1: %--------------------------------------------------------
2: %\documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
3: %\documentstyle[aps]{revtex}
4: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
5: %\documentclass[aps,prl,preprint,groupedaddress]{revtex4}
6: %\documentclass[aps,prl,preprint,superscriptaddress]{revtex4}
7: \documentclass[aps,prl,twocolumn,showpacs,groupedaddress]{revtex4}
8: % Some other (several out of many) possibilities
9: %\documentclass[preprint,aps]{revtex4}
10: %\documentclass[preprint,aps,draft]{revtex4}
11: %\documentclass[prb]{revtex4}% Physical Review B
12: \usepackage{graphicx}% Include figure files
13: \usepackage{dcolumn}% Align table columns on decimal point
14: \usepackage{bm}% bold math
15: \usepackage{psfig,graphics}
16: %\usepackage{psfig,pstricks,pst-grad,fancybox,graphics}
17: \usepackage{latexsym}
18: \usepackage{amsmath}
19: \usepackage{amssymb}
20: \usepackage{enumerate}
21: \usepackage[latin1]{inputenc}
22: \usepackage{bbm}
23: %-----------------begin new commands ------------
24: \newcommand{\C}{\ensuremath{\mathbbm C}}
25: \newcommand{\R}{\ensuremath{\mathbbm R}}
26: \newcommand{\rot}{{\cal R}}
27: \newcommand{\OO}{{\cal O}}
28: \newcommand{\Bra}[1]{\ensuremath{\langle#1|}}
29: \newcommand{\bra}[1]{\ensuremath{\langle#1|}}
30: \newcommand{\Ket}[1]{\ensuremath{|#1\rangle}}
31: \newcommand{\ket}[1]{\ensuremath{|#1\rangle}}
32: \newcommand{\BraKet}[2]{\ensuremath{\langle #1|#2\rangle}}
33: \newcommand{\braket}[2]{\ensuremath{\langle #1|#2\rangle}}
34: \newcommand{\KetBra}[1]{\ensuremath{| #1 \rangle \langle #1 |}}
35: \newcommand{\ketbra}[1]{\ensuremath{| #1 \rangle \langle #1 |}}
36: \newcommand{\KetBraO}[3]{\ensuremath{| #1 \rangle_{#3}\langle #2 |}}
37: \newcommand{\Eins}{\ensuremath{\mathbbm 1}}
38: \newcommand{\eins}{\ensuremath{\mathbbm 1}}
39: \newcommand{\HH}{\ensuremath{\mathcal{H}}}
40: \newcommand{\WW}{\ensuremath{\mathcal{W}}}
41: \newcommand{\HS}{\ensuremath{\mathcal{HS}}}
42: \newcommand{\BH}{\ensuremath{\mathcal{B(H)}}}
43: \newcommand{\kommentar}[1]{}
44: \newcommand{\HSNorm}[2]{\ensuremath{\langle #1|#2\rangle_{HS}}}
45: \newcommand{\Mean}[1]{\ensuremath{\langle #1 \rangle}}
46: \newcommand{\mean}[1]{\ensuremath{{\langle #1 \rangle}}}
47: \newcommand{\qed}{\ensuremath{\hfill \Box}}
48: \newcommand{\be}{\begin{equation}}
49: \newcommand{\ee}{\end{equation}}
50: \newcommand{\ben}{\begin{displaymath}}
51: \newcommand{\een}{\end{displaymath}}
52: \newcommand{\bea}{\begin{eqnarray}}
53: \newcommand{\eea}{\end{eqnarray}}
54: \newcommand{\bean}{\begin{eqnarray*}}
55: \newcommand{\eean}{\end{eqnarray*}}
56: \newcommand{\bc}{\begin{center}}
57: \newcommand{\ec}{\end{center}}
58: \newcommand{\bi}{\begin{itemize}}
59: \newcommand{\ei}{\end{itemize}}
60: \newcommand{\vecb}[1]{{\bf #1}}
61: \newcommand{\cf}{cf.}
62:
63: %--------------------end new commonds
64:
65: \bibliographystyle{apsrev}
66:
67: \begin{document}
68:
69: \title{Differential atom interferometry beyond the standard quantum limit}
70:
71: \author{K. Eckert\mbox{$^{1}$}, P. Hyllus\mbox{$^{1,2}$},
72: D. Bru{\ss}\mbox{$^{1,3}$}, U.V. Poulsen\mbox{$^{1,4,5}$},
73: M.~Lewenstein\mbox{$^{1,6,}$}\footnote{also at Instituci\'o
74: Catalana de recerca i estudis avan\c cats.},
75: C. Jentsch\mbox{$^{7}$}, T.
76: M{\"u}ller\mbox{$^{7}$}, E.M. Rasel\mbox{$^{7}$}, and W.
77: Ertmer\mbox{$^{7}$}}
78: \affiliation{ \mbox{$^{1}$}Institut f{\"u}r Theoretische Physik, Universit\"at Hannover, D-30167 Hannover, Germany.\\
79: \mbox{$^{2}$} The Blackett Lab -- QOLS, Imperial College London, London SW7 2BW, United Kingdom.\\
80: \mbox{$^{3}$}Institut f{\"u}r Theoretische Physik III, Universit{\"a}t D{\"u}sseldorf, D-40225 D{\"u}sseldorf, Germany.\\
81: \mbox{$^{4}$}Dipartimento di Fisica, Universit{\`a} di Trento, I-38050 Povo (TN), Italy;
82: and ECT$^*$, I-38050 Villazzano (TN), Italy.\\
83: \mbox{$^{5}$}Department of Physics and Astronomy, University of Aarhus, DK-8000 Aarhus C., Denmark.\\
84: \mbox{$^{6}$}ICFO - Institut de Ci\`encies Fot\`oniques, 08034 Barcelona, Spain.\\
85: \mbox{$^{7}$}Institut f\"ur Quantenoptik, Universit{\"a}t Hannover, D-30167 Hannover, Germany. }
86:
87: \date{\today}
88:
89: \begin{abstract}
90: We analyze methods to go beyond the standard quantum limit for a class
91: of atomic interferometers, where the quantity of interest is the
92: difference of phase shifts obtained by two independent atomic
93: ensembles. An example is given by an atomic Sagnac interferometer,
94: where for two ensembles propagating in opposite directions in the
95: interferometer this phase difference encodes the angular velocity of the
96: experimental setup.
97: We discuss methods of squeezing separately or jointly
98: observables of the two atomic ensembles, and compare in detail
99: advantages and drawbacks of such schemes. In particular we show that the method of
100: joint squeezing may improve the variance by up to a factor of 2. We take into account
101: fluctuations of the number of atoms in both the preparation and the
102: measurement stage, and obtain bounds on the difference of
103: the numbers of atoms in the two ensembles, as well as on the detection
104: efficiency, which have to be fulfilled in order
105: to surpass the standard quantum limit.
106: Under realistic conditions,
107: the performance of both schemes can be improved
108: significantly by reading out the phase difference {\em via} a
109: quantum non-demolition (QND) measurement.
110: Finally, we discuss a scheme using macroscopically entangled
111: ensembles.
112:
113: \end{abstract}
114:
115: \pacs{42.50.St, 42.50.Ct, 95.75.Kk, 42.81.Pa}
116:
117: \maketitle
118: \date{\today}
119: \newpage
120:
121: %%%%%%%%%%%%%%%%%%%%%%%%%%% the beginning %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
122: \section{I. Introduction}
123:
124: Comparing the phase shifts obtained in two independent interferometric
125: setups has several applications, prominent examples being the comparison of
126: atomic clocks \cite{atomicclocks}, and Sagnac interferometry to discriminate
127: between rotations and accelerations \cite{jentsch}. In addition, the
128: comparison of gravitational forces at different points in space or for
129: different atomic species \cite{weitz} allows to test predictions of possible
130: violations of Einstein's general relativity \cite{snadden:1998,kasevich:2002}.
131: For such {\em differential interferometers}, the
132: quantity of interest is encoded in either the difference, or the sum
133: of the individual phase shifts.
134:
135: Especially for the measurement of inertial forces, atom interferometers
136: promise high resolution. Here atom optical elements like beam splitters
137: and mirrors can be realized using Raman transitions between two
138: atomic ground state levels \cite{KasevichChu}.
139: The accumulated phase difference
140: between the interferometer paths is encoded in the number
141: difference of atoms in the exit ports labeled by different
142: internal states. For example, this can be measured by state
143: selective fluorescence detection. Such schemes have already been
144: successfully implemented to measure inertial forces and the
145: earth's gravity with a high accuracy~\cite{KasevichChu,peters}. In
146: Sagnac atom interferometry, the goal is to measure the phase shift
147: which occurs when the laser setup (laboratory frame) is rotating
148: relative to the frame of the freely flying atomic ensembles. This phase shift is given by
149: $\phi^{\rm at} =4\pi{\bf A}\;\mbox{\boldmath $\Omega$}\;m_{\rm
150: at}/h$ as compared to $\phi^{\rm light}=4\pi{\bf A}\;
151: \mbox{\boldmath $\Omega$}/(\lambda c)$ for laser interferometers,
152: where $m_{\rm at}$ is the mass of the atoms, ${\bf A}$ is the
153: oriented enclosed area of the interferometer, \mbox{\boldmath
154: $\Omega$} is the vector of angular velocity, and $\lambda$ is the wavelength of
155: the light. For an atomic gyroscope working with $^{87}$Rb, the phase
156: $\phi^{\rm at}$ is $10^{11}$ times larger than the corresponding phase
157: $\phi^{\rm light}$ of a light interferometer enclosing the
158: same area and operating at $\lambda=10^3\;$nm. Hence, atom
159: interferometers promise an enormously improved resolution for rotation
160: measurements as compared to ``classical'' photonic devices.
161:
162: The Sagnac phase can be measured by letting two
163: ensembles of atoms pass through the interferometer from
164: opposite sides. They obtain then phase shifts $\pm\phi$
165: due to the rotation of the laboratory frame, where the sign depends
166: on the direction of propagation of the ensembles, and
167: a common phase shift $\theta$ due to effects such as
168: %to environmental noise,
169: an acceleration of the setup.
170: Subtracting the phases of the two ensembles yields the desired phase
171: $2\phi$, which encodes the rotation of the setup around an axis perpendicular
172: to the plane of the interferometer.
173: Collisions between the two ensembles in the
174: interferometer can safely be neglected due
175: to the low atomic densities of the ensembles.
176:
177: In the standard quantum limit for phase measurements in atomic
178: interference experiments, the variance of the phase due
179: to quantum projection noise is given by
180: $(\Delta\phi)^2_{\rm SQL}=1/N,$ where $N$ is the number of
181: atoms in a sample. By
182: feeding the interferometer with non-classical states of atoms, the
183: so-called squeezed states, this limit can be surpassed,
184: with a fundamental bound given by
185: $(\Delta\phi)^2_{\rm H}=1/N^2,$ the so-called
186: Heisenberg limit \cite{inter,heisenberg,oblak}.
187: Squeezed atomic states can be produced by a quantum non-demolition
188: (QND) interaction of the atoms with a light beam \cite{QND}, or by
189: absorption of non-classical
190: states of the light \cite{absorbnonclassical}.
191: Squeezed states of atomic ensembles might also be useful as
192: quantum memory for states of light \cite{memory}.
193: %The QND scheme has been studied
194: %also with respect to the resistance to noise \cite{recent,kuzmichPRL}.
195:
196: It has been shown that for such squeezed states atoms within an ensemble are
197: entangled with each other \cite{Soerensen}. In addition to this
198: entanglement on a microscopic level, it is also possible to
199: entangle macroscopic degrees of freedom of two atomic ensembles
200: with a similar interaction \cite{Duan1,Polzik}. This in principle
201: enables teleportation of the macroscopic state of an ensemble.
202: Furthermore, it has been shown recently that such macroscopically
203: entangled ensembles can improve the efficiency of measurements
204: of the components of a magnetic field \cite{Vivi}.
205: Here, we try to exploit microscopic as well as macroscopic
206: entanglement between two atomic ensembles in the context
207: of differential interferometry, where we focus especially on an
208: atomic Sagnac interferometer setup.
209:
210: In Section II we calculate the phase variance for a differential
211: interferometer using non-squeezed coherent states in order to
212: introduce our methodology. We show also how to include number
213: fluctuations into the calculations. These will turn out to be important later.
214: In Section III we show that the phase uncertainty can be reduced by
215: feeding the interferometer with individually squeezed ensembles. In
216: Section IV we consider squeezing of a joint observable of both ensembles.
217: In Section V we discuss how decoherence affects the interferometer,
218: and in Section VI we compare the variances of the
219: schemes discussed so far for realistic parameters. Finally, in
220: Section VII the use of macroscopically entangled states in the
221: interferometer is discussed.
222:
223: %==========================================================
224: \section{II. Coherent input states}
225: For a single atom we define a pseudo spin-$1/2$ through two ground state
226: atomic hyperfine levels $\ket{1}$ and $\ket{2}$ \cite{jentsch,kasevich:2002}
227: by introducing the relevant spin operators as
228: \begin{eqnarray}\label{eqs:defsigma}
229: \hat{\sigma}_z&=&\frac{1}{2}\left(\ket{1}\bra{1}-\ket{2}\bra{2}\right)\;,
230: \quad
231: \hat{\sigma}_x=\frac{1}{2}\left(\ket{1}\bra{2}+\ket{2}\bra{1}\right)\:,
232: \nonumber\\
233: \hat{\sigma}_y&=&\frac{1}{2i}\left(\ket{1}\bra{2}-\ket{2}\bra{1}\right).
234: \end{eqnarray}
235: We will subsequently only consider the collective spin
236: $\hat{\bf J}=\sum_{i=1}^{N_J}\hat{\mbox{\boldmath $\sigma$}}{\,(i)}$
237: for the first, and in analogy $\hat{\bf L}$ for the second ensemble;
238: $N_{J,L}$ are the number of atoms in the ensembles $J$ and $L$, respectively.
239: The measurements that can typically be performed
240: in atom interferometry are population measurements
241: of the levels $\ket{1}$ and $\ket{2}$ by fluorescence
242: techniques \cite{Kasevich}. Hence, only the $z$
243: components of the collective spin vectors can be
244: measured directly.
245:
246: \subsection{II.A. Description of the interferometer}
247:
248: Initially, all the atoms are assumed to be prepared in the
249: state $\ket{1}$, leading to the following expectation
250: value and variance of the collective spin vector:
251: \bea
252: \mean{\hat{\bf J}}=\frac{N_J}{2}\hat{\bf z},\quad
253: (\Delta \hat{J}_z)^2=0,\quad (\Delta \hat{J}_{x,y})^2=\frac{N_J}{4},
254: \label{eqs:Jin}
255: \eea
256: Analogous values are obtained for the ensemble $\hat{\bf L}$.
257: %%The state from
258: %%Eq.~\ref{eqs:Jin} can be represented by a cone in three
259: %%dimensional space, see Fig.~\ref{fig:Jin}.
260: Corresponding to the
261: definitions in Eq.~(\ref{eqs:defsigma}), the $z$ axis denotes the
262: difference of the number of atoms in states $\ket{1}$ and
263: $\ket{2}$, whereas the phase difference between these two states
264: is encoded in the $x-y$ plane. The uncertainties stem from
265: the single particle uncertainties, and correspond to a state with a {\em fixed}
266: number of atoms.
267: %
268: %%\begin{figure}[h]
269: %%\includegraphics[width=4cm]{Jin}
270: %%\caption{Initial state of the atomic ensembles.}
271: %%\label{fig:Jin}
272: %%\end{figure}
273: %
274:
275: A typical interferometer sequence used for the measurement of
276: inertial forces consists of three atom--light interactions as
277: shown in Fig.~\ref{fig:AI}. The first beam splitting %$\pi/2$
278: Raman pulse transfers all the atoms from the ground state
279: $\ket{1}$ to the superposition
280: $\frac{1}{\sqrt{2}}(\ket{1}+\ket{2})$. Atoms transfered to state
281: $\ket{2}$ obtain a momentum kick of two photon recoil
282: if the two Raman lasers are counter-propagating, so
283: that the partial waves delocalize, as depicted in Fig.~\ref{fig:AI}. The
284: second pulse exchanges the populations,
285: $\ket{1}\leftrightarrow\ket{2}$, and deflects the partial waves.
286: Finally, they are recombined in the last interaction zone, acting as a
287: beam splitter. %again a $\pi/2$
288: %pulse.
289: %
290: \begin{figure}[th]
291: \includegraphics[width=6.5cm]{AI_bw}
292: \caption{Scheme of the atom interferometer. $\frac{\pi}2$ and $\pi$
293: label the beamsplitting and the mirror pulses, respectively. The populations
294: in the two exit ports are detected state-selectively {\it via} fluorescence
295: measurements.
296: }
297: \label{fig:AI}
298: \end{figure}
299: %
300: These pulses can be represented as rotations of the collective
301: spin vectors $\mean{\hat{\bf J}}$ and $\mean{\hat{\bf L}}$ around an axis in the $x-y$
302: plane, the angle being given by $\pi/2$ for beam splitters and by
303: $\pi$ for mirrors. For a fixed coordinate system the angle between
304: the $x$ axis and the rotation axis is given by the laser phase
305: \cite{atominterferometry}. The laser phases change if the setup (laboratory frame)
306: rotates with respect to the path of the freely flying ensembles,
307: which causes the Sagnac phase shift. This
308: change corresponds to a rotation of the collective spin vectors in
309: the $x-y$ plane around the $z$ direction.
310:
311: In order to make the scheme applicable to a more general scenario for
312: differential interferometers, we model it as follows for each of the
313: ensembles labeled by $\hat{\bf J}$ and $\hat{\bf L}$, respectively:
314: the first beam splitter rotates each collective spin
315: vector by $\pi/2$ around the $y$ axis, then we collect all the
316: phase shifts occurring in the interferometer in rotations around $z$
317: of $\hat{\vecb{J}}$ and $\hat{\vecb{L}}$ by $\Phi_J$ and $\Phi_L$,
318: respectively. Finally, a $\pi$ and a $\pi/2$ pulse, both
319: around $x$, implement the mirror and the final
320: beam splitter, respectively. These last two pulses can
321: be combined to a single rotation around the $x$ axis by $-\pi/2$.
322: %
323: \begin{figure}[h]
324: %\includegraphics[width=7cm]{CS}
325: \includegraphics[width=0.9\columnwidth]{CS_bunt}
326: \caption{(Color online) Interferometric scheme for a coherent input state:
327: (a) collective spin after the first beam splitter, (b) the total phase $\phi$
328: accumulated in the interferometer results in a rotation around $z$,
329: before (c) a final rotation around $x$ by $-\pi/2$ implements the
330: mirror and the final beam splitter, encoding the phase in the
331: $z$ direction.}
332: \label{fig:CS}
333: \end{figure}
334: %
335: All relevant interferometric steps are described in Fig.~\ref{fig:CS},
336: where the atomic spin vector is depicted together with a disc
337: representing the corresponding uncertainties of the spin operators.
338: To simplify notation we will take the state after
339: the first beam splitter as the initial state:
340: $\hat{\vecb{J}}^{\rm in}=\rot_y(\pi/2)\hat{\vecb{J}}$, see Fig. \ref{fig:CS}(a).
341: Here $\rot_{i}(\alpha)$, $i\in\{x,y,z\}$, is the matrix rotating a vector
342: by the angle $\alpha$ around the direction $\hat{\bf e}_i$.
343: In the Heisenberg picture, and neglecting collisions between the atoms
344: within the ensemble, the spin operator changes according to
345: $\hat{{\bf J}}^{\rm out}=\rot_x(-\pi/2)\rot_z(\Phi_J)\hat{{\bf J}}^{\rm in}$.
346: This leads to
347: \be
348: \hat{{\bf J}}^{\rm out}=\left(
349: \begin{array}{c}
350: \hat{J}_x^{\,\rm in}\cos\Phi_J -\hat{J}_y^{\,\rm in}\sin\Phi_J \\
351: \hat{J}_z^{\rm in}\\
352: -\hat{J}_x^{\,\rm in}\sin\Phi_J -\hat{J}_y^{\,\rm in}\cos\Phi_J
353: \end{array}
354: \right).
355: \ee
356: It is also possible to consider a balanced atom interferometer
357: in which all rotations are around the $x$ axis. In this
358: case, an extra $\pi/2$ shift around $z$ leads to the same result.
359: The advantage of the extra pulse is that for the initial state of
360: Eq.~(\ref{eqs:Jin})
361: \be
362: \mean{\hat{J}_z^{\rm out}}=-\frac{N_J}{2}\sin\Phi_J\approx-\frac{N_J}{2}\Phi_J
363: \ee
364: for small angles $\Phi_J$, while with the unmodified balanced
365: scheme $\mean{\hat{J}_z^{\rm out}}$ would be proportional
366: to $\cos\Phi_J$, and hence sensitive to $\Phi_J$
367: only in second order around $\Phi_J=0$.
368:
369: As explained in the introduction, the phase
370: will be given by $\Phi_J=\phi+\theta$ for $\hat{\bf J}$
371: and by $\Phi_L=-\phi+\theta$ for $\hat{\bf L}$, because the
372: Sagnac phase $\phi$ takes a different sign, depending on
373: direction in which the ensemble passes the interferometer.
374: We define the phase operator for the
375: case of coherent states (cs) as
376: \be
377: \hat{\phi}_{\rm cs}=-\frac{\hat{J}_z^{\rm out}}{N_J}+\frac{\hat{L}_z^{\rm out}}{N_L}
378: \label{eqs:phiCS}
379: \ee
380: and obtain to the first order in $\phi$ and $\theta$
381: \bea
382: \mean{\hat{\phi}_{\rm cs}}&=&\phi \\
383: (\Delta\hat{\phi}_{\rm cs})^2&=&\frac{1}{4N_J}+\frac{1}{4N_L}.
384: \eea
385: The variance of $\hat{\phi}_{\rm cs}$ corresponds to
386: the standard quantum limit. Note that
387: $\theta$ can be obtained in an analogous way.
388: %the same way merely by changing a sign.
389: From the definition it is obvious that determining the
390: number of atoms is important for the calculation of the
391: phase shift. A major source of error will come from fluctuations
392: in the number of atoms of the ensembles and hence we will discuss
393: how to include this process into the calculations in the next section.
394:
395: %----------------------------------------------------------------
396: \subsection{II.B. Number fluctuations}
397: There are two sources of deviations of the number of atoms:
398: the preparation process and the number measurement process. The atomic
399: ensembles produced from the source are best described as a statistical mixture
400: of states with different atom numbers, but we will assume that the
401: final number measurement projects
402: onto a number state with $N_J$ and $N_L$ atoms in the two
403: ensembles. Defining $\bar{N}=(N_J+N_L)/2$, we assume that
404: $|N_J-N_L|=\gamma\sqrt{\bar{N}}$, which reflects the variances of
405: the number operators $\hat{N}_{J,L}$ of the ensembles after the
406: production. Thus, $\gamma$ is the parameter which describes how
407: well the atom numbers in the two ensembles match.
408:
409: We treat the number measurements by introducing operators
410: $\delta \hat{N}$ with
411: \be
412: \mean{\delta \hat{N}}=0,\quad [\Delta(\delta \hat{N})]^2=\alpha N,
413: \ee where $\alpha$ describes the quality of the number
414: measurement. For fluorescence measurements, $\alpha^{-1}$ is given
415: by the mean number of times an atom goes through the fluorescence
416: cycle and scatters a photon which subsequently is registered in the detectors
417: \cite{numbers}. Typical values in nowadays experiments
418: are $\alpha^{-1}\approx 50\ldots100$. We replace $N_J$ by
419: $N_J^0+\delta \hat{N}_{J}^{(1)}+\delta \hat{N}_{J}^{(2)}$ and $\hat{J}_z^{\rm out}$ by
420: $\hat{J}_z^{{\rm out},0}+(\delta \hat{N}_{J}^{(1)}-\delta \hat{N}_{J}^{(2)})/2$. $N_J^0$
421: refers to the actual number that would have been
422: measured in perfect number projection measurements, and in $\delta\hat{N}_J^{(i)}$
423: the index $i$ corresponds to the atomic levels $\ket{i}$.
424:
425: With these substitutions, we obtain
426: \bea
427: \mean{\hat{\phi}_{\rm cs}}&=&\phi
428: -\alpha\Big(\frac{1}{N_J^0}+\frac{1}{N_L^0}\Big)\\
429: (\Delta\hat{\phi}_{\rm cs})^2&=&\frac{1}{4N_J^0}+\frac{1}{4N_L^0}
430: +\alpha\Big(\frac{1}{N_J^0}+\frac{1}{N_L^0}\Big),
431: \eea
432: where terms of higher order have been neglected.
433: The contribution from the number fluctuations is in
434: agreement with Ref.~\cite{numbers}.
435: The expectation value of $\hat{\phi}_{\rm cs}$
436: is shifted for $\alpha\neq 0$. As this shift is
437: of the order of the second term in
438: $(\Delta\hat{\phi}_{\rm cs})^2$, it is of second order
439: of the standard deviation only,
440: and can thus safely be neglected.
441: From the expressions it is clear that $\alpha\ll 1$ is
442: required in order to reach the fundamental limit
443: for the phase resolution using coherent input states.
444:
445: In the following, we will drop the superscript $0$ on the
446: atom number, as long as there is no danger of confusion.
447:
448: %================================================================
449: \section{III. Separately squeezed ensembles}
450:
451: It is known that by taking squeezed input states
452: it is possible to surpass the standard quantum limit
453: in interferometry \cite{inter}. In this section, we
454: consider the case where both ensembles are squeezed
455: separately with the method introduced in \cite{QND},
456: {\it i.e.}, by a QND interaction with a laser beam shortly
457: after the first beam splitter.
458:
459: \subsection{III.A. Squeezing a single ensemble}
460:
461: We assume to have the situation of Fig.~\ref{levels},
462: {\em i.e.}, the electromagnetic field mode $a_{1}$
463: couples states $\ket{1}$ and $\ket{3}$,
464: and $a_{2}$
465: couples states $\ket{2}$ and $\ket{4}$.
466: Here transitions $\ket{1}\leftrightarrow\ket{4}$
467: and $\ket{2}\leftrightarrow\ket{3}$
468: have to be suppressed to the first order in the
469: coupling constant (\cite{QND}, see also the discussion
470: in Section V).
471: %[Here we follow the
472: %first calculation of \cite{QND} which is for
473: %an ensemble in a running mode cavity. The second
474: %approach deals with flying ensembles through
475: %which a laser is sent. Effectively, both approaches
476: %should reach the same result except for a $Q$ factor
477: %from the cavity somewhere is the formulas.]
478: \begin{figure}[h]
479: \includegraphics[width=4.5cm]{levels2}
480: \caption{For the QND interaction, the levels $\ket{1}$ and $\ket{2}$
481: are coupled off-resonantly to states $\ket{3}$ and $\ket{4}$.}
482: \label{levels}
483: \end{figure}
484: For the light, an effective spin vector can be defined, the so-called
485: Stokes vector. Its components are given by
486: %The light can also be described by a spin vector through
487: \begin{eqnarray}
488: \hat{S}_z&=&\frac{1}{2}\left(\hat{a}_1^{\dagger}\hat{a}_1-\hat{a}_2^{\dagger}\hat{a}_2\right),
489: \
490: \hat{S}_x=\frac{1}{2}\left(\hat{a}_1^{\dagger}\hat{a}_2+\hat{a}_2^{\dagger}\hat{a}_1\right),
491: \\
492: \hat{S}_y&=&\frac{1}{2i}\left(\hat{a}_1^{\dagger}\hat{a}_2-\hat{a}_2^{\dagger}\hat{a}_1\right),
493: \end{eqnarray}
494: where $\hat{a}_1^{\dagger}$ and $\hat{a}_2^{\dagger}$ create a photon in mode $a_1$ and $a_2$,
495: respectively.
496: %If $a_1$ and $a_2$ correspond to orthogonal linearly
497: %polarized modes,
498: $\hat{S}_z$ measures the difference of photons in the two modes.
499: %$S_x$ the difference of photons with $45^{\circ}$ or
500: %$135^{\circ}$ linear polarization, and $S_y$ the difference
501: %of photons with orthogonal circular polarization.
502: By an appropriate choice of the parameters, the
503: Hamiltonian describing the interaction of the light with the
504: first atomic
505: ensemble can be brought to the form \cite{QND}
506: \begin{equation}
507: \hat{H}=\hbar\Omega_J \hat{S}_z \hat{J}_z,
508: \label{HQND}
509: \end{equation}
510: with a frequency $\Omega_J$.
511: Due to the interaction, the collective spin vector of the
512: atoms is rotated around the $z$ axis by $\chi_J \hat{S}_z$,
513: % $\rot_z(\chi_J \hat{S}_z)$
514: %=\left(
515: % \begin{array}{ccc}
516: % \cos(\chi_J \hat{S}_z) & - \sin(\chi_J \hat{S}_z) & 0\\
517: % \sin(\chi_J \hat{S}_z) & \cos(\chi_J \hat{S}_z) & 0\\
518: % 0 & 0 & 1
519: % \end{array}
520: % \right),
521: %\ee
522: with the atom--photon coupling $\chi_J=\Omega_J t$ and the effective interaction
523: time $t$.
524: The Stokes vector undergoes the same variation with
525: $\chi_J \hat{S}_z$ replaced by $\chi_J \hat{J}_z$. This
526: rotation of the Stokes vector is due to the Faraday effect of the light
527: passing the atoms, while the rotation of the atomic spin vector is due
528: to an AC Stark shift originating from the light field. The coupling
529: $\chi_J$ is given by \cite{QND}
530: \be
531: \chi_J=2 g^2\frac{L}{c}\frac{\Delta}{\frac14\Gamma^2+\Delta^2},
532: \quad g=\sqrt{\frac{\omega d^2}{2\hbar\epsilon_0 A L}},
533: \label{eq:chi}
534: \ee
535: where $A$ and $L$ are cross section and length of the atomic
536: sample, $\Gamma$ and $d$ are line-width and dipole moment
537: of the atomic transition, respectively,
538: and $\Delta$ is the detuning from
539: the atomic resonance frequency $\omega$.
540:
541: After the interaction,
542: \be
543: \hat{S}_y^{\rm out}=\sin(\chi_J \hat{J}_z)\hat{S}_x^{\rm in}+\cos(\chi_J \hat{J}_z)\hat{S}_y^{\rm in},
544: \ee
545: and if initially
546: \bea
547: \mean{\hat{\bf S}^{\rm in}}=\frac{n_{\! J}}{2}\hat{\bf x},\quad
548: (\Delta \hat{S}_x^{\rm in})^2=0,\quad (\Delta \hat{S}_{y,z}^{\rm in})^2
549: =\frac{n_{\! J}}{4},
550: \label{eqs:Sin}
551: \eea
552: where $n_{\! J}$ is the number of photons in the ensemble $\hat{\vecb{J}}$,
553: then we can effectively replace $\hat{S}_x^{\;\rm in}$ by its macroscopic
554: expectation value $n_{\! J}/2$. Furthermore, developing the
555: trigonometric expressions and assuming that
556: $N_J\chi_J^2\ll 1$, we obtain in leading order
557: \be
558: \hat{S}_y^{\rm out}\approx\frac{n_{\! J}\chi}{2}\hat{J}_z^{\rm in}+\hat{S}_y^{\rm in}.
559: \ee
560: Hence a measurement of the $y$ component of the
561: outgoing light vector gives information about $\mean{\hat{J}_z}$,
562: while $\hat{J}_z$ itself is not affected by the rotation around
563: the $z$ axis.
564:
565: If such a QND measurement is performed after the
566: first beam splitter of the interferometer, then the operator
567: \be
568: \hat{J}'_z\equiv\hat{J}_z^{\rm out}
569: -\frac{2}{n_{\! J}\chi_J}\hat{S}_y^{\rm out}
570: \ee
571: measures the difference between the $z$ component of
572: the ensemble's atomic spin vector after the
573: second beam splitter and the estimated value after the
574: first one. The fluctuations of this operator are
575: reduced as compared to $\hat{J}_z^{\rm out}$ \cite{QND},
576: while the fluctuations of $\hat{J}_y$ are enlarged,
577: which is depicted in Fig.~\ref{fig:SS}. Hence,
578: the state of the atomic spin vector is squeezed
579: in the $z$ direction.
580:
581: %In effect, this knowledge can be used by replacing
582: %$\hat{J}_z^{\rm out}\to \hat{J}_z^{\rm out}
583: %-\frac{2}{n_J\chi_J}\hat{S}_y^{\rm out}$,
584: %because only the difference between the $z$ component
585: %$\frac2{n_J\chi_J}\hat{S}_y^{\rm out}$
586: %of the atomic spin vector after the projection measurement of the light
587: %directly before the interferometer, and $\hat{J}_z^{\rm out}$ is of
588: %interest to the phase measurement.
589:
590: \subsection{III.B. Modified interferometric scheme}
591:
592: We modify the scheme introduced in Section II by
593: inserting the QND interaction shortly after the first
594: beam splitter, followed by an extra rotation
595: around $x$ by $\pi/2$, which rotates the
596: uncertainty ellipse such that the phase uncertainty is reduced
597: as desired, cf. Fig. \ref{fig:SS}.
598: In the experiment, the latter pulse must not transfer
599: momentum to the particles, which can be achieved
600: by using co-propagating Raman lasers for this step,
601: provided that the two transition frequencies are
602: approximately equal, so that the two recoil momenta
603: cancel in the transitions.
604: %by shining in the two lasers from the same side.
605: \begin{figure}[h]
606: %\includegraphics[width=\linewidth]{SS}
607: \includegraphics[width=0.9\columnwidth]{SS_bunt}
608: \caption{(Color online) The interferometric scheme for squeezed input states:
609: (a) collective spin after the first beam splitter, (b) the
610: QND measurement prepares a spin squeezed state which (c) is rotated
611: around $x$ by $\pi/2$; step (d)$\to$(e) as step (b)$\to$(c) in
612: Fig.~\ref{fig:CS}.}
613: \label{fig:SS}
614: \end{figure}
615:
616: The outgoing spin vector now is calculated as
617: $\hat{\bf J}^{\rm out}=\rot_x(-\pi/2)\rot_z(\Phi_J)\rot_x(\pi/2)
618: \rot_z(\chi \hat{S}_z)\hat{\bf J}^{\rm in}$,
619: %where ${\bf J}^{\rm in}$
620: %is again assumed to be the state of Eq.~(\ref{eqs:Jin})
621: %rotated by $\rot_y(\pi/2)$ to point along the $x$ axis,
622: and in analogy for the second ensemble $\hat{\bf L}$, with
623: corresponding Stokes vector $\hat{\bf T}$.
624:
625: In comparison to Eq.~(\ref{eqs:phiCS}), $\hat{J}_z^{\rm out}$
626: now has to be corrected to incorporate the spin squeezing (ss)
627: as described above:
628: \bea
629: \hat{\phi}_{\rm ss}&=&-\frac{1}{N_J}\Big(\hat{J}_z^{\rm out}
630: -\frac{2}{n_{\! J}\chi_J}\hat{S}_{y}^{\rm out}\Big)\nonumber\\
631: &&+\frac{1}{N_L}\Big(\hat{L}_z^{\rm out}
632: -\frac{2}{n_{\! L}\chi_L}\hat{T}_{y}^{\rm out}\Big).
633: \label{eqs:phiSS}
634: \eea
635: Calculating the expectation value and variance as
636: before yields
637: \bea
638: \mean{\hat{\phi}_{\rm ss}}&=&\phi
639: -\alpha\Big(\frac{1}{N_J}+\frac{1}{N_L}\Big)
640: \label{eqn:meansepsq} \\
641: (\Delta\hat{\phi}_{\rm ss})^2&=&\frac{1}{n\chi^2}
642: \Big(\frac{1}{N_J^2}+\frac{1}{N_L^2}\Big)
643: +\alpha\Big(\frac{1}{N_J}+\frac{1}{N_L}\Big).
644: \label{eqn:varsepsq}
645: \eea
646: Here it has been assumed $n_J=n_L=:n$ and $\chi_J=\chi_L=:\chi$.
647: Deviations from these assumptions enter the variance only in higher
648: order terms as long as $||\chi_J|-|\chi_L||/(|\chi_J|+|\chi_L|)\ll1$,
649: and similar for $n_J$.
650: Further assumptions leading to these expressions are \mbox{$\bar{N}\chi^2\ll 1$}
651: as mentioned before, as well as \mbox{
652: $n\chi^2\ll \sqrt{8}\cdot[\sqrt{\bar{N}}(\theta^2+\phi^2)]^{-1}$}.
653: Now, provided that $\alpha$ is small enough, the variance is dominated
654: by the first two terms scaling as $N_{J/L}^{-2}$. They originate from the
655: projection noise of the light, and in principle allow
656: to improve the resolution below the standard quantum limit. However,
657: $n\chi^2$ equals, except for a factor of order unity, the fraction of
658: atoms which are lost due to spontaneous processes during the squeezing
659: process, \cf~Section V. Thus,
660: $n\chi^2\ll1$ is necessary, and even though we obtained a Heisenberg-like
661: scaling $(\Delta\hat{\phi}_{\rm ss})^2\sim 1/N_{J,L}^2$,
662: we are far from reaching the Heisenberg limit.
663:
664: Furthermore, as it becomes clear from the second term, the resolution
665: is limited by the accuracy of the fluorescence number measurements.
666: These measurements are necessary in any case to determine the
667: phase, because $\mean{\hat{J}_z^{\rm out}}$ and $\mean{\hat{L}_z^{\rm out}}$
668: have to be rescaled properly. However, $\hat{J}_z$ and $\hat{L}_z$ itself
669: can be measured using another QND interaction. As will be shown in the next
670: section, this reduces the dependence on the quality of
671: the number measurements.
672:
673: %--------------------------------------------------------------------
674: \subsection{III.C. QND output measurement}
675: Let us consider now a modification of the scheme using
676: squeezed states, where a second QND laser beam is sent
677: through each ensemble shortly after the last beam splitter.
678: We define
679: \bea
680: \hat{\phi}_{\rm ss+}=-\frac{1}{N_J}\frac{2}{n\chi}
681: \Big(\hat{S}_{y,r}^{\rm out}-\hat{S}_{y}^{\rm out}\Big)\nonumber\\
682: +\frac{1}{N_L}\frac{2}{n\chi}\Big(\hat{T}_{y,r}^{\rm out}
683: -\hat{T}_{y}^{\rm out}\Big)
684: \label{eqs:phiSSQ},
685: \eea
686: where the extra index $+$ is supposed to indicate the
687: additional QND measurement. Furthermore, $\hat{\vecb{S}}_r$ and
688: $\hat{\vecb{T}}_r$ correspond to individually
689: prepared light pulses used to read out $\hat{J}_z$ and $\hat{L}_z$,
690: respectively.
691: The resulting expectation value and variance in leading order are
692: \bea
693: \mean{\hat{\phi}_{\rm ss+}}&=&\phi
694: +\alpha\theta\Big(\frac{1}{N_J}+\frac{1}{N_L}\Big) \\
695: (\Delta\hat{\phi}_{\rm ss+})^2&=&\frac{2}{n\chi^2}
696: \Big(\frac{1}{N_J^2}+\frac{1}{N_L^2}\Big)
697: \nonumber
698: \\
699: & & +\alpha(\theta^2+\phi^2)\Big(\frac{1}{4N_J}+\frac{1}{4N_L}\Big).
700: \label{eqs:varSSQ}
701: \eea
702: Let us compare the modified scheme including an additional
703: QND measurement with the scheme using squeezed states.
704: The magnitude of $\alpha$ in the variance of the modified scheme
705: is effectively reduced for small $\theta$ and $\phi$,
706: $\alpha\to\alpha(\theta^2+\phi^2)/4$,
707: as compared to Eq.~(\ref{eqn:varsepsq}). Hence, the dependence
708: on the number measurement is reduced.
709: Furthermore, the leading term shifting the expectation value
710: from the desired result $\phi$ is smaller by a factor $\theta$ as
711: compared to Eq.~(\ref{eqn:meansepsq}).
712:
713: As a further possible advantage, the effect of a non-symmetric
714: atom-light interaction is compensated if the coupling of the read-out
715: QND pulse is similar to the coupling of the squeezing pulse \cite{kuzmichPRL}.
716: However, this advantage is probably not very relevant
717: for the Sagnac interferometer considered here due to the
718: long time-of-flight of the ensembles in the interferometer
719: between the two QND pulses.
720:
721: A disadvantage of this scheme is that the first term of the variance
722: of Eq.~(\ref{eqs:varSSQ}) comes with
723: a factor of $2$ because of the two projection measurements of the light
724: necessary per atomic ensemble. It is possible, although technically demanding,
725: to reduce this contribution by re-using the light from the first QND
726: interaction for the read-out. In this case,
727: $\chi\to-\chi$ is needed in the second interaction
728: in order to obtain the difference
729: $\hat{J}_z^{\rm out}-\hat{J}_z^{\rm in}$ \cite{kuzmichPRL},
730: and in analogy for the second ensemble.
731: The sign can also be achieved by a $\pi$ rotation of the
732: atom spin vector around the $x$ axis in between the final beam splitter
733: and the QND read-out pulse.
734:
735: %This scheme is compared to the previous ones in Section VI.
736:
737: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
738: \section{IV. Jointly squeezed ensembles}
739:
740: In the preceding section we have seen that the $1/(n\chi^2N_J^2)$ term in the
741: variance comes with a factor given by the number of QND interactions
742: with \emph{different} ensembles of light, \cf~Eq.~(\ref{eqs:varSSQ}).
743: For this reason let us consider the case of preparing
744: the initial state of the two atomic ensembles with only
745: a single QND pulse that interacts with both ensembles consecutively.
746: The first interaction (with ensemble $\hat{\vecb{J}}$) transforms
747: $\hat{\vecb{J}}^{\text{in}}\rightarrow\rot_z(\chi \hat{S}_z^{\rm in})\hat{\vecb{J}}^{\text{in}}$,
748: the second interaction (with ensemble $\hat{\vecb{L}}$) transforms
749: $\hat{\vecb{L}}^{\text{in}}\rightarrow\rot_z(\chi \hat{S}_z^{\rm in})\hat{\vecb{L}}^{\text{in}}$,
750: because $\hat{S}_z^{\rm in}$ itself remains unchanged during the QND interaction. The
751: Stokes vector $\hat{\vecb{S}}$ transforms as
752: $\hat{\vecb{S}}^{\text{out}}=\rot_z(\chi \hat{L}_z^{\rm in})
753: \rot_z(\chi \hat{J}_z^{\rm in})\hat{\vecb{S}}^{\text{in}}$,
754: and the $y$ component of the outgoing light is given by
755: \bea
756: \hat{S}_y^{\text{out}}&=&\cos(\chi \hat{L}_z^{\rm in})
757: \left[\cos(\chi \hat{J}_z^{\,\rm in})\hat{S}_y^{\,\rm in} + \sin(\chi \hat{J}_z^{\,\rm in})\hat{S}_x^{\,\rm in}\right]+\nonumber\\
758: &&\,\sin(\chi \hat{L}_z^{\rm in})
759: \left[\cos(\chi \hat{J}_z^{\,\rm in})\hat{S}_x^{\,\rm in} - \sin(\hat{J}_z^{\,\rm in}\chi)\hat{S}_y^{\,\rm in}\right],
760: \eea
761: such that for $N\chi^2\ll1$, and $\hat{\vecb{S}}$ initially prepared as in
762: Eq.~(\ref{eqs:Sin}), we have
763: \be
764: \hat{S}_y^{\text{out}}\approx\frac{n\chi}2(\hat{J}_z^{\,\rm in}+\hat{L}_z^{\rm in})+\hat{S}_y^{\,\rm in}.
765: \label{eqn:Syjs}
766: \ee
767: Measuring $\hat{S}_y^{\text{out}}$ thus reveals information about $\hat{J}_z^{\rm in}+\hat{L}_z^{\rm in}$
768: and performs a squeezing operation on this joint operator. Now we
769: apply the same operations as before to the ensemble $\hat{\vecb{J}}$,
770: but for the ensemble $\hat{\vecb{L}}$ we perform a rotation by $\pi$
771: around the $x$ axis before the final measurement. After the extra pulse,
772: the Sagnac shift $\phi$ is effectively encoded
773: in the sum of the $z$ components instead of into the
774: difference.
775:
776: %I HAVE TAKEN OUT FIGURE 6!
777: %\begin{figure}[h]
778: %\includegraphics[width=\linewidth]{Joint}
779: %\caption{The interferometric scheme for squeezing $J_z+L_z$.
780: % Due to the extra $\pi$ pulse on $\hat{\bf L}$, the last
781: % pulse is converted to a $+\pi/2$ pulse.}
782: %\label{fig:JS}
783: %\end{figure}
784:
785: We define the phase operator for the scheme employing
786: jointly squeezed (js) ensembles as
787: \be
788: \hat{\phi}_{\rm js}=-\Big(\frac1{N_J}\hat{J}_z^{\,\text{out}}
789: +\frac1{N_L}\hat{L}_z^{\text{out}}-
790: \frac{2}{n\chi\bar{N}}\hat{S}_y^{\,\text{out}}\Big).
791: \ee
792: Note that in this definition $\hat{J}_z^{\,\text{out}}$ and
793: $\hat{L}_z^{\text{out}}$ are divided by the atom numbers of
794: the respective ensembles as before, while the QND measurement
795: yields an estimate of $\hat{J}_z^{\,\text{out}}+\hat{L}_z^{\text{out}}$
796: without such a correction, \cf~Eq.~(\ref{eqn:Syjs}).
797: %
798: %Note that the
799: %operator \mbox{$4\hat{S}_y^{\text{out}}/(n\chi(N_J+N_L))$} is
800: %\emph{not}, as before, exactly the estimation of
801: %\mbox{$\hat{J}_z^{\text{out}}/N_J+\hat{L}_z^{\text{out}}/N_L$}, as the former
802: %does not contain the weights $N_{J/L}^{-1}$.
803: %
804: As a consequence we expect that we loose the advantages of squeezing if the
805: numbers of atoms in the two ensembles differ strongly, {\it i.e.}, if
806: $\gamma\gg1$. We find in leading order \bea
807: \mean{\hat{\phi}_{\rm js}}&=&\phi
808: -\frac{2\alpha}{\bar{N}}\\
809: (\Delta\hat{\phi}_{\rm js})^2&=&\frac{1}{n\chi^2\bar{N}^2}
810: +\frac{2\alpha}{\bar{N}}+
811: \frac{\gamma^2}{8\bar{N}^2}.
812: \label{eqn:jsvar}
813: \eea
814: The importance of the number fluctuations at the preparation stage
815: is reflected in the fact that in order to arrive at these equations,
816: the assumption
817: \be
818: \gamma\sqrt{\bar{N}}(n\chi^2)^2(\theta^2+\phi^2)\ll 1
819: \label{eqn:cond2}
820: \ee
821: is necessary in addition to the assumptions leading
822: to Eq.~(\ref{eqn:varsepsq}). Furthermore, now $1/(n\chi^2\bar{N}^2)$
823: is the leading term only if $n\chi^2\gamma^2/8\ll1$.
824:
825: A QND measurement could also be used after the interferometer
826: to directly read out the joint observable $\hat{J}^{\text{out}}_z+\hat{L}_z^{\text{out}}$
827: by defining
828: \be
829: \hat{\phi}_{\rm js+}=-\frac{2}{n\chi}\frac{1}{\bar{N}}
830: \left(\hat{S}_{y,r}^{\text{out}}-\hat{S}_{y}^{\text{out}}\right),
831: \ee
832: where again the indices $r$ refers to the read-out QND measurement.
833: %labels the Stokes vector $\hat{\vecb{S}}$
834: %distinguish the light ensembles used for squeezing and for read-out, respectively.
835: Notice that in this way it is not possible to measure the
836: correct rescaled observable \mbox{$\hat{J}_z^{\text{out}}/N_J+\hat{L}_z^{\text{out}}/N_L$},
837: and consequently there is an important contribution from the difference
838: of the atom numbers in the expectation value already:
839: \be
840: \mean{\hat{\phi}_{\rm js+}}=\phi+
841: %\frac{\gamma}{\sqrt{\bar{N}}}\frac{\theta}{2}+
842: \frac{N_L-N_J}{N_L+N_J}\theta+
843: \frac{\alpha\phi}{2\bar{N}}.\label{eqn:phijsexp}
844: \ee
845: Compared to the variance of the scheme using jointly squeezed ensembles
846: without the QND read-out measurement, Eq.~(\ref{eqn:jsvar}),
847: the dependence of the variance on the number measurements and on the atom
848: number difference in the two ensembles is reduced for small $\theta,\,\phi$:
849: \bea
850: (\Delta\hat{\phi}_{\rm js+})^2&=&\frac{2}{n\chi^2\bar{N}^2}
851: +\frac{\alpha}{2\bar{N}}(\theta^2+\phi^2)+\nonumber\\
852: & & \,\,+\frac{\gamma^2}{8\bar{N}^2}\alpha(\theta^2+\phi^2).
853: \eea
854: However, the $\gamma$-dependent correction to the
855: expectation value $\mean{\hat{\phi}_{\rm js+}}$ is only negligible
856: compared to the standard deviation
857: $\Delta\phi_{\rm js+}$ if $\gamma^2\bar{N}n\chi^2\theta^2/8\ll1$. This is
858: generally a stronger criterion than the limit on $\gamma$
859: encountered without the QND read-out, \cf Eq.~(\ref{eqn:cond2}).
860:
861: The offset can be compensated by using an estimate for $\theta$ from the
862: final fluorescence measurement
863: \be
864: \hat{\theta}=-\frac1{N_J}\hat{J}_z^{\text{out}}+\frac1{N_L}\hat{L}_z^{\text{out}},
865: \ee
866: to define a corrected phase operator
867: \be
868: \hat{\phi}^c_{js+}=
869: \hat{\phi}_{js+}-\frac{N_L-N_J}{N_L+N_J}\hat{\theta}
870: \ee
871: which takes into account the bias of $\mean{\hat{\phi}_{\rm js+}}.$
872: We find that to leading order
873: \bea
874: \mean{\hat{\phi}^c_{\rm js+}} &=& \phi +\frac{\alpha\phi}{2\bar{N}}-
875: \frac{\gamma^2\alpha}{2\bar{N}^2}\\
876: \left(\Delta\hat{\phi}^c_{\rm js+}\right)^2 &=&
877: \frac{2}{n\chi^2\bar{N}^2}+\frac{\alpha\phi^2}{2\bar{N}}+\frac{\gamma^2}{8\bar{N}^2}.
878: \eea
879: Hence, the $\gamma$-dependent bias in the expectation value is reduced
880: by a factor $\alpha\gamma/(\bar{N}^{3/2}\theta)\ll1$, while
881: the $1/\bar{N}$ term in the variance still has a factor $\alpha\phi^2$.
882:
883: These are the same advantages that we also found in the case of separately
884: squeezed ensembles with QND read-out, cf.~Eq.~(\ref{eqs:varSSQ}). However,
885: the contribution of the term proportional to $1/n\chi^2\bar{N}^2$
886: is reduced by a factor of $2$ in $(\Delta\hat{\phi}^c_{\rm js+})^2$,
887: because only two instead of four projective measurements of the
888: light are necessary in this case.
889:
890: %apart from a factor $2$ by which the variance
891: %$\Delta\hat{\phi}_{\rm jsC}^2$ is reduced in the first term,
892: %because only two light-atom interactions are necessary here as
893: %compared to four in the case of separately squeezed ensembles.
894:
895: %This problem would obviously be avoided, if $\theta$ would be known,
896: %as then $\mean{\hat{\phi}}$ could be corrected because $\gamma$ and $\bar{N}$ are
897: %determined through the final measurement.
898: %For this reason we will know try to identify a scheme
899: %which allows to extract $\theta$ and $\phi$ and performs still better than
900: %just squeezing both ensembles separately.
901:
902: \section{V. Decoherence}
903:
904: The attainable squeezing is limited by the absorption
905: of photons during the interaction between light and atoms \cite{squeezingGaussian}.
906: Each atom which absorbs and subsequently spontaneously emits a photon
907: is no longer correlated to the rest of the atoms, but still adds to the
908: variance. We estimate the number of atoms contributing to such an uncorrelated
909: background as the number of scattered photons $n\kappa$,
910: where $\kappa=N_J\chi\Gamma/\Delta$ is the optical density.
911: %, and $\Gamma$ and $\Delta$ are the line width of the
912: %transition and the detuning from resonance, respectively.
913: Then, in the limit of $\kappa\ll 1$ and for just a single
914: ensemble, one finds \cite{squeezingGaussian}
915: \bea
916: \mean{J_z^{\,\rm out}}&\rightarrow&\mean{J_z^{\,\rm out}}\left(1-\frac{n\chi\Gamma}{\Delta}\right),\label{eqn:dec_ew}\\
917: \mean{\left(J_z^{\,\rm out}\right)^2}&\rightarrow&\left(1-\frac{n\chi\Gamma}{\Delta}\right)^2\mean{\left(J_z^{\,\rm out}\right)^2}+\frac{n N_J\chi\Gamma}{2\Delta}
918: \eea
919: for the collective atomic spin vector and
920: \bea
921: \mean{S_z^{\,\rm out}}&\rightarrow&\mean{S_z^{\,\rm out}}\left(1-\frac{N_J\chi\Gamma}{\Delta}\right),\\
922: \mean{\left(S_z^{\,\rm out}\right)^2}&\rightarrow&\left(1-\frac{N_J\chi\Gamma}{\Delta}\right)^2\mean{\left(S_z^{\,\rm out}\right)^2}+
923: \frac{N_J^2\chi\Gamma}{4\Delta}
924: \eea
925: for the Stokes vector.
926: The leading order corrections to the variance $(\Delta\hat{\phi}_{\rm ss})^2$
927: given in Eq.~(\ref{eqn:varsepsq}) for the case of separately squeezed
928: ensembles reads
929: \bea
930: (\Delta\hat{\phi}_{\rm ss})^2&\rightarrow&
931: (\Delta\hat{\phi}_{\rm ss})^2+\frac{n\chi\Gamma}{\Delta}\frac1{\bar{N}}+
932: \frac{2\Gamma}{n^2\chi\Delta},\label{eq:dec}
933: \eea
934: and similarly for the other schemes. We will
935: use this estimate in the following discussion and leave an in-depth analysis
936: of decoherence processes, e.g., following the lines of \cite{squeezingGaussian},
937: to further investigations. Generally, according to Eq. (\ref{eqn:dec_ew})
938: also the expectation value changes due to decoherence. This can be
939: accounted for by rescaling, and doing so gives only
940: higher-order corrections to Eq.~(\ref{eq:dec}).
941:
942: For usual choices of parameters, the last term
943: in Eq. (\ref{eq:dec}) is negligible, while the contribution
944: proportional to $\bar{N}^{-1}$ is comparable in size to the terms
945: in Eq.~(\ref{eqn:varsepsq}). This limits the coupling $\chi$ and thus
946: the achievable squeezing. For all other experimental parameters
947: fixed, there exists an optimal choice for the detuning $\Delta$
948: and thus for $\chi$ [see Eq. (\ref{eq:chi})]
949: which minimizes the variance. Taking into account only the first term in
950: Eq.~(\ref{eqn:varsepsq}), and in the limit $\Delta\gg\Gamma$,
951: this optimal choice of $\Delta$ leads to a minimal value for
952: $(\Delta\hat{\phi}_{\rm ss})^2$:
953: \be
954: \min_{\Delta}\left[(\Delta\phi_{\rm ss})^2\right]=
955: \frac{2}{N^{\frac32}\,d}\sqrt{\frac{2\epsilon_0\hbar c A \Gamma}{\omega}}.
956: \label{eqn:minvardec}
957: \ee
958: The scaling is thus no longer as in the Heisenberg limit,
959: but it is still better than in shot-noise limited measurements
960: (see also \cite{auzinsh:2004}).
961:
962: In addition, the decay of the states $\ket{1}$ and $\ket{2}$
963: during the interferometer step has to be taken into account.
964: Choosing long-lived hyperfine
965: ground-state levels to implement $\ket{1}$ and $\ket{2}$ minimizes this decay. Also,
966: spin squeezed states have been shown to be robust with respect to both, particle
967: loss and dephasing \cite{stockton:2003}, in contrast to, e.g., GHZ states,
968: which are maximally fragile under particle losses.
969: %We will
970: %leave an in-depth analysis of decoherence in this step to further investigations.
971:
972: %Thus $\mean{J_z}\rightarrow\mean{J_z}(1-\eta)$, and in the limit of
973: %the probability $\eta$ to absorb a photon being small,
974: %$\mean{J_z^2}\rightarrow(1-\eta)^2\mean{J_z^2}+\eta N_J/2$ \cite{squeezingGaussian}.
975: %As each absorbed photon does no longer contributes to the collective Stokes vector, also
976: %$\mean{S_z^2}\rightarrow(1-\kappa)^2\mean{S_z^2}+\kappa N_J/4$ with the optical
977: %density $\kappa$. Estimating $\eta N_J\backsimeq n\kappa$, and
978: %using $\kappa=\chi N_J\Gamma/\Delta$, where $\Gamma$ and $\Delta$ are the
979: %line width and the the detuning from resonance, respectively, we find
980:
981:
982:
983:
984: %=============================================================
985: \section{VI. Comparison of the schemes}
986:
987: \begin{figure*}[pt]
988: %\includegraphics[width=1\linewidth]{scaling_incl_dec4}
989: \includegraphics[width=1\linewidth]{optimizedchi}
990: \caption{
991: Double-logarithmic plot of the variances
992: $(\Delta\phi)^2$ for the methods discussed in the text normalized
993: to the variance for coherent states $(\Delta\phi_{\rm cs})^2$,
994: as a function of the mean number of atoms $\bar{N}$. The scale on the right
995: hand site gives the noise reduction in dB.
996: In both figures $n=10^{11}$, and we fixed $\theta=\phi=0.01$
997: in order to operate close to the point of maximal sensitivity of
998: the interferometer. In (a) we consider the close-to-ideal scenario with
999: $\alpha=2\times10^{-7}$, $\gamma=10$, and decoherence is not included.
1000: $\chi=3.23\times10^{-10}$ is used, corresponding
1001: to a detuning $\Delta=2.28\times10^{10}\;$s$^{-1}$ for the Rb D$_2$ line
1002: and a cross section $A=0.3\;$mm$^2$ of the laser beam.
1003: In this close-to-ideal scenario, the scaling as $\bar{N}^{-2}$ is visible for all
1004: the methods employing squeezing. The graphs for separately squeezed
1005: ensembles (without QND read-out) and for jointly squeezed ensembles
1006: with QND read-out and corrected expectation value lie on top of each other.
1007: In the case of a joint QND read-out,
1008: failing to correct the expectation value for the contribution from $\theta$ results in a
1009: scaling as $1/\bar{N}$ for $\bar{N}\gtrsim10^9$.
1010: (b) Realistic scenario with $\alpha=2\times10^{-2}$, $\gamma=10^4$,
1011: and including decoherence. In each case and for each value of $\bar{N}$
1012: the detuning has been adjusted to
1013: minimize the variance \cite{parameters}. The inset shows the
1014: corresponding optimal values of the coupling parameter $\chi$.
1015: For large atom numbers it is clearly seen that the schemes
1016: which do not employ a final QND measurement are strongly
1017: affected by the limitations from the fluorescence detection. The curve for
1018: jointly squeezed ensembles using a joint QND readout lies outside the range of
1019: the figure.
1020: %
1021: %in all the schemes the variance
1022: %scales as $\bar{N}^{-1}$ due to decoherence, but the schemes
1023: %employing a final QND measurement (including, if necessary, a
1024: %correction of the expectation value), are less effected
1025: %by the limits of the fluorescence detection.
1026: }
1027: \label{fig:scaling_all}
1028: \end{figure*}
1029: %
1030:
1031: To analyze the performance of the schemes discussed in the
1032: preceding sections we will fix the number of photons as
1033: $n=10^{11}$ and take $\bar{N}=10^{10}$ as a reasonable
1034: parameter for the mean atom number per ensemble.
1035: We will first consider a close-to-ideal scenario and
1036: assume that the noise from the fluorescence measurements
1037: can be neglected by setting $\alpha=2\times10^{-7}$. Also, we
1038: will set $\gamma=10$, which for $\bar{N}=10^{10}$ atoms corresponds to
1039: $|N_J-N_L|=10^{-4}\bar{N}$, and we will initially not include decoherence.
1040: Fig.~\ref{fig:scaling_all} (a) shows the scaling with $\bar{N}$ of
1041: $(\Delta\phi)^2/(\Delta\phi_{\rm cs})^2$, {\it i.e.}, of the various variances
1042: normalized to the case of
1043: coherent ensembles. For all the methods involving squeezing of
1044: some observable a Heisenberg like scaling is visible.
1045: The offsets of these curves are given by the
1046: numbers of QND measurements performed, {\it i.e.}, by the factor multiplying
1047: the $1/(n\chi^2\bar{N}^2)$ term in the variance of each of the considered
1048: schemes that involve squeezing. For the measurement of
1049: $\phi$ {\em via} reading out $\hat{J}_z+\hat{L}_z$ through
1050: a QND interaction, the term proportional to $\theta$
1051: shifting the expectation value [see Eq.~(\ref{eqn:phijsexp})]
1052: has been included into the variance, in order to
1053: allow for a fair comparison. It is
1054: this term which makes the variance scale only proportional to $1/\bar{N}$
1055: for $\bar{N}\gtrsim10^9$.
1056: Obviously, correcting this contribution of $\theta$ as described
1057: in Section IV avoids this term and maintains
1058: the improvement by a factor of $2$ compared to the
1059: case of squeezing and QND measuring both ensembles separately.
1060: %\begin{figure}[t]
1061: %\includegraphics[width=1\linewidth]{alphagammasmall}
1062: %\caption{Double-logarithmic plot of the variances $(\Delta\phi)^2$
1063: %as a function of the mean number of atoms $\bar{N}$
1064: %for the methods discussed in the text.
1065: %The parameters have been chosen as $n=10^{10}$, $\chi=2\cdot10^{-8}$, $\theta=\phi=0.01$,
1066: %$\alpha=10^{-7}$ and $\gamma=10$. The latter two correspond to
1067: %a close to ideal scenario. The graphs for separately squeezed
1068: %ensembles (without QND read-out) and for jointly squeezed ensembles
1069: %with QND read-out and corrected expectation value lie on top of each other.
1070: %The scaling as $\bar{N}^{-2}$ is visible for all the methods employing squeezing.
1071: %In the case of a joint QND read-out, failing to correct the expectation
1072: %value for $\theta$ results in a scaling as $1/\bar{N}$ for $\bar{N}\gtrsim10^9$.
1073: %}
1074: %\label{fig:alphagammasmall}
1075: %\end{figure
1076: %
1077: %
1078: %
1079: %
1080: %Now let us consider realistic conditions for the ensemble
1081: %preparation process. A parameter $\gamma=100$ corresponds to
1082: %a difference of the number of atoms in the two ensembles of
1083: %$\sim1$\% of the mean number $\bar{N}$ (at $\bar{N}=10^8$),
1084: %which should be achievable experimentally \cite{collisions}.
1085: %The results are plotted in Fig.~\ref{fig:gamma100}.
1086: %The only curve which is affected
1087: %In the other cases, the variance is
1088: %not affected significantly.
1089: %\begin{figure*}[t]
1090: %\includegraphics[width=0.8\linewidth]{graphs/gamma100bw.eps}
1091: %\caption{The variances $(\Delta\phi)^2$ as a function of $\bar{N}$ for
1092: %$\gamma=100$ and all the other parameters
1093: %as in Fig.~\ref{fig:alphagammasmall}.
1094: %\label{fig:gamma100}
1095: %\end{figure*}
1096:
1097: In Fig.~\ref{fig:scaling_all} (b), the relative variances are plotted for
1098: realistic experimental parameters and including decoherence.
1099: $\alpha=2\times10^{-2}$ corresponds to $50$ fluorescence cycles per atom,
1100: and $\gamma=10^4$ is
1101: equivalent to a difference of the number of atoms in the two
1102: ensembles of $10$\% of the mean number $\bar{N}$ at
1103: $\bar{N}=10^{10}$.
1104: With all other parameters fixed, for each value of $\bar{N}$ the interaction
1105: strength $\chi$ is determined by choosing the detuning $\Delta$ from
1106: atomic resonance such that the variance $(\Delta \phi)^2$
1107: is minimized, see Eq.~(\ref{eqn:minvardec}) \cite{parameters}.
1108:
1109: As can be seen from Fig.~\ref{fig:scaling_all} (b), in such a realistic
1110: scenario the noise reduction obtained from
1111: squeezing decreases for all methods. For large atom numbers, in all cases
1112: the variance scales as $1/\bar{N}$ due to decoherence and the noise from the
1113: fluorescence measurements. For all the procedures not involving a QND read-out,
1114: the strong influence of the latter contribution can be observed, though
1115: for $\bar{N}=10^{10}$ atoms the total noise still is reduced by
1116: around $7\;$dB with respect to the limit set by quantum projection noise.
1117: On the other hand, the read-out {\it via} QND measurements allows to reduce the
1118: noise by more than $10\;$dB compared to this limit, while this method does not require an
1119: additional experimental setup as compared to the scheme where QND measurements
1120: are performed only on the incoming atomic ensembles.
1121:
1122: While in the close-to-ideal case
1123: squeezing of a joint observable gives an advantage of a factor of $2$ in the variance
1124: compared to individual squeezing and read-out (corresponding to a $3\;$dB noise
1125: reduction), for an experimental reasonable scenario this advantage reduces to
1126: about $1.5\;$dB. Of course the method of squeezing a joint observable
1127: of both ensembles has a basic advantage since it only needs a single squeezing
1128: operation instead of two, and thus less technical effort is necessary.
1129:
1130: %\begin{figure*}[pt]
1131: %\includegraphics[width=1\linewidth]{optimizedchi}
1132: %\caption{Same as \ref{fig:scaling_all}, but $\Delta$ has
1133: %been adapted for {\it every} $\bar{N}$ to yield the
1134: %minimal variance.
1135: %}
1136: %\label{fig:scaling_all_optimized}
1137: %\end{figure*}
1138:
1139:
1140:
1141: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1142: \section{VII. Entangled ensembles}\label{sec:entangled}
1143:
1144: Julsgaard {\it et al.}~demonstrated experimentally in Ref.~\cite{Polzik}
1145: the generation of macroscopic entanglement between two
1146: atomic ensembles. The scheme to generate such a macroscopically entangled state,
1147: described first in Ref.~\cite{Duan1}, is motivated by
1148: the fact that under the ideal condition of $\gamma=0$
1149: two commuting joint observables can be constructed from $\hat{\vecb{J}}$
1150: and $\hat{\vecb{L}}$:
1151: \begin{equation}
1152: \mean{[\hat{J}_y-\hat{L}_y,\hat{J}_z+\hat{L}_z]}\propto(N_J-N_L)=0,
1153: \label{commute}
1154: \end{equation}
1155: {\it i.e.}, $\hat{J}_y-\hat{L}_y$ can be measured without affecting $\hat{J}_z+\hat{L}_z$,
1156: and vice versa. This can be seen directly from
1157: \be
1158: (\rot_z(\chi \hat{S}_z)\hat{\vecb{J}}^{\text{in}}-\rot_z(\chi \hat{S}_z)\hat{\vecb{L}}^{\text{in}})_y=
1159: \hat{J}_y^{\text{in}}-\hat{L}_y^{\text{in}},
1160: \ee
1161: {\it i.e.}, the first QND interaction leaves the difference of
1162: the $y$ components unaffected.
1163: Thus after squeezing the sum $\hat{J}_z+\hat{L}_z$, also the
1164: difference $\hat{J}_y-\hat{L}_y$ can be squeezed
1165: without loosing the information gained in the first
1166: measurement. To realize this experimentally in the interferometer,
1167: after the first squeezing
1168: interaction $\hat{\bf J}$ is rotated by a classical $\pi/2$ pulse
1169: around the $x$ axis so that $\hat{J}_y\to \hat{J}_z$ while $\hat{\bf L}$ is
1170: rotated by $-\pi/2$ around $x$ giving $-\hat{L}_y\to \hat{L}_z$. Then a second
1171: laser pulse, prepared again as in Eq. (\ref{eqs:Sin}), interacts consecutively
1172: with both ensembles and thus finally carries information
1173: about $\hat{J}_y-\hat{L}_y$. The outgoing state corresponds to a macroscopically
1174: entangled EPR state \cite{Duan1}.
1175: It is now a natural question to ask whether entangled atomic
1176: ensembles are of use in Sagnac atom interferometry.
1177:
1178: For the schemes discussed so far, the collective spin vectors
1179: lie in the $x-z$ plane after the last step of the interferometer.
1180: Therefore always $\mean{\hat{J}_y-\hat{L}_y}=0$,
1181: and an additional operation is necessary in order to
1182: encode phase information in the $y$ components as well.
1183: This can be achieved by rotating both ensemble vectors $\hat{\vecb{J}}$
1184: and $\hat{\vecb{L}}$ by an angle $\varphi$ and $-\varphi$
1185: around the $x$ axis before and after the interferometric phase is applied, respectively.
1186: In this way the plane of rotation of the phase shift is effectively
1187: tilted by $\varphi$ around the $x$ axis.
1188:
1189: The measurement process now consists of
1190: first rotating $\hat{\vecb{J}}$ and $\hat{\vecb{L}}$ by $\pm\pi/2$ and
1191: using another QND interaction to measure the sum of the $z$ components.
1192: This measurement, scaled correctly by a $\varphi$-dependent factor,
1193: reveals $\phi$. To be more explicit, the corresponding operator
1194: for entangled ensembles (EE) reads
1195: \be
1196: \hat{\phi}_{\rm EE}=\frac1{\cos\varphi}\frac2{\bar{N} n\chi^2}
1197: \left(\hat{S}^{\rm out}_{y,r}-\hat{S}^{\rm out}_{y}\right).
1198: \ee
1199: A measurement of the sum of the $y$ components
1200: can be realized after another rotation around $x$ by either
1201: a QND or a projection measurement. In the former case
1202: \be
1203: \hat{\theta}_{\rm EE}=\frac1{\sin\varphi}\frac2{\bar{N} n\chi^2}
1204: \left(\hat{T}_{y,r}^{\rm out}-\hat{T}^{\rm out}_{y}\right).
1205: \ee
1206: These measurements yield both angles:
1207: \begin{eqnarray}
1208: \mean{\hat{\phi}_{\rm EE}} & = & \phi+
1209: %\frac{\gamma}{\sqrt{\bar{N}}}\frac{\theta}2+
1210: \frac{N_L-N_J}{N_L+N_J}\theta+
1211: \frac{\alpha\;\phi}{2\bar{N}}\\
1212: \mean{\hat{\theta}_{\rm EE}} & = & \theta+
1213: %\frac{\gamma}{\sqrt{\bar{N}}}\frac{\phi}2+
1214: \frac{N_L-N_J}{N_L+N_J}\phi+
1215: \frac{\alpha\;\theta}{2\bar{N}},
1216: \end{eqnarray}
1217: where the offsets can be corrected as above.
1218: For parameters as in Section V. the leading terms
1219: of the corresponding variances read (at $\theta=0$ and $\phi=0$)
1220: \begin{eqnarray}
1221: (\Delta\hat{\phi}_{\rm EE})^2&=&\frac1{\cos^2\varphi}\frac{2}{\bar{N}^2 n\chi^2}+
1222: \frac{\gamma^2\;n\chi^2}{8 \bar{N}}.\label{eqn:sagnac:propop_ent}\\
1223: (\Delta\hat{\theta}_{\rm EE})^2&=&\frac1{\sin^2\varphi}\frac{2}{\bar{N}^2 n\chi^2}+
1224: \frac{\gamma^2\;n\chi^2}{8 \bar{N}}.
1225: \label{eqn:sagnac:trestop_ent}
1226: \eea
1227: Changing $\varphi$ allows to trade in
1228: a lower variance of one component for a higher variance of the other. But
1229: these variances are only scaling with $1/\bar{N}^2$ if $\gamma$ is close to zero,
1230: %XXX
1231: otherwise the last term $\propto \gamma^2n\chi^2/\bar{N}$ in Eqs.
1232: (\ref{eqn:sagnac:propop_ent}) and (\ref{eqn:sagnac:trestop_ent}) is dominating
1233: the scaling. However,
1234: $\gamma\approx0$ is an obvious requirement in this case, as
1235: otherwise the commutator does not vanish in Eq. (\ref{commute}), and thus
1236: the two squeezing operations are not compatible.
1237:
1238: %The measurement process now consists of
1239: %first rotating $\hat{\vecb{J}}$ and $\hat{\vecb{L}}$ by $\pm\pi/2$ and
1240: %using another QND interaction to measure the sum of the $z$ components.
1241: %This measurement, scaled correctly by a $\varphi$ dependent factor,
1242: %reveals $\phi$. To be more explicit, the corresponding operator reads
1243: %\be
1244: %\hat{\phi}_{\rm ent}=\frac1{\cos\varphi}\frac2{\bar{N} n\chi^2}
1245: %\left(\hat{S}^{\rm out}_{y,2}-\hat{S}^{\rm out}_{y}\right),
1246: %\ee
1247: %and find
1248: %\begin{eqnarray}
1249: %\mean{\hat{\phi}_{\rm ent}} & = & \phi+\frac{\gamma}{\sqrt{\bar{N}}}\frac{\theta}2+
1250: %\frac{\alpha\;\phi}{2\bar{N}}\\
1251: %(\Delta\hat{\phi}_{\rm ent})^2&=&\frac1{\cos^2\varphi}\frac{2}{\bar{N}^2 n\chi^2}+
1252: %\frac{\alpha\;\theta^2}{2\bar{N}}+\nonumber\\
1253: %&&+\frac{\gamma^2\alpha(\theta^2-\phi^2)}{8n\bar{N}}+
1254: %\frac{\gamma^2\;n\chi^2}{16 \bar{N}\cos^2\varphi}.\label{eqn:sagnac:propop_ent}
1255: %\end{eqnarray}
1256: %A measurement of the sum of the $y$ components
1257: %can be realized after another rotation around $x$ by either
1258: %a QND or a projection measurement. In the former case
1259: %\be
1260: %\hat{\theta}_{\rm ent}=\frac1{\sin\varphi}\frac2{\bar{N} n\chi^2}
1261: %\left(\hat{T}_y^{\rm out}-\hat{T}^{\rm out}_{y,2}\right),
1262: %\ee
1263: %and
1264: %\bea
1265: %\mean{\hat{\theta}} & = & \theta+\frac{\gamma}{\sqrt{\bar{N}}}\frac{\phi}2+
1266: %\frac{\alpha\;\theta}{2\bar{N}}\\
1267: %(\Delta\hat{\theta}_{\rm ent})^2&=&\frac1{\sin^2\varphi}\frac{2}{\bar{N}^2 n\chi^2}+
1268: %\frac{\alpha\;\theta^2}{2\bar{N}}+\nonumber\\
1269: %&&+\frac{\gamma^2\alpha(\theta^2-\phi^2)}{8n\bar{N}}+\frac{\gamma^2\;n\chi^2}{16 \bar{N} \sin^2\varphi}.
1270: %\label{eqn:sagnac:trestop_ent}
1271: %\eea
1272: %From these two measurements thus $\phi$ and $\theta$ can be inferred by correcting
1273: %the expectation values as above. Changing $\varphi$ allows to trade in
1274: %a lower variance of one component for a higher variance of the other. But
1275: %these measurements are only at the Heisenberg limit if $\gamma$ is close to zero,
1276: %otherwise the last term $\propto \gamma^2n\chi^2/\bar{N}^2$ in Eqs.
1277: %(\ref{eqn:sagnac:propop_ent}) and (\ref{eqn:sagnac:trestop_ent}) is dominating
1278: %the scaling. But $\gamma\approx0$ is an obvious requirement in this case, as
1279: %otherwise the commutator does not vanish in Eq. (\ref{commute}), and thus
1280: %the two squeezing operations are not compatible.
1281:
1282:
1283:
1284: %=============================================================
1285: \section{VIII. Conclusion and outlook}
1286:
1287: We have presented and compared in detail several methods to improve the
1288: detection of a differential phase shift of two atomic
1289: interferometers beyond the standard quantum limit, having in mind
1290: especially the application to Sagnac interferometry. For this
1291: purpose, we have analyzed the squeezing of individual and joint
1292: observables and, in both cases, the read-out of the interferometer
1293: {\em via} fluorescence detection of the atoms only or by an additional
1294: QND interaction.
1295:
1296: If decoherence and measurement imperfections
1297: are neglected, all the methods of squeezing improve
1298: the behavior of the variance of
1299: %XXX
1300: the differential phase to a $1/\bar{N}^2$ scaling, modified
1301: by a factor $k/(n\chi^2)\gg1$,
1302: which is determined by the number $k$ of QND interactions
1303: involved, by the number of photons $n$, and by the coupling $\chi$
1304: between atoms and photons. In the case of jointly squeezed
1305: observables, we found that this limit can only be attained if some
1306: constraints on the difference of the number of atoms in both
1307: ensembles can be fulfilled. In all cases, the achievable
1308: squeezing is limited by decoherence due to the absorption
1309: of photons during the QND measurement.
1310:
1311: Using fluorescence measurements to
1312: read out the atomic spins after the interferometer always produces
1313: additional noise scaling as $1/\bar{N}$ due to the photon shot
1314: noise. As an alternative method, a QND measurement can be
1315: employed to read out the final state of the interferometer. Although in
1316: this case fluorescence measurements are still necessary to determine the number
1317: of atoms in the two ensembles, their contribution to
1318: the noise is reduced to a large extent. We have shown that
1319: the best method to achieve this is to perform squeezing and read-out
1320: {\em via} a QND measurement of a joint observable of the two ensembles,
1321: provided that the difference between the number of atoms in the
1322: two ensembles can be made smaller than approximately $10$\% of
1323: the mean number of atoms. This procedure
1324: minimizes the number of QND interactions necessary, thereby minimizing
1325: the factor multiplying the $1/\bar{N}^2$ term in the variance,
1326: and it reduces the experimental effort.
1327:
1328: Finally, we considered the creation of a macroscopically entangled state
1329: of the two atomic ensembles {\it via} squeezing of two non-local, commuting
1330: observables. We showed that in this case both the sum and the difference
1331: of the phase shifts can be measured with a variance scaling with
1332: $1/\bar{N}^2$, and that
1333: the relative uncertainty can be shifted between both quantities. However,
1334: %XXX
1335: this scaling can only be reached here if the number
1336: difference between the two ensembles can be made very small.
1337: %, which
1338: %makes this approach a challenge for future experiments.
1339: %renders this scheme rather impractical in current experimental setups.
1340: Therefore, it would be desirable to identify methods to
1341: control the numbers of atoms in the ensembles, for instance by
1342: employing the superfluid -- Mott insulator transition in an
1343: optical lattice embedded in a weakly confining harmonic potential
1344: \cite{micirac,mibloch}. Controlling
1345: this confinement, which plays the role of a local chemical potential, the number
1346: of atoms in the Mott phase at $T>0$ can be controlled and should only depend
1347: mildly on the total number of atoms in the system.
1348: A detailed analysis of this idea is left for future investigations.
1349: %This idea however needs
1350: %further theoretical investigation.
1351:
1352: \section{VIII. Acknowledgments}
1353: It is a pleasure to thank H.A. Bachor, M. Drewsen, J.~Eschner,
1354: P. Grangier, O.~G{\"u}hne, K.~M{\o}lmer, and A.~Sanpera for
1355: illuminating discussions.
1356: %This work has been supported by the DFG (Schwerpunktprogramme
1357: %'Quanteninformationsverarbeitung' and
1358: %'Wechselwirkung in ultrakalten Atom- und Molekülgasen'
1359: %and SFB 407), by the EC (QUPRODIS, FINAQS, and QUDEDIS).
1360: This work has been supported by the DFG (SPP 1116, SPP 1078, GRK 282, POL 436, CRK, and SFB 407),
1361: by the EC (QUPRODIS, FINAQS), and by the ESF (QUDEDIS).
1362: UVP acknowledges support from the Danish Natural Science Research Council.
1363:
1364:
1365: \begin{thebibliography}{99}
1366:
1367: \bibitem{atomicclocks} R. Bluhm, V.A. Kostelecky, C.D. Lane, and N. Russell,
1368: Phys. Rev. Lett. {\bf 88}, 090801 (2002).
1369:
1370: \bibitem{jentsch} C. Jentsch, T. M{\"u}ller, E. Rasel, and W. Ertmer,
1371: Gen. Rel. Grav. {\bf 36}, 2197 (2004).
1372:
1373: \bibitem{weitz} S. Fray, C. Alvarez-Diez, T.W. H{\"a}nsch, and M. Weitz,
1374: Phys. Rev. Lett. {\bf 93}, 240404 (2004).
1375:
1376: \bibitem{snadden:1998}
1377: M.J.~Snadden, J.M.~McGuirk, P.~Bouyer, K.G.~Haritos, and M.A.~Kasevich,
1378: %"Measurement of the Earth's gravity gradient with an atom interferometer-based gravity gradiometer",
1379: Phys. Rev. Lett. {\bf 81}, 971 (1998).
1380:
1381: \bibitem{kasevich:2002}
1382: J.M.~McGuirk, G.T.~Foster, J.B.~Fixler, M.J.~Snadden, and M.A.~Kasevich,
1383: %title = "Sensitive absolute-gravity gradiometry using atom interferometry",
1384: Phys. Rev. A {\bf 65}, 033608 (2002).
1385:
1386: \bibitem{KasevichChu} M.A. Kasevich and S. Chu, Phys. Rev. Lett.
1387: {\bf 67}, 181 (1991)
1388:
1389: \bibitem{peters} A. Peters, K.Y. Chung, and S. Chu,
1390: Metrologia {\bf 38}, 25 (2001).
1391:
1392: \bibitem{inter} B. Yurke, S.L. McCall, and J.R. Klauder,
1393: Phys. Rev. A {\bf 33}, 4033 (1986).
1394:
1395: \bibitem{heisenberg} P. Kok, S.L. Braunstein, and J.P. Dowling,
1396: Jour. Opt. B {\bf 6}, 811 (2004).
1397:
1398: \bibitem{oblak} D. Oblak, J.K. Mikkelsen, W. Tittel, A.K. Vershovski, J.L. S{\o}rensen,
1399: P.G. Petrov, C.L. Garrido Alzar, and E.S. Polzik, preprint: quant-ph/0312165 (2003).
1400:
1401: \bibitem{QND} A. Kuzmich, N.P. Bigelow, and L. Mandel,
1402: Europhys. Lett. {\bf 42}, 481 (1998).
1403:
1404: \bibitem{absorbnonclassical}
1405: A.~Kuzmich, K.~M{\o}lmer, and E.S.~Polzik,
1406: %title = "Spin Squeezing in an Ensemble of Atoms Illuminated with Squeezed Light",
1407: Phys. Rev. Lett. {\bf 79}, 4782 (1997);
1408: J.~Hald, J.L.~S{\o}rensen, C.~Schori, and E.S.~Polzik,
1409: %title = "Spin Squeezed Atoms: A Macroscopic Entangled Ensemble Created by Light",
1410: Phys. Rev. Lett. {\bf 83}, 1319 (1999).
1411:
1412: %\bibitem{recent} A. Andr{\'e}, A.S. S{\o}rensen, and M.D. Lukin,
1413: % Phys. Rev. Lett. {\bf 92}, 230801 (2004).
1414:
1415:
1416: % A. Kuzmich and T.A.B. Kennedy,
1417: % Phys. Rev. Lett. {\bf 92}, 030407 (2004).
1418:
1419: \bibitem{memory} C. Mewes and M. Fleischhauer, Phys. Rev. A {\bf 66},
1420: 033820 (2002); H. Kaatuzian, A.~Rostami, A.A.~Oskouei, preprint: quant-ph/0402057 (2004);
1421: D.N. Matsukevich and A. Kuzmich, Science {\bf 306}, 663 (2004);
1422: B. Julsgaard, J. Sherson, J.I. Cirac, J. Fiur{\'a}\v{s}ek, and E.S. Polzik,
1423: Nature {\bf 432}, 482 (2005).
1424:
1425: \bibitem{Soerensen} A. S. S{\o}rensen and K. M{\o}lmer, Phys. Rev. Lett. {\bf 86},
1426: 4431 (2001).
1427:
1428: \bibitem{Duan1} L.M. Duan, J.I. Cirac, P. Zoller, and E.S. Polzik,
1429: Phys. Rev. Lett. {\bf 85}, 5643 (2000).
1430:
1431: \bibitem{Polzik} B. Julsgaard, A. Kozhekin, and E. Polzik,
1432: Nature {\bf 413}, 400 (2001).
1433: % ; D. Oblak {\em et al.}, quant-ph/0312165.
1434:
1435: \bibitem{Vivi} V. Petersen, L.B. Madsen, and K. M{\o}lmer,
1436: Phys. Rev. A {\bf 71}, 012312 (2005).
1437:
1438: \bibitem{Kasevich} T.L. Gustavson, A. Landragin, and M.A. Kasevich,
1439: Class. Quantum Grav. {\bf 17}, 2385 (2000).
1440:
1441: \bibitem{atominterferometry} B. Young, M. Kasevich, and S. Chu,
1442: \emph{Precision Atom Interferometry with Light Pulses}, in:
1443: Atom interferometry, ed. by P.R. Berman, Academic Press 1997.
1444:
1445: \bibitem{numbers} G. Santarelli {\em et al.}, Phys. Rev. Lett.
1446: {\bf 82}, 4619 (1999).
1447:
1448: \bibitem{kuzmichPRL} A. Kuzmich and T. A. B. Kennedy,
1449: Phys. Rev. Lett. {\bf 92}, 030407 (2004).
1450:
1451: \bibitem{squeezingGaussian}
1452: L.B. Madsen and K. M{\o}lmer, Phys. Rev. A {\bf 70}, 052324 (2005).
1453:
1454: %\bibitem{fluorescence} Reference on fluorescence. Todo! Necessary?
1455: % Oder: see \cite{Kasevich,atominterferometry}.
1456:
1457: \bibitem{auzinsh:2004}
1458: M. Auzinsh, D. Budker, D.F. Kimball, S.M. Rochester, J.E. Stalnaker, A.O. Sushkov, and V.V. Yashchuk,
1459: Phys. Rev. Lett. {\bf 93}, 173002 (2004).
1460:
1461: \bibitem{stockton:2003}
1462: J.K.~Stockton, J.M.~Geremia, A.C.~Doherty, H.~Mabuchi,
1463: %Characterizing the entanglement of symmetric many-particle spin-(1/2) systems
1464: Phys. Rev. A {\bf 67}, 022112 (2003).
1465:
1466: \bibitem{parameters}
1467: For the parameters in the figure, the optimal values for the detuning
1468: $\Delta$ and the interaction parameter $\chi$ are (at $\bar{N}=10^{10}$ atoms
1469: for the Rb D$_2$ line and a laser beam cross section of $0.3\;$mm$^2$):
1470: separately squeezed ensembles $\Delta=2.45\times10^{10}$s$^{-1}$, $\chi=3.01\times10^{-10}$;
1471: separately squeezed ensembles (QND readout) $\Delta=2.07\times10^{10}$s$^{-1}$, $\chi=3.56\times10^{-10}$;
1472: joint squeezing $\Delta=3.01\times10^{10}$s$^{-1}$, $\chi=2.44\times10^{-10}$;
1473: joint squeezing (QND readout and correction) $\Delta=2.53\times10^{10}$s$^{-1}$, $\chi=2.91\times10^{-10}$.
1474:
1475:
1476: \bibitem{micirac} D. Jaksch, C. Bruder, J.I. Cirac, C. W. Gardiner, and P. Zoller,
1477: Phys.~Rev.~Lett.~{\bf 81}, 3108 (1998).
1478:
1479: \bibitem{mibloch} M.~Greiner, O.~Mandel, T.~Esslinger, T.W. H{\"a}nsch, and I.~Bloch,
1480: Nature {\bf 415}, 39 (2002).
1481:
1482: \end{thebibliography}
1483:
1484: \end{document}
1485:
1486: %%%%%%%%%%%%%% FINITO %%%%%%%%%%%%%%%%%%%