{32}+{\cal K}_{34}},
240: \end{eqnarray}
241: %
242: and evaluating the correlators for the four different combinations
243: of directions ${\bf a},~\bar{\bf a}$ and ${\bf b},~\bar{\bf b}$,
244: we arrive at the Bell inequality
245: %
246: \begin{equation}
247:       E_{\scriptscriptstyle\rm BI} =
248:       | E({\bf a},{\bf b}) - E({\bf a},\bar {\bf b})
249:       + E(\bar {\bf a},{\bf b}) + E(\bar {\bf a},\bar {\bf b}) |
250:       \leq 2.
251:       \label{BI1}
252: \end{equation}
253: %
254: 
255: We proceed further by separating the current
256: correlators in Eq.~(\ref{ncor}) into irreducible parts
257: $C_{ij}({\bf a},{\bf b}; t_1,t_2)=\langle \delta \hat I_i(t_1)
258: \delta \hat I_j(t_2) \rangle$ with $\delta\hat I_i(t)= \hat
259: I_i(t)- \langle \hat I_i(t)\rangle$ and products of average
260: currents and rewrite $E({\bf a},{\bf b})$ in the form
261: %
262: \begin{equation}
263:       E({\bf a},{\bf b}) =
264:       \frac{K_{12} - K_{14} - K_{32} + K_{34} + \Lambda_-}
265:       {K_{12} + K_{14} + K_{32} + K_{34}+\Lambda_+},
266: \end{equation}
267: %
268: where we have defined $\Lambda_\pm = [\langle \hat{N_1} \rangle
269: \pm \langle \hat{N_3} \rangle][\langle \hat{N_2}\rangle \pm
270: \langle \hat{N_4} \rangle]$ with the irreducible particle
271: number correlator
272: %
273: \begin{eqnarray}
274:    \label{K}
275:    K_{ij}(t_\mathrm{ac})
276:    &=& \langle \delta\hat N_i(t_\mathrm{ac})
277:    \delta\hat N_j(t_\mathrm{ac})\rangle \\
278:    &=& \int_0^{t_\mathrm{ac}} dt_1 dt_2 \,
279:    C_{ij}({\bf a},{\bf b};t_1,t_2).
280:    \nonumber
281: \end{eqnarray}
282: %
283: The average currents are related via $\langle \hat
284: I_1(t) \rangle=\langle\hat I_3(t)\rangle=\langle\hat I_\mathrm{u}
285: (t)\rangle/2$ and $\langle \hat I_2(t)\rangle =\langle \hat
286: I_4(t)\rangle =$ $\langle \hat I_\mathrm{d}(t)\rangle/2$ and thus
287: $\Lambda_-=0$, $\Lambda_+=\langle \hat N_\mathrm{u}\rangle \langle
288: \hat N_\mathrm{d}\rangle$. The irreducible current-current
289: correlator factorizes into a product of spin and orbital parts,
290: $C_{ij}({\bf a},{\bf b};t_1,t_2) =|\langle {\bf a}_i|{\bf b}_j
291: \rangle|^2 C_\mathrm{ud}(t_1,t_2)$ with ${\bf a}_{1,3}= \pm{\bf
292: a}$ and ${\bf b}_{2,4}=\pm{\bf b}$. The spin projections involve
293: the angle $\theta_{{\bf a}{\bf b}}$ between the directions $\bf a$
294: and $\bf b$ of the polarizers, $\langle \pm {\bf a}|\pm {\bf
295: b}\rangle = \cos^2 (\theta_{{\bf a}{\bf b}}/2)$ and $\langle \pm
296: {\bf a}|\mp {\bf b}\rangle = \sin^2 (\theta_{{\bf a}{\bf b}}/2)$,
297: and the Bell inequality assumes the form
298: %
299: \begin{equation}
300:       \left|\frac{K_\mathrm{ud}
301:       [\cos\theta_{{\bf a}{\bf b}}-\cos\theta_{{\bf a}\bar{\bf b}}
302:       +\cos\theta_{\bar{\bf a}{\bf b}}+\cos\theta_{\bar{\bf a}\bar{\bf b}}]}
303:       {2K_\mathrm{ud}+\langle \hat N_\mathrm{u}\rangle \langle
304:       \hat N_\mathrm{d}\rangle}
305:       \right|\leq1,
306:       \label{BI2}
307: \end{equation}
308: %
309: where $K_\mathrm{ud}(t_\mathrm{ac})= \int_0^{t_\mathrm{ac}} dt_1
310: dt_2\, C_\mathrm{ud}(t_1,t_2)$ is the (irreducible) number
311: cross-correlator between the upper and lower leads of the fork.
312: The maximal violation of the Bell inequality is attained for the
313: standard orientations of the detector polarizations $\theta_{{\bf
314: a}{\bf b}}=\theta_{\bar{\bf a}{\bf b}} =\theta_{\bar{\bf
315: a}\bar{\bf b}}=\pi/4$, $\theta_{{\bf a}\bar{\bf b}} =3\pi/4$; the
316: Bell inequality (\ref{BI2}) then reduces to
317: %
318: \begin{equation}
319:       E_{\scriptscriptstyle\rm BI} = \left|
320:       \frac{2K_\mathrm{ud}}{2K_\mathrm{ud}+\langle \hat N_\mathrm{u}
321:       \rangle \langle \hat N_\mathrm{d}\rangle}\right| \leq
322:       \frac1{\sqrt2}.
323:       \label{BI3}
324: \end{equation}
325: %
326: 
327: \section{Number Correlators for a Single Pulse}\label{sec:onepulse}
328: 
329: The orbital part $C_\mathrm{ud}(t_1,t_2)$ of the current
330: cross-corre\-lator between the upper and lower leads can be
331: calculated within the standard scattering theory of noise
332: \cite{lesovik_89,buttiker_90,lesovik_99,blanter}. We assume
333: that the time dependent voltage drop $V(t)$ at the splitter can
334: be treated adiabatically (i.e., the voltage changes slowly
335: during the electron scattering time). The electrons incident
336: from the source lead `s' and scattered to the `up' or `down'
337: lead then acquire an additional time dependent phase $\phi(t)
338: =\int_{-\infty}^t dt^\prime\, eV(t^\prime)/\hbar$. The
339: scattering states (for one spin component) describing the
340: electrons in the upper and lower leads, $\hat \Psi_\mathrm{u}
341: (x,t)$ and $\hat \Psi_\mathrm{d}(x,t)$, take the form
342: %
343: \begin{eqnarray}
344:       &&\hat \Psi_\mathrm{u} =
345:       \int \frac{d\epsilon}{\sqrt{h v_\epsilon}}\Bigl[
346:       \bigr( t_\mathrm{su} e^{i\phi(t-x/v_\epsilon)}
347:       \hat c_\epsilon + r_\mathrm{u} \hat a_\epsilon +
348:       t_\mathrm{du}\hat b_\epsilon \bigr) e^{ikx}
349:       \nonumber\\
350:       && \qquad\qquad
351:       +\,\hat a_\epsilon e^{-ikx} \Bigr] e^{-i\epsilon t/\hbar},
352:       \label{scu}
353:       \\
354:       &&\hat \Psi_\mathrm{d} =
355:       \int \frac{d\epsilon}{\sqrt{h v_\epsilon}}\Bigl[
356:       \bigr( t_\mathrm{sd} e^{i\phi(t-x/v_\epsilon)}
357:       \hat c_\epsilon + r_\mathrm{d} \hat b_\epsilon +
358:       t_\mathrm{ud}\hat a_\epsilon \bigr) e^{ikx}
359:       \nonumber\\
360:       && \qquad\qquad
361:       +\,\hat b_\epsilon e^{-ikx} \Bigr] e^{-i\epsilon t/\hbar},
362:       \label{scd}
363: \end{eqnarray}
364: %
365: where $v_\epsilon = \sqrt{2m\epsilon}$; $\hat a_\epsilon$, $\hat
366: b_\epsilon$, and $\hat c_\epsilon$ denote the annihilation
367: operators for spinless electrons at energy $\epsilon$ in leads
368: `u', `d', and `s'; the scattering amplitudes $t_\mathrm{su}$
369: ($t_\mathrm{du}$) and $t_\mathrm{sd}$ ($t_\mathrm{ud}$) describe
370: particle transmission from the source (down) lead into the upper
371: lead and from the source (up) lead into the lower (`d') lead;
372: $r_\mathrm{u}$, $r_\mathrm{d}$ denote the reflection amplitudes
373: into the leads `u' and `d'. Such adiabatically deformed scattering
374: states (\ref{scu}) and (\ref{scd}) have first been used
375: in the calculation of the spectral noise power in an $ac$-driven
376: system \cite{ll_94}; the validity of this approach has been
377: confirmed in several experiments \cite{ksp_00}.
378: 
379: We substitute these expressions into the current operator
380: $\hat I_\mathrm{u(d)}(x,t)$ and drop all terms small in
381: the parameter $|\epsilon-\epsilon^\prime|/
382: \epsilon_{\scriptscriptstyle\rm F}$ (we assume a linear
383: dispersion). The irreducible current cross-correlator
384: $C_\mathrm{ud}(t_1,t_2) = \langle \delta \hat I_\mathrm{u}
385: (x,t_1) \delta \hat I_\mathrm{d} (y,t_2) \rangle$ measured
386: at the positions $x$ and $y$ in the leads `u' and `d' can
387: be splitted into two terms, one due to equilibrium
388: fluctuations, $C_\mathrm{ud}^\mathrm{eq} (t_1-t_2)= \int
389: (d\omega/2\pi)\,S^\mathrm{eq}(\omega)e^{i\omega(t_1-t_2)}$ with
390: %
391: \begin{equation}
392:       S^\mathrm{eq}(\omega) = -\frac{2e^2}{h}\, T_\mathrm{ud}
393:       \cos(\omega\tau^+)\,\frac{\hbar\omega}{1-e^{\hbar\omega/\theta}},
394: \end{equation}
395: %
396: and a second term describing the excess correlations at finite
397: voltage,
398: %
399: \begin{equation}
400:       C_\mathrm{ud}^\mathrm{ex}(t_1,t_2) = -\frac{4e^2}{h^2}
401:       T_\mathrm{u} T_\mathrm{d}
402:       \sin^2 \frac{\phi(\xi_1)\!-\!\phi(\xi_2)}2\,
403:       \alpha(\tau\!-\!\tau^-,\theta),
404:       \label{cor_eq_ex}
405: \end{equation}
406: %
407: with $\alpha(\tau,\theta)= \pi^2\theta^2 / \sinh^2 [ \pi\theta\tau
408: / \hbar]$ ($\theta$ is the temperature of electronic reservoirs),
409: $\tau=t_1-t_2$, $\tau^\pm =(x\pm y)/v_{\scriptscriptstyle\rm F}$,
410: $\xi_1 = t_1-x/v_{\scriptscriptstyle\rm F}$, and
411: $\xi_2=t_2-y/v_{\scriptscriptstyle\rm F}$. The coefficients
412: $T_\mathrm{u} = |t_\mathrm{su}|^2$, $T_\mathrm{d} =
413: |t_\mathrm{sd}|^2$, and $T_\mathrm{ud} = |t_\mathrm{ud}|^2 =
414: |t_\mathrm{du}|^2$ denote the transmission probabilities from the
415: source to the `up', `down' leads, and from the `down' to the `up'
416: lead.
417: 
418: The equilibrium part of the current cross-correlator
419: $C_\mathrm{ud}^\mathrm{eq}(t_1-t_2)$ describes the correlations
420: of the electrons in the Fermi sea propagating ballistically
421: from lead `u' to lead `d' (or vice versa) with the retardation
422: $\tau^+=(x_1+x_2)/v_{\scriptscriptstyle\rm F}$. The corresponding
423: equilibrium part of the particle-number cross-correlator,
424: $K_\mathrm{ud}^\mathrm{eq} = \int_0^{t_\mathrm{ac}} dt_1 dt_2
425: \, C_\mathrm{ud}^\mathrm{eq}(t_1-t_2)$ then takes the form
426: %
427: \begin{equation}
428:       K_\mathrm{ud}^\mathrm{eq} \approx \frac{e^2}{\pi^2}
429:       T_\mathrm{ud} \ln \frac{t_\mathrm{ac}}{\tau},\quad
430:       \tau = \mbox{max}\{\hbar/\epsilon_{\scriptscriptstyle\rm
431:       F}, \tau^+\}, \label{ncceq}
432: \end{equation}
433: %
434: where we have assumed the zero temperature limit and an accumulation
435: time $t_\mathrm{ac}\gg\tau$. The logarithmic divergence in
436: $t_\mathrm{ac}$ reduces the violation of the Bell inequality
437: Eq.~(\ref{BI3}) at large accumulation times and one has
438: to suppress the equilibrium correlations between the upper
439: and the lower leads in the setup. This can be achieved via
440: a reduction in the transmission probability $T_\mathrm{ud}$,
441: however, in the fork geometry of Fig.~\ref{fig:fork}(a) the
442: probability $T_\mathrm{ud}$ cannot be made to vanish.
443: Alternatively, one may chose a setup with a reflectionless
444: four-terminal beam splitter as sketched in Fig.~\ref{fig:fork}(b)
445: with no exchange amplitude between the upper and lower outgoing
446: leads; using such a fork geometry, the equilibrium fluctuations
447: $K_\mathrm{ud}^\mathrm{eq}$ can be made to vanish 
448: \cite{note}.
449: 
450: Next, we concentrate on the excess part $K_\mathrm{ud}^\mathrm{ex}$
451: of the particle-number cross-correlator $\langle \hat N_\mathrm{u}
452: (t_\mathrm{ac}) \hat N_\mathrm{d}(t_\mathrm{ac}) \rangle $.
453: Note that the excess fluctuations are the same for both setups
454: Fig.~\ref{fig:fork}(a) and (b) and we can carry out all the
455: calculations for the fork geometry. We consider a sharp voltage
456: pulse applied at time $t_0$, $0<t_0<t_\mathrm{ac}$, with short
457: duration $\delta t$. The total accumulated phase $\phi(t)$ then
458: exhibits a step-like time dependence with the step height
459: $\Delta\phi= \phi(t_0 +\delta t/2)-\phi(t_0-\delta t/2)=
460: -2\pi \Phi/\Phi_0$, where we have introduced the Faraday
461: flux $\Phi=-c\int V(t) dt$ and $\Phi_0 = hc/e$ is the flux
462: quantum. The excess part of the particle-number cross-correlator
463: $K_\mathrm{ud}$ then takes the form (we consider again the zero
464: temperature limit)
465: %
466: \begin{equation}
467:       K_\mathrm{ud}^\mathrm{ex} = -
468:       \frac{e^2}{\pi^2}T_\mathrm{u} T_\mathrm{d}\!\!
469:       \int\limits_0^{t_\mathrm{ac}} \!\! dt_1 dt_2\,
470:       \frac{\sin^2[(\phi(t_1)-\phi(t_2))/2]}{(t_1-t_2)^2}.
471:       \label{ncord}
472: \end{equation}
473: %
474: For a sharp pulse with $\delta t \ll t_0, t_\mathrm{ac}$ we can
475: identify two distinct contributions arising from the integration
476: domains $|t_1-t_2|\ll \delta t$ and $|t_1-t_2|\gg \delta t$,
477: cf.\ Refs.\ \onlinecite{ll_93} and \onlinecite{lll_96};
478: we denote them with $K^<$ and $K^>$. Introducing the average
479: and relative time coordinates $t=(t_1+t_2)/2$ and $\tau=t_1-t_2$
480: and expanding the phase difference $\phi(t_1)-\phi(t_2)=
481: \phi(t+\tau/2)-\phi(t-\tau/2) \approx \dot \phi(t)\tau$, the
482: first contribution $K^<$ reads
483: %
484: \begin{eqnarray}
485:       K^< &=& -\frac{e^2}{\pi^2} T_\mathrm{u} T_\mathrm{d}
486:       \int\limits_0^{t_\mathrm{ac}} dt
487:       \int d\tau \, \frac{\sin^2[\dot \phi(t)\tau/2]}{\tau^2}
488:       \nonumber\\
489:       &=& -\frac{e^2}{2\pi} T_\mathrm{u} T_\mathrm{d}
490:       \int\limits_0^{t_\mathrm{ac}} dt\, |\dot \phi(t)|.
491: \end{eqnarray}
492: %
493: Assuming that the phase $\phi(t)$ is a monotonic function of $t$
494: (guaranteeing a unique sign for $\dot \phi(t)$) the last equation
495: can be rewritten in terms of the Faraday flux $\Phi$,
496: %
497: \begin{equation}
498:       K^< = -e^2
499:       T_\mathrm{u} T_\mathrm{d}\,\frac{|\Phi|}{\Phi_0};
500: \end{equation}
501: %
502: this contribution to the particle-number cross-correlator
503: $K_\mathrm{ud}^\mathrm{ex}$ describes the correlations
504: arising from the $n=|\Phi|/\Phi_0$ additional particles
505: pushed through the fork by the voltage pulse $V(t)$,
506: see Eq.\ (\ref{N_Phi}) below.
507: 
508: The second contribution $K^>$ to $K_\mathrm{ud}^\mathrm{ex}$
509: originates from the time domains $0<t_{1(2)}<t_0-\delta t/2$
510: and $t_0+\delta t/2<t_{2(1)}<t_\mathrm{ac}$, where $|\phi(t_1)
511: -\phi(t_2)|=2\pi \Phi/\Phi_0$, hence
512: %
513: \begin{equation}
514:       K^> \approx
515:       -\frac{2 e^2}{\pi^2} T_\mathrm{u} T_\mathrm{d} \sin^2
516:       \frac{\pi\Phi}{\Phi_0}\ln \frac{t_\mathrm{m}}{\delta t};
517:       \label{response}
518: \end{equation}
519: %
520: here, we have kept the most divergent term in the measurement
521: time $t_\mathrm{m}=t_\mathrm{ac}-t_0$, the time during which
522: the pulse manifests itself in the detector. The above
523: expression describes the response of the electron
524: gas to the sudden perturbation $V(t)$; the logarithmic
525: divergence in the measurement time $t_\mathrm{m}$ can be
526: interpreted \cite{ll_93} along the lines of the
527: orthogonality catastrophe \cite{anderson}, with the
528: isolated perturbation in space, the impurity, replaced
529: by the sudden perturbation in time. The periodicity of
530: the response in the Faraday flux $\Phi$ is due to the
531: discrete nature of electron transport as expressed
532: through the binomial character of the distribution
533: function of transmitted particles \cite{ll_93,lll_96}.
534: Remarkably, the above logarithmically divergent
535: contribution to $K_\mathrm{ud}^\mathrm{ex}$ vanishes
536: for voltage pulses carrying an integer number of electrons
537: $n=|\Phi|/\Phi_0$, see (\ref{N_Phi}) below. This follows
538: quite naturally from the invariance of the scattering
539: amplitudes $t_\mathrm{su}$ and $t_\mathrm{sd}$ in
540: Eqs.~(\ref{scu}) and (\ref{scd}) under the (adiabatic)
541: voltage pulses carrying integer flux $\pm n \Phi$,
542: $t_\mathrm{sx} \rightarrow t_\mathrm{sx} e^{\pm 2\pi n}$
543: with  $\mathrm{x}=\mathrm{u,d}$; transmitting an integer
544: number of particles at Faraday fluxes $\Phi = n \Phi_0$
545: avoids the system shakeup and the associated logarithmic
546: divergence.
547: 
548: We proceed with the determination of the average number of
549: transmitted (spinless) particles $\langle \hat N_\mathrm{u(d)}
550: (t_\mathrm{ac}) \rangle = \int_0^{t_\mathrm{ac}} dt\,
551: \langle \hat I_\mathrm{u(d)}(x,t) \rangle$. Within the
552: scattering matrix approach the average currents in the
553: upper and lower leads are given by the expression
554: %
555: \begin{eqnarray}
556:       \langle \hat I_\mathrm{u(d)}(x,t) \rangle &=& \frac{e}{h}
557:       T_\mathrm{u(d)}\, eV(t-x/v_{\scriptscriptstyle\rm F})
558:       \nonumber\\
559:       &=&\frac{e}{2\pi} T_\mathrm{u(d)}
560:       \dot \phi(t-x/v_{\scriptscriptstyle\rm F}).
561:       \label{Iud}
562: \end{eqnarray}
563: %
564: The time integration provides the average number of
565: transmitted particles
566: %
567: \begin{equation}
568:       \langle \hat N_\mathrm{u(d)}(t_\mathrm{ac})\rangle
569:       = e T_\mathrm{u(d)}\,\frac{\Phi}{\Phi_0}.
570:       \label{N_Phi}
571: \end{equation}
572: %
573: With $T_\mathrm{u}+T_\mathrm{d}=1$, the result (\ref{N_Phi}) tells
574: that a voltage pulse corresponding to $n=|\Phi|/\Phi_0$ flux units
575: pushes $n$ spinless electrons through the fork, in forward direction
576: from the source lead `s' to the prongs `u' and `d' if $\Phi>0$ and
577: in the backward direction for $\Phi<0$.
578: 
579: \section{Results}\label{sec:results}
580: 
581: Substituting the above expressions for the particle-number
582: cross-correlators and for the average number of transmitted particles
583: into (\ref{BI3}) we arrive at the following general result for
584: the Bell inequality
585: %
586: \begin{equation}
587:       E_{\scriptscriptstyle\rm BI} =
588:       \left|
589:       \frac{n+(2/\pi^2)\sin^2(\pi n) \ln(t_\mathrm{m}/\delta t)}
590:       {2n^2-n-(2/\pi^2)\sin^2(\pi n) \ln(t_\mathrm{m}/\delta t)}
591:       \right|.
592:       \label{BI4}
593: \end{equation}
594: %
595: 
596: \subsection{Pulse with integer flux}\label{sec:if}
597: 
598: For a voltage pulse with integer $n$ the above expression simplifies
599: dramatically as all logarithmic terms vanish, leaving us with
600: the Bell inequality
601: %
602: \begin{equation}
603:       E_{\scriptscriptstyle\rm BI} =
604:       \left|
605:       \frac{1}{2n-1}
606:       \right|\leq\frac1{\sqrt{2}},
607:       \label{BI5}
608: \end{equation}
609: %
610: which we find maximally violated for $n=1$ and never violated
611: for larger integers $n>1$ --- any additional particle
612: accumulated in the detector spoils the violation of the Bell
613: inequality. Furthermore, this violation is independent of
614: the transparencies $T_\mathrm{u}$, $T_\mathrm{d}$ and hence
615: universal; moreover, the Bell inequality~(\ref{BI5}) does
616: not depend on the particular form or duration of the applied
617: voltage pulse but involves only the number of electrons $n$
618: carried by the voltage pulse.
619: 
620: A voltage pulse with $n=1$ pushes two electrons with opposite spin
621: polarization towards the beam splitter. Such a pair appears in a
622: singlet state \cite{lebedev_05}
623: and can be described by the wave
624: function $\Psi_\mathrm{in}^{\scriptscriptstyle 12} =
625: \phi_\mathrm{s}^{\scriptscriptstyle 1}
626: \phi_\mathrm{s}^{\scriptscriptstyle 2}
627: \chi_\mathrm{sg}^{\scriptscriptstyle 12}$ with the spin-singlet
628: state $\chi_\mathrm{sg}^{\scriptscriptstyle 12} =
629: [\chi_{\uparrow}^{\scriptscriptstyle 1}
630: \chi_{\downarrow}^{\scriptscriptstyle 2} -
631: \chi_{\downarrow}^{\scriptscriptstyle 1}
632: \chi_{\uparrow}^{\scriptscriptstyle 2}]/\sqrt{2}$;
633: $\phi_\mathrm{s}$ is the orbital part of the wave function
634: describing a particle in the source lead `s' and the upper indices
635: 1 and 2 denote the particle number. This local spin-singlet pair
636: is scattered at the splitter and the wave function
637: $\Psi_\mathrm{in}^{\scriptscriptstyle 12}$ transforms to
638: $\Psi_\mathrm{scat}^{\scriptscriptstyle 12} = t_\mathrm{su}^2
639: \phi_\mathrm{u}^{\scriptscriptstyle 1}
640: \phi_\mathrm{u}^{\scriptscriptstyle 2}
641: \chi_\mathrm{sg}^{\scriptscriptstyle 12}+ t_\mathrm{sd}^2
642: \phi_\mathrm{d}^{\scriptscriptstyle 1}
643: \phi_\mathrm{d}^{\scriptscriptstyle 2}
644: \chi_\mathrm{sg}^{\scriptscriptstyle 12}+ t_\mathrm{su}
645: t_\mathrm{sd} [\phi_\mathrm{u}^{\scriptscriptstyle 1}
646: \phi_\mathrm{d}^{\scriptscriptstyle 2} +
647: \phi_\mathrm{d}^{\scriptscriptstyle 1}
648: \phi_\mathrm{u}^{\scriptscriptstyle
649: 2}]\chi_\mathrm{sg}^{\scriptscriptstyle 12}$, where the last term
650: describes two particles in a singlet state shared between the
651: upper and lower leads of the fork. The Bell inequality test is
652: only sensitive to pairs of particles propagating in different
653: arms, implying a projection of the scattered wave function
654: $\Psi_\mathrm{scat}^{\scriptscriptstyle 12}$ onto the spin-entangled
655: component. Thus the origin of the entanglement is found in the
656: post-selection during the cross-correlation measurement effectuated
657: in the Bell inequality test~\cite{samuelsson_04,lebedev_04}.
658: From an experimental point of view it may be difficult to produce
659: voltage pulses driving exactly one (spinless) particle $n=1$.
660: However, as follows from the full expression Eq.\ (\ref{BI4}),
661: for a sufficiently small deviation $|\delta n| =|n-1| \ll 1$ the
662: logarithmic terms are small in the parameter $(\delta n)^2$ and
663: thus can be neglected, provided the measurement time $t_m$
664: satisfies the condition $(\delta n)^2\ln (t_\mathrm{m}/\delta
665: t)\ll 1$.
666: 
667: \subsection{Weak pumping regime}\label{sec:wp}
668: 
669: The weak pumping regime with $n < 1$, corresponding to small
670: voltage pulses carrying less then one electron per spin
671: channel, deserves special attention. Inspection of (\ref{BI4})
672: shows that the Bell inequality can be formally violated in this
673: regime. We believe that this violation of the Bell inequality
674: has no physical meaning. Below, we will show that for non-integer
675: $n$ many-particle effects generate backflow of particles in our
676: setup. We argue that this backflow leads to an inconsistency
677: in the derivation of the Bell inequality itself, as the
678: assumption $0\leq |x/X|,|\bar x/X|, |y/Y|,|\bar y/Y| \leq 1$
679: may no longer hold true for $n <1$.
680: 
681: In order to understand this weak pumping regime better, we analyze
682: the sign of the full current cross-correlator ${\cal C}_\mathrm{ud}
683: (t_1,t_2)=\langle \hat I_\mathrm{u}(t_1) \hat I_\mathrm{d}(t_2)
684: \rangle$ (for one spin component); expressing this quantity
685: through the phase $\phi(t)$ we find the form
686: %
687: \begin{eqnarray}
688:       &&{\cal C}_\mathrm{ud}(t_1,t_2) =
689:       \frac{e^2}{(2\pi)^2} T_\mathrm{u} T_\mathrm{d}
690:       \nonumber\\
691:       &&\times
692:       \left[
693:       \dot \phi(t_1) \dot \phi(t_2) -4
694:       \frac{\sin^2[(\phi(t_1)-\phi(t_2)/2]}{(t_1-t_2)^2}
695:       \right].
696: \end{eqnarray}
697: %
698: We consider again the case of a narrow voltage pulse applied
699: at $t=t_0$ and assume a specific shape $\phi(t) = 2n \arctan
700: [(t-t_0)/\delta t]$. Choosing times $t_1 < t_0 < t_2$ before
701: and after the application of the pulse at $t_0$, we find that
702: the currents in the leads `u' and `d' predominantly flow in opposite
703: directions: for a sharp pulse with $|t_{1,2}-t_0| \gg \delta t$
704: we can assume that $\phi(t_1) - \phi(t_2) = 2\pi n$ and thus
705: the correlator ${\cal C}_\mathrm{ud} (t_1,t_2)$ takes the form
706: %
707: \begin{eqnarray}
708:       &&{\cal C}_\mathrm{ud}(t_1,t_2) =
709:       \frac{e^2}{\pi^2 (\delta t)^2} T_\mathrm{u} T_\mathrm{d}
710:       \nonumber\\
711:       &&\times
712:       \left[
713:       \frac{n^2}{(1+z_1^2)(1+z_2^2)}-\frac{\sin^2 \pi n}{(z_1-z_2)^2}
714:       \right],
715: \end{eqnarray}
716: %
717: where we have introduced $z_{1,2}=(t_{1,2}-t_0)/\delta t$. For
718: $|z_{1,2}|\gg 1$ the second (negative) term $\propto 1/(|z_1|
719: +z_2)^2$ describing the irreducible correlations dominates
720: over the first (positive) term $\propto 1/(z_1^2\, z_2^2)$
721: and hence the full current cross-correlator is negative. This
722: negative sign tells us that, despite application of a positive
723: voltage pulse with  $n>0$, the currents at times $t_1 < t_0
724: < t_2$ in leads `u' and `d' flow in opposite directions on
725: average. Note that this unusual behavior is a specific
726: feature of time dependent voltage pulses and does not
727: appear for a constant dc voltage with  $\phi(t) = eVt/\hbar$
728: --- in this case the current cross-correlator is
729: always positive.
730: 
731: As a consequence, the time-integrated full particle-number
732: cross-correlator (per one spin component) may turn out
733: negative as well and it does so for voltage pulses carrying
734: less then one electron per spin channel $n<1$,
735: %
736: \begin{equation}
737:       {\cal K}_\mathrm{ud}^\mathrm{ex} =
738:       e^2 T_\mathrm{u} T_\mathrm{d} \Bigl[
739:       n^2 - n -
740:       \frac{2}{\pi^2}\sin^2(\pi n)\,\ln
741:       \frac{t_\mathrm{m}}{\delta t}\Bigr].
742: \end{equation}
743: %
744: Hence, in the weak pumping regime the particles in the
745: outgoing leads `u' and `d' are preferentially transmitted in
746: opposite directions. Note that both (negative) terms in the
747: correlator, the one ($-n$) from short time differences as
748: well as the contribution ($\propto -\sin^2(n\pi)\ln
749: (t_\mathrm{m}/\delta t)$ related to the `orthogonality
750: catastrophe' dominate over the (positive) product term
751: ($n^2$), with the second one becoming increasingly important
752: at large measuring times. Furthermore, this second term
753: also drives the particle-number cross-correlator negative
754: at large non-integer $n>1$ and long measuring times, again
755: signalling the presence of particle backflow in the device.
756: 
757: The derivation of the Bell inequality relies on
758: the assumption that the quantities $|x/X|,~|y/Y|$ etc.\ are
759: bounded by unity. For our setup this implies that the
760: particle number ratios of the type $|x/X|=|(N_1-N_3)/
761: (N_1+N_3)|$ are bounded by unity, which is only guaranteed
762: for particle numbers with equal sign, $N_1\, N_3 >0$;
763: hence, particles detected in the pair of spin filters
764: with polarization $\pm{\bf a}$ in the upper arm have to
765: be transmitted in the same direction. Next, we note that
766: particles with opposite spin propagate independently and
767: hence our finding that particles preferentially propagate
768: in opposite directions of the outgoing leads `u' and `d'
769: for $n < 1$ also implies that the particle numbers $N_1$
770: and $N_3$ can be of opposite sign, hence the condition
771: $x < 1$ is not necessarily satisfied for $n < 1$.
772: On the contrary, for $n=1$ the full particle-number
773: cross-correlator (for one spin component) vanishes,
774: ${\cal K}_\mathrm{ud}=0$: in the simplest interpretation
775: we may conclude that the single transmitted particle
776: is propagating either through the upper or the lower
777: lead, thus either $N_\mathrm{u} =0$ and $N_\mathrm{d}=1$
778: or vice versa and the quantity $|x/X|$ is properly bounded.
779: The above arguments cannot exclude the relevance of
780: additional many-particle effects, i.e., the appearance of
781: additional particle-hole excitations in the system
782: contributing to the particle count in the various
783: detectors. A formal proof of the unidirectional
784: propagation of particles confirming the applicability
785: of the Bell inequality for the present non-stationary
786: situation relies on the calculation of the full
787: counting statistics of particles measured in the
788: detectors 1 and 3, etc.; such a calculation has not
789: been done yet.
790: 
791: \subsection{Many integer-flux pulses}\label{sec:mifp}
792: 
793: Above, we have concentrated on the situation where only
794: a single voltage pulse has been applied. Let us consider
795: another situation where a sequence of voltage pulses driving
796: an integer number of electrons is applied to the source lead
797: `s'.  In contrast to the previous analysis, we study the
798: total transmitted charge from $t=-\infty$ to $t=\infty$, $\hat
799: N_i(\infty)= \int_{-\infty}^\infty dt \, \hat I_i(t)$; the excess
800: part of the irreducible particle-number cross-correlator
801: takes the form [we remind that the equilibrium part can be
802: quenched in going to the four-terminal beam splitter of
803: Fig.\ \ref{fig:fork}(b)]
804: %
805: \begin{equation}
806:       K^\mathrm{ex}_\mathrm{ud} = -
807:       \frac{e^2}{\pi^2}T_\mathrm{u} T_\mathrm{d}\!\!
808:       \int\limits_{-\infty}^{\infty} \!\! dt_1 dt_2\,
809:       \frac{\sin^2[(\phi(t_1)-\phi(t_2))/2]}{(t_1-t_2)^2}.
810:       \label{ncordi}
811: \end{equation}
812: %
813: In our further analysis, we closely follow the technique
814: developed in Ref.\ \onlinecite{ll_95}. The double
815: integral in the above expression is logarithmically
816: divergent at large times $t_1,t_2$, producing the logarithmic
817: dependence on the measurement time $t_\mathrm{m}$ noted above
818: for the finite accumulation time. However, for pulses with an
819: integer number of electrons, this problematic term disappears;
820: in this case we are allowed to regularize the integral in
821: (\ref{ncordi}) with the help of
822: %
823: \begin{equation}
824:       \frac{1}{(t_1-t_2)^2}
825:       \rightarrow \frac12 \Bigl[ \frac1{(t_1-t_2+i\delta)^2}+
826:       \frac1{(t_1-t_2-i\delta)^2}\Bigr]
827:       \label{reg}
828: \end{equation}
829: %
830: and $\delta \to 0$ a small cutoff. Expressing the factor
831: $\sin^2(...)$ in Eq.~(\ref{ncordi}) in terms of exponential
832: functions, we arrive at the form
833: %
834: \begin{eqnarray}
835:       K_\mathrm{ud}^\mathrm{ex} &=& \frac{e^2
836:       T_\mathrm{u} T_\mathrm{d}}{(2\pi)^2} \int\!\! dt_1 dt_2 \,
837:       \biggl[\frac{e^{i\phi(t_1)-i\phi(t_2)}}{(t_1-t_2+i\delta)^2}
838:       \label{K_ex} \\
839:       &&\qquad\qquad\qquad\qquad\qquad
840:       +\frac{e^{i\phi(t_1)-i\phi(t_2)}}{(t_1-t_2-i\delta)^2}
841:       \biggr].\nonumber
842: \end{eqnarray}
843: %
844: In order to proceed further, we split the exponential into two
845: terms, $e^{i\phi(t)} = f_+(t)+f_-(t)$, with $f_+(t)$ and $f_-(t)$
846: two bounded analytic functions in the upper and lower complex-$t$
847: plain. Substitution into the above expression and using Cauchy's
848: formula for the derivative,
849: %
850: \[
851:    \dot f_\pm(t) = \pm \frac{i}{2\pi}
852:    \int dt' \frac{f_\pm(t')}{(t-t'\pm i\delta)^2},
853: \]
854: %
855: allows us to write the particle-number correlator in the form
856: %
857: \begin{equation}
858:       K_\mathrm{ud}^\mathrm{ex} = - \frac{e^2}{2\pi i}
859:       T_\mathrm{u} T_\mathrm{d} \int dt\, \bigl[\dot f_+(t)
860:       f_+^*(t) - \dot f_-(t) f_-^*(t) \bigr].
861:       \label{ncorf}
862: \end{equation}
863: %
864: In (\ref{ncorf}) we have made use of the analytical properties of
865: $f_\pm(t)$; in particular, with the complex conjugate functions
866: $f^*_+(t)$ and $f^*_-(t)$ bounded and analytic in the lower
867: and upper half-planes, respectively, we easily find that
868: $\int dt \, \dot f_+(t) f_-^*(t) = \int dt \, \dot f_-(t)
869: f_+^*(t) = 0$. In addition, we can also express the average
870: number of transmitted particles in terms of the functions
871: $f_\pm(t)$ introduced above,
872: %
873: \begin{eqnarray}
874:       \langle \hat N_\mathrm{u(d)} \rangle &=&
875:       \frac{e}{\pi} T_\mathrm{u(d)} \int dt\, \dot \phi(t)
876:       \label{nave}
877:       \\
878:       &=& \frac{e}{\pi i} T_\mathrm{u(d)} \int dt\, e^{-i\phi(t)}
879:       \frac{d}{dt} e^{i\phi(t)}
880:       \nonumber\\
881:       &=& \frac{e}{\pi i} T_\mathrm{u(d)} \int dt\,
882:       \bigl[ \dot f_+(t) f_+^*(t) + \dot f_-(t) f_-^*(t) \bigr].
883:       \nonumber
884: \end{eqnarray}
885: %
886: Rewriting Eqs.~(\ref{ncorf}) and (\ref{nave}) in terms of the
887: real numbers $n_\pm$,
888: %
889: \begin{equation}
890:       n_\pm = \pm \frac1{2\pi i} \int dt\, \dot f_\pm(t) f_\pm^*(t),
891:       \label{npm}
892: \end{equation}
893: %
894: we obtain the particle-number cross-correlator and the average number
895: of transmitted particles in the form
896: %
897: \begin{eqnarray}
898:       K_\mathrm{ud}^\mathrm{ex} &=& - e^2\, T_\mathrm{u}
899:       T_\mathrm{d}\, (n_+ + n_-),
900:       \label{ncorn} \\
901:       \langle N_\mathrm{u(d)}\rangle &=& 2e\,
902:       T_\mathrm{u(d)}\,(n_+-n_-).
903:       \label{naven}
904: \end{eqnarray}
905: %
906: Substituting these expressions into the Bell inequality
907: Eq.~(\ref{BI3}), we arrive at the result
908: %
909: \begin{equation}
910:       E_{\scriptscriptstyle\rm BI} = \left|
911:       \frac{n_+ + n_-}{ 2 (n_+ - n_-)^2 - (n_+ + n_-)}
912:        \right|\leq \frac1{\sqrt{2}}.
913:       \label{BI_n}
914: \end{equation}
915: %
916: The physical meaning of the numbers $n_\pm$ is easily identified
917: for the specific form of Lorentzian voltage pulses
918: %
919: \begin{equation}
920:       V(t) = \sum_i n_i\,
921:       \frac{2\hbar\gamma_i/e}{1+(t-t_i)^2\gamma_i^2},
922:       \label{vpl}
923: \end{equation}
924: %
925: where the index $i$ denotes the number of the pulse in the sequence,
926: $t_i$ is the moment of its appearance, $\gamma_i^{-1}$ is the
927: pulse width, and $n_i$ the number of spinless electrons carried
928: by the $i$-th pulse with the sign of $n_i$ defining the sign
929: of the applied voltage. Such a sequence of pulses produces
930: the phase
931: %
932: \begin{equation}
933:       e^{i\phi(t)} = \prod_i
934:       \Bigl(\frac{t-t_i-i/\gamma_i}{t-t_i+i/\gamma_i}\Bigr)^{n_i},
935:       \label{ephi}
936: \end{equation}
937: %
938: from which the decomposition into the terms $f_\pm(t)$ can be
939: found. The further analysis is straightforward for unidirectional
940: pulse sequences with all $n_i > 0$, in which case $\exp[i\phi(t)]
941: = f_+(t)$ and $n_+ = \sum_i n_i$, $n_- = 0$, or all $n_i < 0$
942: whence $\exp[i\phi(t)] = f_-(t)$ and $n_+ = 0$, $n_- = - \sum_i
943: n_i$. It then turns out \cite{ll_95,lll_96} that all results
944: for the irreducible particle-number cross-correlator
945: (\ref{ncorn}), the average currents (\ref{naven}),
946: and the Bell inequality (\ref{BI_n}) do neither depend on
947: the separation $t_{i+1} - t_i$ between the pulses nor
948: on their widths $\gamma_i^{-1}$. Furthermore, the result
949: (\ref{BI_n}) for the Bell inequality agrees with the
950: previous expression (\ref{BI5}) where a single pulse
951: is carrying $n=n_+$ (or $n_-$) electrons in one go and
952: we confirm our finding that the violation of the Bell
953: inequality is restricted to pulses containing only one
954: pair of electrons with opposite spin. Also, we note that
955: for the case of well separated pulses we can restrict
956: the accumulation time over the duration of the
957: individual pulses, in which case the Bell inequality
958: is violated for all pulses with $|n_i| = 1$.
959: 
960: Another remark concerns the case of an alternating
961: voltage signal with no net charge transport and hence
962: zero accumulated particle numbers $\langle N_{\mathrm{u(d)}}
963: \rangle = 0$. Equation (\ref{naven}) then tells us
964: that $n_+ = n_-$ and the Bell inequality (\ref{BI_n})
965: is formally violated. However, we argue that this
966: violation is again unphysical and due to the same
967: improper normalization of the basic quantities
968: $|x/X|,~|y/Y|$, etc.\ as encountered previously
969: for the case of small Faraday flux $n < 1$:
970: concentrating on the expression $x/X
971: = (N_1-N_3)/(N_1+N_3)$, we note that two pulses
972: with opposite signs allow for processes where the
973: charge driven through the two spin detectors satisfies
974: $N_1 N_3 < 0$ and hence $|x/X| > 1$, in contradiction
975: with the requirements of the lemma. Note that the manner
976: of violating the Bell inequality is quite different for the
977: physical cases involving pulses with a single particle
978: (see the discussion of single integer-flux
979: pulses in  Sec.\ \ref{sec:if} with $n=1$, or the
980: discussion of many integer-flux pulses in Sec.\
981: \ref{sec:mifp} with $n_+ = 1,~n_-=0$ and $n_+=0,~n_-
982: = 1$) and for the unphysical situation of an
983: alternating signal with $n_+ = n_-$ discussed above:
984: in the first case the small denominator results
985: from a cancellation between the product term $2n$ and
986: the negative number correlator $-1$, hence $E_{\rm
987: \scriptscriptstyle BI} = |1/(2-1)| = 1$, while in the
988: second case, the product term vanishes and there is no
989: compensation, although the final result is the same,
990: $E_{\rm \scriptscriptstyle BI} = |1/-1| = 1$. The
991: same apparent violation appears at large non-integer
992: values of $n>1$ and long measuring times, where the
993: term $\propto \sin^2(\pi n)\ln(t_\mathrm{m}/\delta t)$
994: becomes dominant, cf.\ (\ref{BI4}).
995: 
996: \section{Conclusion}\label{sec:conc}
997: 
998: The application of voltage pulses to a mesoscopic fork
999: allows to generate spin-entangled pairs of electrons
1000: through post-selection; the presence of these entangled pairs
1001: can be observed in a Bell inequality measurement based on
1002: particle-number cross-correlators. A number of items have
1003: to be observed in producing these entangled objects:
1004: {\it i)} Equilibrium fluctuations competing with the
1005: pulse signal have to be eliminated. This can be achieved
1006: with the help of a four-channel beam splitter as sketched
1007: in Fig.\ \ref{fig:fork}(b) where the channel mixing is
1008: tuned such that the transmission $T_\mathrm{ud}$ between
1009: the upper and lower channel is blocked.
1010: {\it ii)} Pulses $V(t)$ with integer Faraday flux $\Phi
1011: = -c \int dt  V(t) = n \Phi_0$ injecting an integer number
1012: of particles shall be used. Otherwise, the `fractional
1013: injection' of a particle induces a long-time perturbation
1014: in the system producing a logarithmically divergent
1015: contribution to the excess number correlator. The flux
1016: $\Phi = (n+\delta n)\Phi_0$, $n = $ an integer, associated
1017: with the voltage pulse has to be precise within the limit
1018: $(\delta n)^2 \ll 1/\ln(t_m/\delta t)$, with $t_m$ the
1019: measurement time of the pulse and $\delta t$ the pulse width.
1020: {\it iii)} The Bell inequality is violated for
1021: pulses injecting a single pair of electrons with opposite
1022: spin, i.e., pulses with one Faraday flux and hence $n=1$.
1023: The maximal violation of the inequality points to the
1024: full entanglement of the pair --- the question what
1025: type of pulses produce only partially entangled states
1026: (as quantified in terms of concurrence or negativity
1027: of the partially transposed density matrix \cite{peres_96})
1028: has not been addressed here.
1029: {\it iv)} Although weak pumping with pulses carrying less
1030: than one Faraday flux, i.e., $n <1$, formally violate
1031: the Bell inequality (note the proviso {\it ii)}, however),
1032: we associate this spurious violation with an improper
1033: normalization of the particle-number ratios $(N_i-N_j)
1034: /(N_i+N_j)$ entering the Bell inequality.
1035: {\it v)} The same argument also applies to the case of
1036: pumping with an alternating signal --- we find the Bell
1037: inequality always violated when the average injected
1038: current vanishes (i.e., when the number of carriers
1039: transmitted in the forward and backward directions are
1040: equal). Again, the origin of this spurious violation is
1041: located in the improper normalization of the particle-number
1042: ratios $(N_i-N_j)/(N_i+N_j)$ for this situation.
1043: 
1044: The above points suggest the following physical interpretation:
1045: An integer-flux pulse with Faraday flux $n\Phi_0$ extracts
1046: exactly $n$ electron pairs from the reservoir which then
1047: are tested in the Bell measurement setup. For $n=1$ we find
1048: the Bell inequality maximally violated, implying that the
1049: electrons within the pair are maximally entangled and not
1050: entangled with the remaining electrons in the
1051: Fermi sea. On the other hand, the application of a
1052: fractional-flux pulse with non-integer $n$ produces
1053: a superposition of states with different number of
1054: excess electron pairs in the fork. The electrons
1055: injected into the fork then remain entangled with
1056: those in the Fermi sea and their analysis in the
1057: Bell measurement setup makes no sense.
1058: 
1059: In our analysis of the spurious violations of Bell
1060: inequalities for weak pumping and for alternating drives we
1061: have identified the presence of reverse particle flow as
1062: the problematic element.  In the weak pumping limit this
1063: conclusion has been conjectured from the appearance of
1064: negative values in the current cross-correlator, implying
1065: negative values of the particle-number correlator for $n
1066: < 1$. Although we believe that these are strong arguments
1067: supporting our interpretation, we are not aware of a formal
1068: analysis of the backflow appearing in this type of systems.
1069: The question to be addressed then is: Given a bias signal
1070: driving particles through the device in the forward direction,
1071: what are the circumstances and what is the probability to
1072: find particles moving in the opposite direction (backflow)?
1073: A related problem has been addressed by Levitov
1074: \cite{levitov_01} (see also Ref.\ \onlinecite{makhlin_01})
1075: who has derived the full counting statistics for the charge
1076: transport across a quantum point contact in the weak $ac$
1077: pumping regime and has identified parameters producing a
1078: strictly unidirectional flow. The corresponding analysis
1079: for our system remains to be done.
1080: 
1081: A similar scheme for producing spin-entangled pairs of
1082: electrons has been discussed in Ref.~\onlinecite{lebedev_05},
1083: where a constant voltage $V$ has been applied to the source
1084: lead `s'. In this case, the source reservoir injects a regular
1085: sequence of spin-singlet pairs of electrons separated by the
1086: voltage time $\tau_V = h/eV$; the
1087: Bell inequality then is violated at short times only. The main
1088: novelty of the present proposal is the generation of
1089: well separated spin-entangled electron pairs in response to
1090: distinguished voltage pulses, thus avoiding the short time
1091: correlation measurement at time scales of order
1092: $\tau_V$.
1093: 
1094: Our mesoscopic fork device produces entangled pairs of
1095: electrons with a probability of 50 \%, i.e., half of the
1096: single-flux pulses will produce a useful pair with one
1097: spin propagating in the upper and the other in the lower
1098: channel. The competing events with both particles moving
1099: in one channel produce no useful outcome. This is similar
1100: to the finding of Beenakker {\it et al.} \cite{beenakker_05}
1101: who derive a concurrence corresponding to the production
1102: of one entangled pair per two pumping cycles. In how
1103: far this represents an upper limit in the performance
1104: of this type of devices or what type of entanglement generators
1105: are able to reach (at least ideally) 100 \% efficiency is
1106: an interesting problem.
1107: 
1108: We thank D. A. Ivanov and L. S. Levitov for discussions and
1109: acknowledge the financial support from the Swiss National
1110: Foundation (through the program MaNEP and the CTS-ETHZ), the
1111: Forschungszentrum J\"ulich within the framework of the Landau
1112: Program, the Russian Science Support Foundation, and the program
1113: 'Quantum Macrophysics' of the RAS.
1114: 
1115: \begin{thebibliography}{99}
1116: 
1117: \bibitem{lesovik_01} G.\ Lesovik, T.\ Martin, and G.\ Blatter,
1118:    Eur.\ Phys.\ J.\ B {\bf 24}, 287 (2001); 
1119:    P.\ Recher, E.V.\ Sukhorukov, and D.\ Loss, 
1120:    Phys.\ Rev.\ B {\bf 63}, 165314 (2001);
1121:    C.\ Bena, S.\ Vishveshwara, L.\ Balents, and M.P.A.\ Fisher,
1122:    Phys.\ Rev.\ Lett.\ {\bf 89}, 037901 (2002).
1123: 
1124: \bibitem{ionicioiu_01} R.\ Ionicioiu, P.\ Zanardi, and F.\ Rossi,
1125:    Phys.\ Rev.\ A {\bf 63}, 050101 (2001).
1126: 
1127: \bibitem{oliver_02} W.D.\ Oliver, F.\ Yamaguchi, and Y.\ Yamamoto, 
1128:    Phys.\ Rev.\ Lett.\ {\bf 88}, 037901 (2002); 
1129:    D.S.\ Saraga and D.\ Loss, 
1130:    Phys.\ Rev.\ Lett.\ {\bf 90}, 166803 (2003).
1131: 
1132: \bibitem{beenakker_03} C.W.J.\ Beenakker, C.\ Emary, M.\
1133:    Kindermann, and J.L.\ van Velsen, 
1134:    Phys.\ Rev.\ Lett.\ {\bf 91}, 147901 (2003).
1135: 
1136: \bibitem{samuelsson_03}  P.\ Samuelsson, E.V.\ Sukhorukov, and
1137:    M.\ B\"uttiker, 
1138:    Phys.\ Rev.\ Lett.\ {\bf 91}, 157002 (2003).
1139: 
1140: \bibitem{samuelsson_04} P.\ Samuelsson, E.V.\ Sukhorukov, and 
1141:    M.\ B\"uttiker, 
1142:    Phys.\ Rev.\ Lett.\ {\bf 92}, 026805 (2004).
1143: 
1144: \bibitem{fazio_04} L.\ Faoro, F.\ Taddei, and R.\ Fazio, 
1145:    Phys.\ Rev. B {\bf 69}, 125326 (2004).
1146: 
1147: \bibitem{lebedev_05} A.V.\ Lebedev, G.B.\ Lesovik, and G.\ Blatter, 
1148:    Phys.\ Rev.\ B {\bf 71}, 045306 (2005).
1149: 
1150: \bibitem{lebedev_04} A.V.\ Lebedev, G.\ Blatter, C.W.J.\ Beenakker, 
1151:    and G.B.\ Lesovik, 
1152:    Phys.\ Rev.\ B {\bf 69}, 235312 (2004).
1153: 
1154: \bibitem{loss_98} A.\ Barenco, D.\ Deutsch, A.\ Ekert, and R.\ Jozsa, 
1155:    Phys.\ Rev.\ Lett.\ {\bf 74}, 4083 (1995); 
1156:    D.\ Loss and D.P.\ DiVincenzo, 
1157:    Phys.\ Rev.\ A {\bf 57}, 120 (1998).
1158: 
1159: \bibitem{loss_04}  V.\ Cerletti, O.\ Gywat, and D.\ Loss,
1160:    cond-mat/0411235; 
1161:    C.\ Emary, B.\ Trauzettel, and C.W.J.\ Beenakker, cond-mat/0502550.
1162: 
1163: \bibitem{samuelsson_04b}  P.\ Samuelsson and M.\ B\"uttiker,
1164:    Phys.\ Rev.\ B {\bf 71}, 245317 (2005).
1165: 
1166: \bibitem{beenakker_05} C.W.J.\ Beenakker, M.\ Titov, and B.\ Trauzettel,
1167:    New J.\ Phys.\ {\bf 7}, 186 (2005).
1168: 
1169: \bibitem{divincenzo_05} M.\ Blaauboer and D.P.\ DiVincenzo,
1170:    cond-mat/0502060.
1171: 
1172: \bibitem{bell} J.S.\ Bell,
1173:    Physics {\bf 1},195 (1965).
1174: 
1175: \bibitem{nike_02} N.M.\ Chtchelkatchev, G.\ Blatter,
1176:    G.B.\ Lesovik, and  T.\ Martin, Phys.\ Rev.\ B {\bf 66},
1177:    161320(R) (2002).
1178: 
1179: \bibitem{clauser} J.F.\ Clauser and M.A.\ Horne, 
1180:    Phys.\ Rev.\ D {\bf 10}, 526 (1974).
1181: 
1182: \bibitem{lesovik_89} G.B.\ Lesovik,
1183:   JETP Lett.\ {\bf 49}, 592 (1989).
1184: 
1185: \bibitem{buttiker_90} M.\ B\"uttiker,
1186:    Phys.\ Rev.\ Lett.\ {\bf 65} 2901 (1990).
1187: 
1188: \bibitem{lesovik_99} G.B.\ Lesovik,
1189:   JETP Lett.\ {\bf 70}, 208 (1999).
1190: 
1191: \bibitem{blanter}  Ya.M.\ Blanter and M.\ B\"uttiker,
1192:    Phys.\ Rep.\ {\bf 336}, 1 (2000).
1193: 
1194: \bibitem{ll_94} L.S.\ Levitov and G.B.\ Lesovik,
1195:    Phys.\ Rev.\ Lett.\ {\bf 72}, 538 (1994);
1196: 
1197: \bibitem{ksp_00} A.A.\ Kozhevnikov, R.J.\ Schoelkopf, and
1198:    D.E.\ Prober, Phys.\ Rev.\ Lett.\ {\bf 84}, 3398 (2000);
1199:    L.-H.\ Reydellet, P.\ Roche, D.C.\ Glattli, B.\ Etienne,
1200:    and Y.\ Jin,
1201:    Phys.\ Rev.\ Lett.\ {\bf 90}, 176803 (2003).
1202: 
1203: \bibitem{note} The setup in Ref.\ \onlinecite{lebedev_05} involves
1204:    a short-time measurement (on a time scale $\tau \sim h/eV$)
1205:    with well separated detectors $x+y \ll \tau v_F$; under these
1206:    conditions the contribution of the equilibrium part of the 
1207:    correlator is small compared to the excess part.
1208: 
1209: \bibitem{ll_93} H.\ Lee and L.S.\ Levitov, cond-mat/9312013.
1210: 
1211: \bibitem{lll_96} L.S.\ Levitov, H.\ Lee, and G.B.\ Lesovik,
1212:    J.\ Math.\ Phys.\ {\bf 37}, 4845 (1996).
1213: 
1214: \bibitem{anderson} P.W.\ Anderson, 
1215:    Phys.\ Rev.\ Lett.\ {\bf 18}, 1049 (1967).
1216: 
1217: \bibitem{ll_95} H.\ Lee and L.S.\ Levitov, cond-mat/9507011.
1218: 
1219: \bibitem{peres_96} A.\ Peres,
1220:    Phys.\ Rev.\ Lett.\ {\bf 77}, 1413 (1996).
1221: 
1222: \bibitem{levitov_01} L.S.\ Levitov, cond-mat/0103617.
1223: 
1224: \bibitem{makhlin_01} Y.\ Makhlin and A.D.\ Mirlin,
1225:    Phys.\ Rev.\ Lett.\ {\bf 87}, 276803 (2001).
1226: 
1227: \end{thebibliography}
1228: 
1229: \end{document}
1230: 
1231: 
1232: !DSPAM:433bd706937211575118318!
1233: