1: %
2: % Choose preprinttrue for something readable.
3: % Set it false for two columns.
4: %
5:
6: %\documentstyle[prl,aps,epsfig,preprint]{revtex}
7: \documentstyle[prl,aps,epsfig,twocolumn]{revtex}
8:
9: \draft
10:
11: \newlength{\textwidthm}
12: \setlength{\textwidthm}{\columnwidth}
13: \addtolength{\textwidthm}{-\parindent}
14: \addtolength{\textwidthm}{-\parindent}
15: \begin{document}
16: \title{Atom correlations and spin squeezing near the Heisenberg limit:
17: finite size effect and decoherence}
18: \author{A. Andr\'e and M.D. Lukin}
19: \address{Physics Department, Harvard University and \\
20: ITAMP, Harvard-Smithsonian Center for Astrophysics, Cambridge, MA~~02138}
21: \date{\today}
22: \maketitle
23: \begin{abstract}
24:
25: We analyze a model for spin squeezing based on the so-called counter-twisting
26: Hamiltonian, including the effects of dissipation and finite system size.
27: We discuss the conditions under which the Heisenberg limit, i.e. phase
28: sensitivity $\propto 1/N$, can be achieved.
29: A specific implementation of this model based on atom-atom
30: interactions via quantized photon exchange is presented in detail.
31: The resulting
32: excitation corresponds to the creation of spin-flipped atomic pairs
33: and can be used for
34: fast generation of entangled atomic ensembles, spin squeezing and
35: applications in quantum information processing.
36: The conditions for achieving strong spin squeezing with this mechanism are
37: also analyzed.
38:
39: \end{abstract}
40:
41: \pacs{PACS numbers 03.67.-a, 42.50.-p, 42.50.Gy}
42:
43: %\nobreak
44: %\begin{twocolumn}
45:
46: \section{Introduction}
47:
48: Interacting quantum systems that start in uncorrelated states generally
49: evolve towards entangled states due to quantum correlations building up in
50: time. These correlations and the form they take depend crucially on the
51: interaction that gives rise to them. For example in parametric down
52: conversion or in the optical parametric oscillator (OPO) pairs of photons
53: can be created in distinct modes of the electromagnetic field. The fact
54: that {\it pairs} of photons are generated leads to quantum correlations
55: between the two modes. Since each mode is described by a harmonic
56: oscillator, one can think of the state of the field as the quantum state
57: of two fictitious particles in harmonic oscillator potentials. The
58: quantum correlations correspond to e.g. the positions of the particles
59: being strongly correlated, in the ideal case $\Delta
60: (X_1-X_2)^2\rightarrow 0$ and their
61: momenta being anticorrelated $\Delta (P_1+P_2)^2\rightarrow 0$. For the
62: electromagnetic field modes, the position and momenta correspond to
63: quadratures of the field modes and it is between these that correlations
64: are produced \cite{Kimble-Houches,Walls-Milburn}.
65: These correlations
66: are essential to quantum communication e.g. quantum teleportation of
67: information from one location to another \cite{Q-teleport}.
68: Entanglement is also crucial for many schemes in quantum cryptography and
69: for long-distance quantum communication through lossy channels
70: \cite{Longdist}.
71:
72: Since the mechanism for producing correlations in electromagnetic field
73: modes is at the fundamental level so simple (photons created in pairs) it
74: is natural to wonder if such a simple mechanism may lead to entanglement
75: of atoms interacting in a similar manner. In complete analogy to the OPA
76: mechanism, a process that transfers {\it pairs} of atoms from their ground
77: state to two well defined final states also gives rise to quantum
78: correlations between atoms. When a collection of $N$ two level atoms is
79: thought of as an ensemble of effective spin ${\tiny \frac{1}{2}}$
80: particles with total pseudo-angular momentum $J=N/2$, it turns out
81: \cite{Ueda} that the quantum correlations produced by an interaction that
82: transfers atoms in pairs from the lower state to the upper state shows up
83: as reduced fluctuations in a component of the angular momentum e.g.
84: $\Delta J_x^2 \rightarrow 0$.
85: We will discuss entanglement of atoms with
86: one another in an atomic ensemble for which an effective interaction
87: leads to the transfer of atoms in pairs to well defined final states and
88: we will use the concept of spin squeezing to quantify the amount of
89: quantum correlations produced in such a case. As for squeezed states of
90: light, decoherence mechanisms and dissipation are acting in such a way as
91: to destroy or limit the amount of squeezing achievable in practice. We
92: also analyze the influence of such dissipation mechanisms and
93: find relations between the spin squeezing interaction rate, the
94: dissipation rate and the amount of squeezing achievable in the presence of
95: damping mechanisms.
96: The coherent control of the dynamical evolution of complex systems such as
97: atomic ensembles may lead to the
98: production of entangled non-classical states such as spin squeezed states
99: \cite{Ueda} (analogous to squeezed states of light \cite{Walls-Milburn})
100: and correlated collective atomic modes (similar to twin
101: photons generated by a non-degenerate OPO).
102:
103: The main result of this paper is that for a collection of $N$ atoms with
104: average single atom nonlinearity $\chi$ (two atom interaction energy) and
105: with single atom loss rate $\Gamma$, the condition for achieving some
106: spin squeezing is that $N\chi > \Gamma$.
107: In order to achieve reduction of uncertainty in say $J_x$ compared to the
108: uncertainty in the Bloch state $|J=N/2,J_z=N/2\rangle$ for which $(\Delta
109: J_x)^2=N/4$ by an amount $s$ (i.e. $(\Delta J_x)^2=N/(4s)$)
110: with $1\leq s \leq N$, one requires that $N\chi>s\Gamma$ and the
111: interaction time needed scales as $t\sim (\log{s})/(N\chi)$ while the
112: maximum number of atoms than can be lost without destroying the squeezing
113: scales as $\Delta N\sim (N/s)\log{s}$.
114: To achieve Heisenberg limited precision (i.e. maximum
115: spin squeezing $s\sim N$), one needs a large single atom nonlinearity
116: $\chi>\Gamma$. This means that the interaction time needed to achieve this
117: strongly correlated state is $t\sim (\log{N})/(N\chi)$ and the maximum
118: number of atoms that can be
119: lost without compromising this optimal squeezing is $\Delta N\sim
120: \log{N}$ i.e. a very small number of atoms lost may prevent reaching the
121: Heisenberg limit. This analysis remains valid and agrees with a specific
122: implementation
123: based on an effective atom-atom interaction via quantized photon
124: exchange in a cavity, for which the decoherence mechanism corresponds to
125: spontaneous emission and leakage of photons from the cavity.
126:
127: The possibility of coherently controlling interacting quantum systems
128: has lead to many new developments in the field of quantum information
129: science \cite{Bouwmeester}.
130: These are expected to have an impact in a broad area ranging from quantum
131: computation and quantum communication \cite{Nielsen-Chuang} to precision
132: measurements \cite{Wineland} and controlled modeling of complex quantum
133: phenomena \cite{Sachdev}.
134: Entangled systems realized in the laboratory range in size from few qbits
135: \cite{Wineland2}, to macroscopic ensemble of particles \cite{Polzik}.
136: Controllable coherent interactions between atoms \cite{Lloyd,Meystre} may
137: also open the
138: way for modelling of complex quantum phenomena such as quantum phase
139: transitions \cite{Sachdev} in which quantum correlations play a crucial role.
140:
141: Entanglement of a single atomic ensemble, i.e. quantum correlations between
142: atoms in the same ensemble, has been shown to be potentially very
143: useful in the field of precision measurements \cite{Wineland}.
144: Certain types of interactions between atoms lead to
145: entanglement and spin squeezing,
146: characterized by reduced variance in an
147: observable and increased fluctuations in the canonically conjugate observable.
148: This reduction of fluctuations directly translates into an improved
149: accuracy for
150: measurements sensitive to that observable. A typical figure of merit for spin
151: squeezed states is the phase accuracy $\delta\phi$ on
152: estimating accumulated dynamical phase in the Ramsey interferometric
153: experiment. With all experimental uncertainties controlled below this noise
154: level, the dominant source of noise in such experiments is the ``quantum
155: projection noise'' \cite{Wineland} associated with e.g. the noise in
156: measurements of the
157: x-component of the spin of an ensemble of two-level atoms (effective spin
158: ${\tiny \frac{1}{2}}$) all
159: prepared in the lower level (the $|\downarrow\rangle$ state). This noise leads
160: to a lower limit on phase accuracy $\delta\phi=1/\sqrt{N}$ called the
161: standard quantum limit (SQL), where $N$ is
162: the number of atoms in the ensemble.
163: The Heisenberg uncertainty principle
164: however allows for phase accuracies consistent with the basic principles of
165: quantum mechanics that are as low as $\delta\phi=1/N$, called the
166: Heisenberg limit.
167:
168: We also discuss in more detail a technique \cite{Paper1} based on a
169: resonantly enhanced
170: nonlinear process involving Raman scattering into a
171: ``slow'' optical mode \cite{slow-light}, which creates a
172: pair of spin-flipped atom and slowly propagating coupled excitation
173: of light and matter (dark-state polariton). When the group
174: velocity of the polariton is reduced to zero \cite{stop-light,darkpolar},
175: this results in pairs of spin flipped atoms.
176: The dark-state polariton can be easily converted into corresponding states
177: of photon wavepackets ``on demand'' \cite{darkpolar},
178: which makes the present approach most suitable for implementing protocols
179: in quantum information processing that require a combination of
180: deterministic sources of entangled states and
181: long-lived quantum memory \cite{Longdist,Q-mem}.
182:
183:
184:
185: This paper is divided into V sections. In Section II, we discuss Ramsey
186: spectroscopy and the use of spin-squeezed states in precision
187: measurements.
188: In particular we analyze the situation where $N$ two level atoms with levels
189: $|g\rangle$ and $|e\rangle$ are prepared in a correlated state and
190: subsequently
191: probed by separated fields of frequency $\omega$ in the Ramsey
192: interferometric configuration, which we review in Appendix A.
193: We also discuss spin-squeezed states and develop pictorial
194: representation of those states which we compare to squeezed states of
195: light and in appendix B we introduce the Wigner representation for a
196: particular class of spin squezed states.
197:
198: In Section III, we analyze a model for spin squeezing based on the analogy
199: with
200: the optical parametric oscillator. We also seek to understand the influence of
201: loss processes on the coherent spin-squeezing interaction and the way in which
202: it limits the correlations achievable for a given interaction rate. The
203: model
204: consists of two bosonic modes (a ``spin up'' state and a ``spin down'' state)
205: with loss
206: rates and a coherent interaction that transfers pairs of atoms from one mode
207: to the other.
208: For our simple model, analytical results can be obtained
209: in the perturbative regime of small number of excitations (most atoms in the
210: lower state) and low loss rate. We estimate the
211: conditions for which Heisenberg-limited spin squeezed states can be
212: produced.
213:
214:
215: In Section IV, we present a scheme for inducing effective coherent
216: interactions
217: between atoms in an atomic ensemble. These coherent interactions
218: lead to massive entanglement of the ensemble and to characterize the degree
219: of entanglement thus obtained, we calculate the squeezing or reduction in
220: fluctuations of one particular observable. The coherent interaction is based
221: on Raman scattering into a cavity mode for which the atomic medium is made
222: transparent by Electromagnetically Induced Transparency (EIT). The slowly
223: propagating mode is then best described by a polariton: a collective
224: excitation that is partly photonic and partly ``spin'' excitation of the atomic
225: ensemble (the up and down states of the spin being two metastable states).
226: The overall process leads to the creation of pairs of
227: excitations, one being a ``spin flip'' created by Raman scattering, the other
228: being a polariton which can be ``steered'' into a photon or spin flip
229: excitation ``on demand''. We find that substantial spin squeezing can be
230: obtained for atomic ensembles in low finesse cavities, without the
231: strong coupling requirement
232: of cavity QED. In the limit of unity finesse this corresponds to free-space
233: configuration and substantial correlations can still be produced in this
234: case. In the
235: opposite limit of high finesse, very strong correlations are obtained and
236: in particular we estimate the regime for which Heisenberg limited spin-squeezed
237: states are produced.
238:
239:
240: \section{Ramsey spectroscopy with correlated atoms}
241:
242: In appendix A, Ramsey spectroscopy is reviewed and in particular
243: we show how
244: the phase accuracy in phase estimation based on the Ramsey fringe signal is,
245: at the maximum sensitivity point, given by
246:
247: \begin{equation}
248: \delta\phi=\frac{\Delta J_x}{|\langle \hat{J}_z \rangle|}
249: \end{equation}
250:
251: where $\Delta J_x$ is the variance in the x-component of the pseudo angular
252: momentum (of length $J=N/2$)
253: representing the state of $N$ two-level atoms and $\langle \hat{J}_z \rangle$
254: is the expectation value of the z-component of the pseudo angular momentum
255: (both the variance and the expectation value are
256: calculated in the initial state).
257:
258: For an uncorrelated state of atoms e.g. with all atoms in their lower
259: state so that the state of the ensemble is described by
260: $|J_z=-N/2\rangle$, it is found that
261: $\Delta J_x=\sqrt{J/2}$ and $\langle \hat{J}_z \rangle=-J$ so that
262: $\delta\phi=\sqrt{1/N}$.
263: In order to improve the phase accuracy, one must use
264: a state for which the variance in
265: $\hat{J}_x$ is reduced while $\langle \hat{J}_z \rangle$ is little
266: changed.
267: Consider therefore a state such as an
268: eigenstate of $\hat{J}_x$, for example $|J_x=0\rangle$. Calculating the
269: expectation values and variances we find $\langle
270: \hat{J}_x \rangle =
271: \langle \hat{J}_y \rangle = \langle \hat{J}_z \rangle = 0$, $\Delta J_x = 0$
272: and $\Delta J_y = \Delta J_z = \sqrt{J(J+1)/2}$. However the Ramsey signal
273: has amplitude proportional to $\langle J_z \rangle$ and therefore vanishes
274: for all phase angles $\phi$, which means that even though the noise or
275: fluctuation properties of the signal may be improved, its average is zero.
276: Note that this is
277: because we have chosen $\hat{J}_z(\phi)$ as our observable, other
278: observables such as $\hat{J}_z^2(\phi)$ for example may lead to non-zero
279: average signal together with reduced variance
280: \cite{Kasevich,Holland}. However, it turns out their signal to noise ratio
281: is very much reduced compared to that of the Ramsey scheme \cite{Holland}.
282: It is thus necessary to consider
283: states that lead to a reduced variance $\Delta J_x$ while maintaining a large
284: signal amplitude, i.e. a large $\langle \hat{J}_z \rangle$. We therefore
285: consider states such as
286:
287: \begin{equation}
288: |\psi(a)\rangle = \frac{1}{\sqrt{1+a^2}}(i|J_x=0\rangle +
289: a\frac{|J_x=+1\rangle - |J_x=-1\rangle}{\sqrt{2}})
290: \label{psia}
291: \end{equation}
292:
293: where $a$ is a real number parametrizing the state $|\psi(a)\rangle$.
294: It is straightforward to calculate the expectation values
295: \begin{eqnarray}
296: \langle \hat{J}_x \rangle &=& 0 \nonumber \\
297: \langle \hat{J}_y \rangle &=& 0 \nonumber \\
298: \langle \hat{J}_z \rangle &=& \frac{2a}{1+a^2} \sqrt{\frac{J(J+1)}{2}}
299: \end{eqnarray}
300:
301: and the variances
302: \begin{eqnarray}
303: \Delta J_x &=& \frac{a}{\sqrt{1+a^2}} \nonumber \\
304: \Delta J_y &=& \frac{1}{\sqrt{1+a^2}}\sqrt{\frac{J(J+1)}{2}} \nonumber \\
305: \Delta J_z &=& \sqrt{\frac{J(J+1)}{2}}\left[1-\frac{4a^2}{(1+a^2)^2}\right]
306: ^{1/2} .
307: \label{varpsia}
308: \end{eqnarray}
309:
310: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
311:
312: \begin{figure}[ht]
313: \begin{center}
314: \epsfig{file=fig1.eps,width=8.5cm}
315: \leavevmode
316: \end{center}
317: \vspace*{2ex}
318: \caption{Number of atoms detected in the upper state ($N_e$) relative
319: to total number of atoms (the total number is $N=N_e+N_g$
320: and thus $N_e/(N_e+N_g)=(N/2+\langle J_z(\phi) \rangle )/N$)
321: vs. accumulated phase
322: $\phi=(\omega-\omega_0)T$ for a) uncorrelated atoms and b) correlated atoms in
323: a spin-squeezed state $|\psi(a)\rangle$, for $N=100$ and $a=-1.1$ (note
324: that error bars have been magnified by a factor of $10$ for clarity).
325: Note how squeezing of the variance improves the phase accuracy.
326: }
327: \end{figure}
328:
329: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
330:
331: The signal amplitude which depends on $\langle \hat{J}_z \rangle$ can thus
332: be rather large ($O[N]$) while the noise amplitude characterized by
333: $\Delta J_x$ is minimized ($O[1]$).
334: The Ramsey signal and phase accuracy for such a state is shown in Fig. 1b,
335: compared
336: to the case of uncorrelated atoms (Fig. 1a).
337:
338: These states are minimum uncertainty states i.e.
339: $(\Delta J_x)(\Delta J_y)={\tiny \frac{1}{2}}|\langle \hat{J}_z \rangle |$
340: for all values of the parameter $a$. Also, their phase accuracy is given by
341:
342: \begin{equation}
343: \delta\phi(\pm\pi/2) = \sqrt{\frac{1+a^2}{2}}\frac{1}{\sqrt{J(J+1)}}
344: \end{equation}
345:
346: which is of order $1/N$.
347: The best phase accuracy is obtained for $a\rightarrow 0$ in which case the
348: optimal phase accuracy is $\delta\phi(\pm\pi/2)=\sqrt{2}/N$. Note that in
349: this case the signal amplitude ($\propto \langle\hat{J}_z\rangle$) becomes
350: vanishingly small ($\langle\hat{J}_z(\phi)\rangle\rightarrow 0$ for all
351: $\phi$) and also the
352: range of values of $\phi$ for which improved phase accuracy is achieved
353: becomes vanishingly small around $\phi=\pm\pi/2$. For these reasons, the
354: optimally spin squeezed state $|\psi(a=0)\rangle$ may prove impractical.
355: Note however that for finite $a$ i.e. for $|a|=1$, the signal amplitude is
356: large ($\sim N/\sqrt{8}$) and
357: the phase accuracy is independent of $\phi$
358:
359: \begin{equation}
360: \delta\phi(\phi)=\frac{1}{\sqrt{J(J+1)}}\simeq\frac{2}{N}
361: \end{equation}
362:
363: i.e. twice the Heisenberg limit.
364:
365: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
366:
367: \begin{figure}[ht]
368: \begin{center}
369: \epsfig{file=fig2.eps,width=8.5cm}
370: \leavevmode
371: \end{center}
372: \vspace*{2ex}
373: \caption{Number of atoms detected in the upper state vs. accumulated phase
374: $\phi=(\omega-\omega_0)T$ for correlated atoms in various spin-squeezed states
375: $|\psi(a)\rangle$, for $N=100$ and $a=-0.9$, $a=-1$ and $a=-1.1$.
376: Also shown is the phase accuracy $\delta\phi(\phi)$ vs. accumulated phase
377: (the dashed line represents the standard quantum limit
378: $\delta\phi=1/\sqrt{N}$).
379: Note how $\delta\phi(\phi)$ gets to a minimum value of order $2/N=0.02$.
380: }
381: \end{figure}
382:
383: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
384:
385: In Fig. 2 we show the signal and variance for various spin squeezed states
386: along with the phase accuracy $\delta\phi(\phi)$.
387:
388: We can gain a better understanding of the squeezing in the states
389: (\ref{psia})
390: $|\psi(a)\rangle$ by looking at various representations of them.
391: The simplest representation is to project the state onto eigenstates of the
392: three components of the angular momentum
393:
394: \begin{equation}
395: P_i(m) = |\langle J_i=m|\psi(a)\rangle |^2
396: \end{equation}
397:
398: where $|J_i=m\rangle$ is the eigenstate of the i-component of angular momentum
399: with eigenvalue $m$.
400:
401: From Fig. 3 it is clear that the expectation value of $\hat{J}_x$ and
402: $\hat{J}_y$ are zero in such a state, whereas (for $a=-1$) the expectation
403: value of $\hat{J}_z$ is large and negative, the variances are clearly given by
404: (\ref{varpsia}). It is interesting to note the similarity of these angular
405: momentum squeezed states and those of a harmonic oscillator (i.e.
406: squeezed states of light). In both cases, the probability distributions vanish
407: for odd number of quanta (note that for simplicity we consider only $N$ even
408: here).
409:
410: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
411:
412: \begin{figure}[ht]
413: \begin{center}
414: \epsfig{file=fig3.eps,width=8.5cm}
415: \leavevmode
416: \end{center}
417: \vspace*{2ex}
418: \caption{Projection of the state $|\psi(a)\rangle$ onto eigenstates of the
419: angular momentum operators $\hat{J}_x$, $\hat{J}_y$ and $\hat{J}_z$ for
420: $J=N/2=20$ and $a=-1$. The $P_x(m)$ distribution is sharply peaked since
421: the
422: state $|\psi(a)\rangle$ is a superposition of the $m_x=-1,0,+1$ components
423: only;
424: the $P_y(m)$ distribution is broad and symmetric; the $P_z(m)$ distribution
425: vanishes for $m$ odd and the even components decrease roughly exponentially
426: with $m$.}
427: \end{figure}
428:
429: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
430:
431: For even number of quanta, the behaviour is
432: nearly exponential $P_z(m)\propto e^{-c(m+j)}$ for some constant $c$.
433: A mechanism for generating such states starting from the uncorrelated state
434: $|J_z=-J\rangle$ must therefore be one in which atoms are excited in pairs,
435: i.e.
436: two atoms in the ground state are transferred to the excited state
437: $|g\rangle |g\rangle \rightarrow |e\rangle|e\rangle$.
438: Consider the similarities with squeezed states of light: in particular the
439: photon number distribution vanishes
440: for odd photon number in the case of squeezed vacuum due to the form of the
441: squeezing Hamiltonian $\hat{H}=-i\chi [\hat{a}^{\dagger 2}-\hat{a}^2]$, which
442: creates and destroys photons in pairs. Since we find a similar cancellation
443: of the probability of there being odd number of excitations for the states
444: $|\psi(a)\rangle$, the interaction giving rise to such states starting from
445: all atoms in their lower states must likewise create and destroy excitations
446: in pairs and thus be of the form
447: $\hat{H}=-i\hbar\chi[\hat{L}_+^2-\hat{L}_-^2]$.
448: This process can
449: also be viewed as a coherent collision mechanism. Moreover, for the whole
450: atomic ensemble to become entangled (not just particular atom pairs), this
451: process must occur completely symmetrically for all atoms. It should
452: not be two particular atoms that get transferred to the excited state,
453: rather it should be two collective excitations that get created
454:
455: \begin{eqnarray}
456: |g_1\cdots g_N\rangle & \rightarrow & \sqrt{\frac{2}{N(N-1)}}\sum_{i>j}\left[
457: |g\cdots e_i \cdots e_j \cdots g\rangle\right]
458: \end{eqnarray}
459:
460: which is the state obtained by letting $\hat{J}_+^2$ operate on
461: $|J_z=-J\rangle$. The simplest Hamiltonian giving rise to this type of
462: interaction is analyzed in section III to quantify the squeezing generated by
463: this mechanism.
464: This form of interaction was considered by Kitegawa and Ueda \cite{Ueda} in
465: their classic study of spin squeezing and was dubbed the ``two axis
466: countertwisting'' interaction.
467:
468: \section{Two axis counter twisting model}
469:
470: We now turn to the analysis of the two axis countertwisting
471: Hamiltonian \cite{Ueda}
472:
473: \begin{eqnarray}
474: \hat{H} &=& -i\frac{\hbar\chi}{2}\left( \hat{L}_+^2-\hat{L}_-^2 \right)
475: \nonumber \\
476: &=& \hbar\chi \left(\hat{L}_x\hat{L}_y+\hat{L}_y\hat{L}_x\right).
477: \label{ctham}
478: \end{eqnarray}
479:
480: As argued in last section it is this type of Hamiltonian that most
481: closely parallels squeezed state generation for light.
482:
483: We now present a general theory that allows to quantify atom correlations
484: and takes into account decoherence and finite system size.
485: Specifically, we consider two bosonic modes (such as for a two
486: component Bose-Einstein condensate or for an atomic ensemble with two relevant
487: atomic levels) with annihilation
488: operators $\hat{a}_1$
489: and $\hat{a}_2$. The system is also subject to damping i.e. loss of atoms at
490: rates which may depend on the internal state.
491: The equations of motion for the two modes are
492: then (with $\hat{L}_+=\hat{a}_2^\dagger\hat{a}_1$ and
493: $\hat{L}_-=\hat{a}_1^\dagger\hat{a}_2$)
494:
495: \begin{eqnarray}
496: \dot{\hat{a}}_1 = -\gamma_1 \hat{a}_1 + \chi\hat{a}_1^\dagger\hat{a}_2^2 +
497: \hat{F}_1(t) \nonumber \\
498: \dot{\hat{a}}_2 = -\gamma_2 \hat{a}_2 - \chi\hat{a}_2^\dagger\hat{a}_1^2 +
499: \hat{F}_2(t)
500: \end{eqnarray}
501:
502: where $\hat{F}_j(t)$ are delta-correlated Langevin noise forces with
503: appropriate diffusion coefficients
504: $D_{ij}=\langle \hat{F}_i(t)\hat{F}_j(t) \rangle$.
505:
506:
507: In order to discuss spin squeezing, it is easier to rewrite the equations of
508: motion in terms of the Stokes parameters
509:
510: \begin{eqnarray}
511: \hat{L}_0 &=& \hat{N} = \hat{a}_1^\dagger\hat{a}_1+\hat{a}_2^\dagger\hat{a}_2 \nonumber \\
512: \hat{L}_x &=& (\hat{a}_2^\dagger\hat{a}_1+\hat{a}_1^\dagger\hat{a}_2)/2 \nonumber \\
513: \hat{L}_y &=& (\hat{a}_2^\dagger\hat{a}_1-\hat{a}_1^\dagger\hat{a}_2)/2i \nonumber \\
514: \hat{L}_z &=& (\hat{a}_2^\dagger\hat{a}_2-\hat{a}_1^\dagger\hat{a}_1)/2
515: \end{eqnarray}
516: for which the equations are
517:
518: \begin{eqnarray}
519: \dot{\hat{L}}_0 &=& -2\Gamma\hat{L}_0+4\gamma\hat{L}_z+\hat{F}_0(t)
520: \nonumber \\
521: \dot{\hat{L}}_x &=& -2\Gamma\hat{L}_x+\chi(\hat{L}_x\hat{L}_z+
522: \hat{L}_z\hat{L}_x)+\hat{F}_x(t) \nonumber \\
523: \dot{\hat{L}}_y &=& -2\Gamma\hat{L}_y-\chi(\hat{L}_y\hat{L}_z+
524: \hat{L}_z\hat{L}_y)+\hat{F}_y(t) \nonumber \\
525: \dot{\hat{L}}_z &=& -2\Gamma\hat{L}_z+\gamma\hat{L}_0-2\chi(\hat{L}_x^2-
526: \hat{L}_y^2)+\hat{F}_z(t)
527: \label{fullqmeqns}
528: \end{eqnarray}
529:
530: where $\Gamma=(\gamma_1+\gamma_2)/2$, $\gamma=(\gamma_1-\gamma_2)/2$ and
531: $\hat{F}_j(t)$ are new delta-correlated noise forces associated with the
532: damping.
533:
534: Since (\ref{fullqmeqns}) are non-linear operator equations,
535: in the equations of motion for the first order moments
536: $\langle \hat{L}_i \rangle$ there are terms that depend on those first order
537: moments but also terms depending on the second order moments
538: $\langle \hat{L}_i\hat{L}_j \rangle$. Similarly the equations of
539: motion for the second order
540: moments depend on themselves and also on the third order moments, and so
541: on, leading to
542: the BBGKY hierarchy of equations of motion for the expectation values of
543: operator products.
544: In order to solve this set of equations, the hierarchy must be truncated
545: at some order \cite{Vardi}. Keeping the first and second order
546: moments,
547: we truncate the BBGKY hierarchy by the approximation
548:
549: \begin{eqnarray}
550: \langle\hat{L}_i\hat{L}_j\hat{L}_k\rangle & \approx &
551: \langle\hat{L}_i\hat{L}_j\rangle\langle\hat{L}_k\rangle +
552: \langle\hat{L}_j\hat{L}_k\rangle\langle\hat{L}_i\rangle +
553: \langle\hat{L}_k\hat{L}_i\rangle\langle\hat{L}_j\rangle \nonumber \\
554: & - & 2\langle\hat{L}_i\rangle\langle\hat{L}_j\rangle\langle\hat{L}_k\rangle .
555: \end{eqnarray}
556:
557: The equations of motion for the expectation values $l_i\equiv \langle \hat{L}_i
558: \rangle$ and the second order moments $\Delta_{ij}\equiv \langle
559: \hat{L}_i\hat{L}_j + \hat{L}_j\hat{L}_i \rangle - 2\langle\hat{L}_i\rangle
560: \langle\hat{L}_j\rangle$
561: are then obtained from (\ref{fullqmeqns}). We are interested in the case
562: when all atoms start in mode $1$, i.e.
563: $l_0(0)=N,l_x(0)=l_y(0)=0,l_z(0)=-N/2$ and
564: $\Delta_{xx}(0)=\Delta_{yy}(0)=N/2$ (all other second moments vanish) and
565: for simplicity we take $\gamma_1=\gamma_2=\Gamma$, $\gamma=0$.
566: Writing only the relevant equations and omitting vanishing terms (such as those
567: proportional to $\Delta_{xz}$ and $\Delta_{yz}$ which are zero for all times),
568: we have (after some algebra)
569:
570: \begin{eqnarray}
571: \dot{l}_0 &=& -2\Gamma l_0 \nonumber \\
572: \dot{l}_z &=& -2\Gamma l_z -\chi(\Delta_{xx}-\Delta_{yy}) \nonumber \\
573: \dot{\Delta}_{xx} &=& -4\Gamma\Delta_{xx}+\Gamma l_0 +4\chi l_z\Delta_{xx}
574: \nonumber \\
575: \dot{\Delta}_{yy} &=& -4\Gamma\Delta_{yy}+\Gamma l_0 -4\chi l_z\Delta_{yy}
576: \label{bbgky1}
577: \end{eqnarray}
578:
579: and $l_x(t)=l_y(t)=0$. These equations are non-linear and cannot be solved
580: analytically nor perturbatively in $\Gamma/\chi$. For short enough times,
581: the number of excitations into mode $2$ is small and $l_z\simeq -N/2$, so
582: that plugging this in (\ref{bbgky1}) we have
583:
584: \begin{eqnarray}
585: \Delta_{xx}(t) &\simeq& {\tiny \frac{N}{2}}e^{-2N\chi t}+O[\Gamma/\chi]
586: \nonumber \\
587: \Delta_{yy}(t) &\simeq& {\tiny \frac{N}{2}}e^{2N\chi t}+O[\Gamma/\chi]
588: \end{eqnarray}
589:
590: i.e. the variance of the x-component of the pseudo-angular momentum is
591: squeezed while
592: that of the y-component is anti-squeezed. Plugging these back in the equation
593: of motion of $l_z$, we obtain $l_z(t)\simeq -N/2+(\cosh{2N\chi t}-1)/2+
594: O[\Gamma/\chi]$. This equation predicts growth of $l_z$ without bound, however
595: we know that because $l_z$ is the z-component of an angular momentum vector,
596: we must have $|l_z|\leq N/2$. The phase space of this angular momentum vector
597: is the Bloch sphere and in essence we have neglected the small
598: curvature of the Bloch sphere (of radius $R=N/2$) and have approximated
599: the
600: phase space
601: by the flat planar phase space of a harmonic oscillator. We call this
602: approximation the bosonic approximation, since it predicts infinite
603: squeezing in the long time limit and in the absence of dissipation,
604: similar to the case of squeezed light.
605: Formally this is equivalent to assuming
606:
607: \begin{eqnarray}
608: [\hat{a}_2^\dagger \hat{a}_1,\hat{a}_1^\dagger\hat{a}_2] &=&
609: \hat{a}_2^\dagger\hat{a}_2-\hat{a}_1^\dagger\hat{a}_1 \nonumber \\
610: &\simeq& -N
611: \end{eqnarray}
612:
613: i.e. the operator $\hat{S}_+=\hat{a}_2^\dagger \hat{a}_1/\sqrt{N}$ obeys
614: bosonic commutation relations. Under this approximation the Hamiltonian
615: (\ref{ctham}) becomes $\hat{H}=-i(\hbar\chi N/2)(\hat{S}_+^2-\hat{S}_-^2)$
616: which
617: is identical to the Hamiltonian describing squeezing of light
618: \cite{Walls-Milburn}.
619:
620:
621: In order to take into account the curvature of phase space and the non-bosonic
622: nature of the angular momentum operators, we use the following
623: transformation
624:
625: \begin{eqnarray}
626: \hat{N} &=& N \hat{h}_0 \nonumber \\
627: \hat{L}_x &=& \sqrt{N} \hat{h}_x \nonumber \\
628: \hat{L}_y &=& \sqrt{N} \hat{h}_y \nonumber \\
629: \hat{L}_z &=& \hat{h}_z-\frac{N}{2}\hat{h}_0
630: \label{groupcon}
631: \end{eqnarray}
632:
633: in terms of which the commutation relations become
634:
635: \begin{eqnarray}
636: \left[ \hat{h}_{x,y,z},\hat{h}_0 \right] & = & 0 \nonumber \\
637: \left[ \hat{h}_z,\hat{h}_{\pm} \right] & = & \pm \hat{h}_{\pm} \nonumber \\
638: \left[ \hat{h}_+,\hat{h}_- \right] & = & 2\frac{\hat{h}_z}{N}-\hat{h}_0
639: \end{eqnarray}
640:
641: where $\hat{h}_\pm=\hat{h}_x\pm i\hat{h}_y$. In the limit $N\rightarrow \infty$
642: these commutation relations become those of bosonic operators i.e.
643: $\lim_{N\rightarrow\infty}[\hat{h}_0,\hat{h}_z,\hat{h}_+,\hat{h}_-]=[\hat{1},
644: \hat{a}^\dagger\hat{a},\hat{a}^\dagger,\hat{a}]$, a process formally known as a
645: group contraction \cite{Courtens}. The linear transformation of operators
646: (\ref{groupcon}) does not introduce any extra approximation.
647:
648: The Hamiltonian (\ref{ctham}) can be re-expressed as
649:
650: \begin{eqnarray}
651: \hat{H} &=& \hbar\chi N\left(\hat{h}_x\hat{h}_y+\hat{h}_y\hat{h}_x\right)
652: \nonumber \\
653: &=& -i\frac{\hbar\xi}{2}\left(\hat{h}_+^2-\hat{h}_-^2\right)
654: \end{eqnarray}
655:
656: where we have defined $\xi=\chi N$.
657: We can now obtain equations of motion for the expectation
658: values $h_j=\langle \hat{h}_j \rangle$ and the second order moments
659: $\delta_{ij}=\langle \hat{h}_i\hat{h}_j \rangle -2 \langle \hat{h}_i \rangle
660: \langle \hat{h}_j \rangle$ from (\ref{bbgky1}). Letting $\tau=N\chi
661: t=\xi t$ be a rescaled time, $\kappa=\Gamma/(N\chi)=\Gamma/\xi$ be
662: the rescaled dissipation rate and writing $\epsilon=1/N$ and $\dot{x}={\rm
663: d}x/{\rm d\tau}$,
664: we have
665:
666: \begin{eqnarray}
667: \dot{h}_0 &=& -2\kappa h_0 \nonumber \\
668: \dot{h}_z &=& -2\kappa h_z -(\delta_{xx}-\delta_{yy}) \nonumber \\
669: \dot{\delta}_{xx} &=& -4\kappa\delta_{xx}+\kappa h_0-2h_0\delta_{xx}+4\epsilon
670: h_z\delta_{xx} \nonumber \\
671: \dot{\delta}_{yy} &=& -4\kappa\delta_{yy}+\kappa h_0+2h_0\delta_{yy}-4\epsilon
672: h_z\delta_{yy}.
673: \label{bbgky2}
674: \end{eqnarray}
675:
676: Note that these equations are formally equivalent to (\ref{bbgky1}), no
677: approximation has been made from (\ref{bbgky1}) to (\ref{bbgky2}). Letting
678: $\epsilon\rightarrow 0$ reproduces the results of the bosonic
679: approximation obtained
680: above in the limit of $l_z\simeq -N/2$. Terms of order $\epsilon$ and higher
681: represent corrections to the bosonic approximation and, as shown below,
682: they give rise to a limit to the amount of squeezing achievable.
683:
684: Solving (\ref{bbgky2}) to first order in $\epsilon$ and $\kappa$ we
685: obtain, writing only the relevant terms,
686:
687: \begin{eqnarray}
688: \delta_{xx}(\tau) &=& \frac{1}{2}\left[e^{-2\tau}+(\kappa+\epsilon/2)+\cdots
689: \right]
690: \label{squbbgky}
691: \end{eqnarray}
692:
693: which shows that the variance $\Delta J_x=\sqrt{(N/2)\delta_{xx}}$ is squeezed.
694: Second order terms in $\kappa$ and $\epsilon$ come multiplied by an
695: exponentially
696: growing term $e^{2\tau}$ so that as a function of time, the variance reaches
697: a minimum value $\delta_{xx}\sim {\rm max}[\kappa,\epsilon]$ at a time
698: $e^{-\tau_*}\sim {\rm max}[\kappa,\epsilon]$, after which it grows
699: exponentially and the squeezing is lost. Note that this behaviour
700: ($\delta_{xx}(t)$ reaches a minimum value and then increases again) also
701: occurs when $\kappa\rightarrow 0$, indicating that it is a generic feature
702: of the finite system size.
703: This model predicts that
704: a variance $\delta_{xx}\sim \epsilon=1/N \rightarrow \Delta J_x^2\sim 1$
705: is achievable as long as losses are small enough, i.e.
706: $\kappa\sim\epsilon$,
707: which in terms of $\chi$ and $\Gamma$ means
708:
709: \begin{equation}
710: \chi \sim \Gamma\;\;\;{\rm or}\;\;\xi \sim N\Gamma
711: \label{heislimit}
712: \end{equation}
713:
714: where $\chi$ corresponds to the single-atom nonlinear interaction rate and
715: $\Gamma$ represents the single-atom loss rate.
716:
717: In order to achieve any squeezing ($\delta_{xx}\leq 1/2$) it is
718: necessary to have $\kappa<1$ i.e. $N\chi > \Gamma$.
719: In the regime $N\chi \gg \Gamma$ very
720: strong correlations can be obtained. Note that the single-atom
721: nonlinearity can still be relatively weak compared to the single particle
722: loss rate ($\chi\ll\Gamma$).
723: For example when the dissipation rate is such that
724: $\kappa\sim\sqrt{\epsilon}$ i.e. $\sqrt{N}\chi\sim\Gamma$, the amount of
725: squeezing obtained (\ref{squbbgky}) is $\delta_{xx}\sim 1/\sqrt{N}$. It
726: takes a time $e^{-2\tau}\sim\sqrt{\epsilon}$ to reach this state and the
727: number of particles lost during that time is $\Delta N\sim N\times
728: 2\kappa\tau\sim\sqrt{N}\log{N}$.
729: This number can therefore also be thought of
730: as the maximum number of particles that can be lost from the ensemble
731: without destroying squeezing beyond $\delta_{xx}\sim 1/\sqrt{N}$.
732:
733: In order to reach the Heisenberg limit it is required that the
734: single atom nonlinearity $\chi$ be larger than the decay rate $\Gamma$.
735: Note that in this case, the number of atoms lost by the time optimal
736: squeezing is achieved is $\Delta N\sim\log{N}$
737: which indicates that a very small number of atoms is lost. This
738: number also corresponds to the maximum number of particles that can be
739: lost from the ensemble and not destroy squeezing at the Heisenberg limit
740: level. Clearly the more squeezed the state of the atoms is, the more
741: sensitive it becomes to atom loss and in general to any form of
742: dissipation.
743:
744:
745: \section{Coherent atom interactions via slow light}
746:
747: We now describe a technique to induce effective coherent
748: interactions between atoms in metastable states \cite{Paper1}.
749: The technique is based on a resonantly enhanced
750: nonlinear process involving Raman scattering into a
751: ``slow'' optical mode \cite{slow-light}, which creates a
752: pair of spin-flipped atom and slowly propagating coupled excitation
753: of light and matter (dark-state polariton). When the group
754: velocity of the polariton is reduced to zero \cite{stop-light,darkpolar},
755: this results in pairs of spin flipped atoms.
756: The fact that pairs of atomic excitations are created in this
757: process can also be viewed as a coherent interaction between atoms, i.e. a
758: controlled ``collision'' leading to entanglement of the state of each atom
759: with that of every other atom in the ensemble.
760:
761: A number of proposals
762: have been made for generating entangled states of atomic ensembles
763: and resulting in so-called spin squeezed states. Some are based
764: on interatomic interactions at ultra-cold temperatures \cite{squ-cold}, whereas
765: others involve mapping the states of non-classical light fields into
766: atoms \cite{squ-Polzik}, QND measurements of spins \cite{squ-Bigelow} with
767: light or dipole blockade for Rydberg atoms \cite{Bouchoule}.
768: In contrast to some of these mechanisms the present approach does not require
769: coherence of the atomic motion or sources of non-classical light
770: and is completely deterministic thereby
771: significantly simplifying possible experimental realizations.
772: We further show that the present technique can be made robust with
773: respect to realistic decoherence processes such as spontaneous emission
774: and leakage of slow photons from the medium.
775:
776: We consider a system of $N$ atoms (Fig. 4a) interacting with two
777: classical driving fields $\Omega_{1,2}$ and
778: one quantized mode $\hat{a}$ of a running wave cavity that is initially in a
779: vacuum state. Note that we consider a cavity configuration for ease of
780: theorerical treatment, the results of this analysis however remain valid in the
781: limit of unity finesse, i.e. in free space configuration.
782: Relevant atomic sublevels
783: include two manifolds of metastable states (e.g hyperfine sublevels of
784: electronic ground state) and excited states that may be accessed by optical
785: transitions. The atoms are initially prepared in their ground states
786: $|g\rangle$. One of the classical fields, of Rabi frequency $\Omega_1$,
787: is detuned from the atomic resonance by an amount roughly equal to the
788: frequency splitting between ground state manifolds.
789: The other field of Rabi frequency $\Omega_2$ is resonant with an atomic
790: transition $|b_2\rangle\rightarrow|a_2\rangle$.
791: The quantized field can be involved in two Raman transitions
792: corresponding to Stokes and anti-Stokes processes. Whereas the former
793: corresponds to the usual Stokes scattering in the forward direction, the
794: latter establishes an Electromagnetically Induced Transparency (EIT) and
795: its group velocity is therefore substantially reduced.
796:
797: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
798:
799: \begin{figure}[ht]
800: \begin{center}
801: \epsfig{file=fig4.eps,width=8.5cm}
802: \leavevmode
803: \end{center}
804: \vspace*{2ex}
805: \caption{
806: Energy levels scheme for the effective coherent interaction leading to creation
807: of pairs of atoms a) in different final states (``non-degenerate'' scheme) and
808: b) in identical final states(``degenerate'' version).}
809: \end{figure}
810:
811: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
812:
813: The pair excitation can be viewed as resulting from quantized photon
814: exchange between atoms (Fig. 5) in a two-step process. The first flipped
815: spin is created due to Stokes Raman scattering, which also results in
816: photon emission in a corresponding Stokes mode. In the presence of EIT,
817: this photon is directly converted into a dark-state polariton which becomes
818: purely atomic when the group velocity is reduced to zero.
819: This implies that atomic spins are always flipped in pairs.
820: In Fig. 4a the two final states involved in Raman
821: transitions are different and atomic pairs in different states
822: are created. In Fig. 4b the final states of the two Raman processes are
823: identical, in which case atomic pairs in the same state result. The analysis
824: of this ``degenerate'' version of the scheme is similar to the
825: ``non-degenerate'' case and we will consider only the latter case
826: here.
827:
828: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
829:
830: \begin{figure}[ht]
831: \begin{center}
832: \epsfig{file=fig5.eps,width=6.0cm}
833: \leavevmode
834: \end{center}
835: \vspace*{2ex}
836: \caption{Diagram illustrating coherent atom-atom interaction mediated by
837: dark-state polariton, leading to the creation of a pair of spin-flipped atoms.}
838: \end{figure}
839:
840: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
841:
842: For conceptual simplicity we assume that the quantized field
843: corresponds to a single mode of a running-wave cavity with a creation
844: operator $\hat{a}^\dagger$ and atom-field coupling constants $g_1$ and
845: $g_2$.
846: The interaction
847: Hamiltonian for the system of N atoms and light can be split into two
848: parts $H = H_{Ram} + H_{res}$ corresponding to the Stokes Raman process and the
849: anti-Stokes process respectively:
850:
851: \begin{eqnarray}
852: H_{Ram}&=&-\hbar\Delta \Sigma_{a1a1} - \hbar\delta_1 \Sigma_{b_1b_1}
853: \nonumber \\
854: &+&[\hbar\Omega_1 \Sigma_{a_1g}
855: + \hbar g_1 a \Sigma_{a_1b_1} + {\rm
856: h.c.}], \label{hamram} \\
857: H_{res}&=&\hbar\delta_2 \Sigma_{b_2b_2} + \hbar\delta_2 \Sigma_{a_2a_2}
858: \nonumber \\
859: &+& [\hbar g_2 a \Sigma_{a_2g} + \hbar\Omega_2 \Sigma_{a_2b_2} + {\rm h.c.}], \label{hamres}
860: \end{eqnarray}
861:
862: where $\Sigma_{\mu\nu} = \sum_i
863: |\mu\rangle_{ii}\langle \nu|$ are collective atomic operators corresponding to
864: transitions between atomic states $|\mu\rangle,|\nu\rangle$, $\Delta$ is the
865: detuning of the classical field $\Omega_1$ from the single-photon transition
866: $|g\rangle \rightarrow |a_1\rangle$, $\delta_1$ and $\delta_2$ are the
867: two-photon detunings from the $|g\rangle \rightarrow |b_1\rangle$ and
868: $|g\rangle \rightarrow |b_2\rangle$ transitions respectively as shown in
869: Fig. 4.
870:
871: In the limit of large detuning $\Delta$ and ignoring two-photon detunings for
872: the moment, the Hamiltonian $H_{Ram}$ describes
873: off-resonant Raman scattering.
874: We take into account realistic decoherence
875: mechanisms such as spontaneous emission from the
876: excited states in all directions and decay of the cavity mode
877: with a rate $\kappa$. The evolution of atomic operators is then described
878: by Heisenberg-Langevin equations:
879: \begin{eqnarray}
880: {\dot \Sigma}_{\mu\nu} = -\gamma_{\mu\nu} \Sigma_{\mu\nu} + {i \over \hbar}
881: [H, \Sigma_{\mu\nu}] + F_{\mu\nu},
882: \end{eqnarray}
883: where $\gamma_{\mu\nu}$ is a decay rate of coherence $\mu\rightarrow \nu$
884: and $F_{\mu\nu}$ are associated noise forces. The latter have zero average
885: and are $\delta$-correlated with associated diffusion coefficients that
886: can be found using the Einstein relations.
887:
888: After a canonical transformation corresponding to adiabatic elimination of
889: the excited state (see Appendix C for details), $H_{Ram}$ becomes equivalent
890: to the effective Hamiltonian
891:
892: \begin{equation}
893: {\tilde H}_{Ram} = \hbar\chi\hat{a}^\dagger\hat{S}_1^\dagger
894: + {\rm h.c.}
895: \label{hamrameff}
896: \end{equation}
897:
898: where $\hat{S}_1=\Sigma_{gb_1}/\sqrt{N}$ and
899: $\chi=g_1\sqrt{N}\Omega_1^*/\Delta$.
900: This effective hamiltonian thus describes the process in which a Stokes
901: photon is emitted necessarily accompanied by a spin flip. The quantum
902: state of the Stokes mode is thus perfectly correlated with the state of
903: the atomic spin flip mode.
904:
905: The resonant part of
906: the Hamiltonian $H_{res}$ is best analyzed in terms of dark and bright-state
907: polaritons \cite{brightpolar}
908: \begin{eqnarray}
909: P_D &=& {\Omega_2 a - g_2 \sqrt{N} S_2 \over \sqrt {g_2^2 N + \Omega_2^2}}, \nonumber \\
910: P_B &=& {g_2 \sqrt{N} a + \Omega_2 S_2 \over \sqrt {g_2^2 N + \Omega_2^2}},
911: \label{polaritons}
912: \end{eqnarray}
913: which are superpositions of photonic and atomic excitations $\hat{a}$ and
914: $S_2=\Sigma_{gb_2}/\sqrt{N}$..
915: In particular, $H_{res}$
916: has an important family of dark-states:
917: \begin{eqnarray}
918: |D^n\rangle \sim (P_D^\dagger)^n |g\rangle|{\rm vac}\rangle
919: \label{dark}
920: \end{eqnarray}
921: with zero eigenenergies. This means that once in the dark state, the system
922: stays in the dark state. Note that all other eigenstates of $H_{res}$
923: have, in general, non-vanishing interaction energy.
924: Under conditions of Raman resonance and sufficiently slow excitation
925: (``adiabatic condition'', see Appendix D for details)
926: the Stokes photons emitted by Raman scattering, Eq.(\ref{hamrameff}), will
927: therefore couple solely to the dark-states (\ref{dark}). In this case
928: the coherent part of the evolution of the entire system is described by an
929: effective
930: Hamiltonian:
931: \begin{eqnarray}
932: H_{eff} = -i \hbar \xi (P_D^\dagger S_1^\dagger - S_1 P_D),
933: \label{eff}
934: \end{eqnarray}
935: with $\xi = \Omega_1 \Omega_2/ \Delta \times g_1 \sqrt{N}/
936: \sqrt{g_2^2 N + \Omega_2^2}$ (without loss of generality, $\xi$ was chosen
937: imaginary here for simplified calculations).
938: The Hamiltonian (\ref{eff}) describes the coherent
939: process of generation of pairs of excitations involving polaritons and
940: spin-flipped atoms.
941: Note that for small number of excitations the spin waves and
942: polaritons obey bosonic commutation relations and this Hamiltonian is formally
943: equivalent to that describing optical parametric amplification (OPA) of two
944: modes \cite{Walls-Milburn}.
945: In the non-bosonic limit, this Hamiltonian is also analogous to the
946: ``counter-twisting'' model of (\ref{ctham}).
947: In appendix D we show that the coupled equations for the polariton $P_D$
948: and the spin flip $S_1$ are given by
949:
950: \begin{eqnarray}
951: \dot{S}_1^\dagger &=& (\frac{|g_1|^2}{|g_2|^2}\gamma_L-\gamma_L-i\delta_1)
952: S_1^\dagger + \xi P_D +{\tilde F}_{S_1}^\dagger(t)
953: \label{finspin}
954: \\
955: \dot{P}_D &=& -(\kappa/\eta+\gamma_L+i\delta_2)P_D +\xi S_1^\dagger +
956: {\tilde F}_D(t)
957: \label{finpol}
958: \end{eqnarray}
959:
960: where the polariton decay rate includes an atomic part $\gamma_L$ and a
961: photonic part $\kappa/\eta$ due to leakage of photons out of the medium
962: (at a rate reduced by the factor $\eta=|g_2|^2N/|\Omega_2|^2$ equal to
963: the ratio of vacuum light velocity to the group
964: velocity of slowly propagating Stokes photons).
965: The spin flip operator
966: equation (\ref{finspin}) is seen to
967: contain both a decay term and a gain term due to spontaneous emission into
968: the bright polariton mode. Note that this apparent decrease in dissipation
969: is however accompanied by increased fluctuations denoted by the new noise
970: force operator ${\tilde F}_{S_1}(t)$.
971: The effective detuning between
972: the polariton and spin flip mode is seen to correspond to the difference
973: in two-photon detunings
974: $\delta_1-\delta_2$.
975:
976: We now consider the scenario in which the system is evolving for a time $\tau$
977: under the Hamiltonian $H_{eff}$, after which both fields are turned off.
978: If the procedure is adiabatic upon turn-off of the coupling fields
979: $\Omega_{1,2}$ the polaritons are converted into pure spin excitations
980: $P_D \rightarrow S_2$. Hence the entire procedure will correspond to the
981: following state of the system:
982: \begin{eqnarray}
983: |\Psi\rangle = && { 1\over \cosh\xi\tau} \sum_n (\tanh\xi\tau)^n {1 \over n!}
984: (P_D^\dagger)^n (S_1^\dagger)^n |g\rangle|{\rm vac}\rangle \nonumber \\
985: && \rightarrow
986: { 1\over \cosh\xi\tau} \sum_n (\tanh\xi\tau)^n |n_{b_1},n_{b_2} \rangle|{\rm vac}\rangle.
987: \end{eqnarray}
988: Here $|n_{b_1},n_{b_2} \rangle = 1/n! (S_2^+)^n (S_1^+)^n |g\rangle$ are
989: Dicke-like
990: symmetric states of atomic ensemble and we assumed $n_{b_1,b_2} \ll N$.
991: For non-zero $\xi \tau$ this state describes an entangled state,
992: for which relative fluctuations between the two modes
993: decreases exponentially to values well below the standard quantum limit
994: (SQL) corresponding to uncorrelated atoms.
995:
996: The present technique can also be viewed as a new mechanism for
997: coherent ``collisions'' \cite{Meystre}
998: between atoms mediated by light. In particular, the case when
999: atomic pairs are excited into two different levels
1000: (as e.g. in Fig. 4a) closely resembles coherent spin-changing
1001: interactions that occur in degenerate atomic samples \cite{Ketterle},
1002: whereas the case
1003: when atomic pairs are stimulated into the identical state (Fig. 4b)
1004: is reminiscent of dissociation of a molecular condensate
1005: \cite{Heinzen}.
1006: To put this analogy in perspective we note
1007: that the rate of the present optically induced process can exceed that of
1008: weak interatomic interactions by orders of magnitude. Therefore the
1009: present work may open up interesting new possibilities
1010: for studying many-body phenomena of strongly interacting atoms.
1011:
1012: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1013:
1014: \begin{figure}[ht]
1015: \centerline{\epsfig{file=fig6.eps,width=8.5cm}}
1016: \vspace*{2ex}
1017: \caption{(a) Quadrature variance $\Delta Y_+^2$ vs. single-photon detuning
1018: $\Delta$ and interaction time $\xi t$, (b) same for $\Delta=\Delta_{opt}$ and
1019: $\delta_1=\delta_2$ showing maximum squeezing $\Delta Y_+^2\simeq 0.02$
1020: (for $\sqrt{g^2N/\gamma\kappa}=100$), (c) Number of excitations pumped
1021: in the system vs. time (same conditions as in b) and (d) $\Delta Y_+(t^*)^2$
1022: vs. two-photon detuning ${\bar \delta}\equiv (\delta_1-\delta_2)/2$ for
1023: $\Delta=\Delta_{opt}$ and where $t^*$ gives maximum squeezing.}
1024: \end{figure}
1025:
1026: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1027:
1028: To quantify the resulting correlations established between the polariton mode
1029: $P_D$ and the pure spin flip mode $S_1$, we introduce the quadratures of both
1030: modes (which are bosonic for small number of excitations) in direct analogy to
1031: the optical parametric case. We define
1032: the quadratures $X_1=(S_1+S_1^+)/\sqrt{2}$, $Y_1=i(S_1-S_1^+)/\sqrt{2}$
1033: and similarly for the polarition mode;
1034: these can be measured e.g. by converting spin excitations to light.
1035: Correlations between the modes appear due to dynamical evolution and
1036: squeezing is found in certain quadratures of the sum and difference modes
1037: $X_-=(X_1-X_D)/\sqrt{2}$ and $Y_+=(Y_1+Y_D)\sqrt{2}$. In the language of
1038: harmonic oscillators, the positions in mode $1$ and $2=D$ are correlated
1039: ($X_1\simeq X_D$), while the momenta are anti-correlated ($Y_1\simeq -Y_D$).
1040: For small number
1041: of excitations the sum and difference modes obey standard commutation
1042: relations $\left[X_\alpha,Y_\beta\right]=-i\delta_{\alpha,\beta}$
1043: where $\alpha,\beta=+,-\,{\rm or}\,1,D$. A quadrature $Y_\pm$ is squeezed
1044: when $\Delta Y_\pm(t)^2<1/2$ and the Heisenberg limit corresponds to
1045: $\Delta Y_\pm(t)^2\sim 1/N$.
1046:
1047: We find that squeezing is optimal under
1048: conditions of four-photon resonance
1049: ($\delta_1=\delta_2$) and in the limit of $\eta\gg 1$ (Fig. 6).
1050: Evolution leads to squeezing of $Y_+$ and $X_-$, anti-squeezing of $Y_-$
1051: and
1052: $X_+$. The squeezing in $Y_+$ reaches a minimum value at $t=t^*$ after
1053: which the growing fluctuations in $X_+$ give rise to increased noise in
1054: $Y_+$.
1055: Note that the total number of excitations (both modes) in the system,
1056: equal to $\langle X_+^2+X_-^2+Y_+^2+Y_-^2 \rangle$,
1057: grows exponentially with time (Fig. 6c).
1058: Specifically, in the case $g_1=g_2=g$ and thus
1059: $\gamma_1=\gamma_2\equiv\gamma$, for $\xi t>1$, we have:
1060:
1061: \begin{eqnarray}
1062: \left(\Delta Y_+(t)\right)^2 &=&1/2\left\{e^{-2\xi t} +
1063: \frac{3\gamma_L+\kappa/\eta}{2\xi}\right. \\ \nonumber
1064: &+& \left. \left(\frac{\gamma_L+\kappa/\eta}{2\xi}\right)^2e^{2\xi t}\right\}
1065: \label{quadt}
1066: \end{eqnarray}
1067:
1068: where we have neglected terms of higher order in $\gamma_L/\xi$ and
1069: $\kappa/(\eta\xi)$.
1070: The maximum amount of squeezing is obtained after an interaction time
1071: $t^*$ such that $e^{-2\xi t^*}=(\gamma_L+\kappa/\eta)/2\xi$ and is given by
1072:
1073: \begin{equation}
1074: (\Delta Y_+)^2=\frac{5\gamma_L+3\kappa/\eta}{4\xi}
1075: \label{squeez1}
1076: \end{equation}
1077:
1078: i.e. of the order of the damping rate divided by the coherent interaction rate.
1079:
1080: Since both the interaction parameter $\xi$ and the relaxation rate of the
1081: polariton $\gamma_D=\gamma_L+\kappa/\eta$ depend on the single photon
1082: detuning $\Delta$ (Fig. 6a), we find that squeezing is optimized for
1083: \begin{equation}
1084: \Delta_{opt}=\gamma\sqrt{\frac{5|\Omega_1|^2}{3|\Omega_2|^2}\frac{|g|^2N}
1085: {\gamma\kappa}}
1086: \end{equation}
1087: and with this optimal value of the detuning, the squeezing reaches a minimum
1088: value of
1089: \begin{equation}
1090: \left(\Delta Y_{+opt}\right)^2 = \frac{\sqrt{15/4}}
1091: {\sqrt{|g|^2N/\gamma\kappa}}.
1092: \label{maxsqueez}
1093: \end{equation}
1094: Note that the denominator is equal to the atomic
1095: density-length product multiplied by the empty cavity finesse and can
1096: easily
1097: exceed $10^4$ even for modest values of the density-length product and
1098: cavity finesse. We further emphasize that typical generation rate
1099: resulting in such optimal squeezing
1100: $\Omega_1\Omega_2/\Delta_{opt}$ can easily be on the order of fraction of
1101: MHz. In such a case other decoherence mechanisms are negligible.
1102: Doppler shifts can also be disregarded as long as all fields are
1103: co-propagating.
1104:
1105: For the ``degenerate'' version of the interaction (i.e. with identical final
1106: states for the spin flips, see Fig. 4b), the effective hamiltonian can be
1107: written as
1108:
1109: \begin{equation}
1110: H_{eff} = i\hbar\xi(\hat{S}^{\dagger 2}-\hat{S}^2)
1111: \end{equation}
1112:
1113: where the limit $\eta\gg 1$ has been used to write $P_D\simeq -S$, with
1114: $S=1/\sqrt{N}\Sigma_{gb}$ the spin flip operator. In this
1115: case the correlations lead to squeezing of $X=(S+S^\dagger)/\sqrt{2}$ and
1116: anti-squeezing of $Y=i(S-S^\dagger)/\sqrt{2}$. The analysis for this
1117: configuration is
1118: very similar to the non-degenerate version, in particular the maximum amount
1119: of squeezing achievable is also given by an expression of the form
1120: (\ref{maxsqueez}).
1121:
1122: We can now obtain a condition for achieving Heisenberg-limited spin
1123: squeezed
1124: states, i.e. $\left(\Delta Y_{+opt}\right)^2\simeq 1/N$. We see from
1125: (\ref{squeez1}) that this requires
1126:
1127: \begin{equation}
1128: \xi \sim N\Gamma
1129: \end{equation}
1130:
1131: where $\Gamma=(5\gamma_L+4\kappa/\eta)/4$ is the effective damping rate of the
1132: system. This is in complete agreement with the estimate based on our
1133: simple bosonic model of section III (\ref{heislimit}).
1134: In terms of the single photon Rabi
1135: frequency $g$, the cavity decay rate $\kappa$, the spontaneous emission rate
1136: $\gamma$ and the number of atoms $N$, the condition for achieving some
1137: squeezing i.e. $(\Delta Y_+)^2<1/2$ is
1138:
1139: \begin{equation}
1140: |g|^2N > \kappa\gamma
1141: \end{equation}
1142:
1143: which can be easily achieved in the laboratory since it simply
1144: corresponds to the condition that the
1145: density length product multiplied by the cavity finesse be larger than
1146: one.
1147: In the cavity QED regime of strong coupling $|g|^2\sim\kappa\gamma$, very
1148: strong quantum correlations i.e. $(\Delta Y_+)^2\sim 1/\sqrt{N}$
1149: between atoms can be produced.
1150: In order to obtain Heisenberg limited spin squeezed states i.e.
1151: $(\Delta Y_+)^2 \sim 1/N$, one requires a more stringent condition
1152:
1153: \begin{equation}
1154: |g|^2\sim N\kappa\gamma
1155: \end{equation}
1156:
1157: which can be fulfilled only in the strong coupling regime of cavity QED
1158: for a limited number of atoms. Note that this regime has been achieved
1159: experimentally by
1160: several groups \cite{cavityQED} and would allow for Heisenberg limited
1161: spin squeezing for as many as $\sim 10^3$ atoms.
1162: We have analyzed in this paper the situation of a running-wave
1163: cavity, so that all atoms couple equally apart from a possible phase to
1164: the cavity mode irrespective of their position.
1165: In order to fulfill the cavity QED regime, small cavity volume is needed
1166: i.e. standing wave cavities.
1167: For atoms in such a cavity the coupling to the cavity mode is position
1168: dependent and it becomes necessary to localize atoms accurately at the
1169: antinodes of a trapping mode. Note that significant experimental progress
1170: has been made towards this direction by several groups \cite{1atom}.
1171: Once the atoms are well localized in the cavity,
1172: the interaction can proceed via a neighbouring mode $b$ (e.g. different
1173: from the trapping mode $a$) so that
1174: for atoms localized within a small region in the cavity the two modes have
1175: essentially the same wavelength and atoms would therefore couple equally
1176: to the $b$ mode as well, irrespective of their position.
1177:
1178: \section{Discussion and Conclusion}
1179:
1180: We have reviewed Ramsey spectroscopy and the use of spin squeezed states
1181: in precision measurements of this type. With the experimental motivation
1182: of minimizing the phase accuracy in phase estimation with Ramsey
1183: fringes, we introduced a particular class of squeezed states. These states
1184: lead to Heisenberg limited phase accuracy and we developed various
1185: pictorial representations for them. The strong similarities of these
1186: representations of spin squeezed states to those of squeezed states of
1187: light suggests an analogy extending to the type of interaction that gives
1188: rise to squeezing. We are thus lead to consider the so-called "counter
1189: twisting" Hamiltonian, which has been shown to lead to maximal spin
1190: squeezing. We have studied this model for spin squeezing in the presence
1191: of a dissipation mechanism and analyzed the effect of damping and finite
1192: system size on the amount of squeezing achievable with such an
1193: interaction. The analysis was based on a decorrelation approximation to
1194: the BBGKY hierarchy of equations of motion, followed by the use of a
1195: linear transformation which in the limit of large number of atoms
1196: $1/N\rightarrow 0$ "contracts" the angular momentum operators onto bosonic
1197: operators. This allows for the systematic inclusion of finite system
1198: size effects. It appears that Heisenberg limited spin squeezed states may
1199: be produced when the single atom nonlinearity exceeds the single atom loss
1200: rate. In this case the maximum number of atoms that can be lost before
1201: quantum correlations are destroyed to the point of compromising the spin
1202: squeezing is of the order $\Delta N\sim\log{N}$. For spin squeezing at a
1203: more modest level than the Heisenberg limit, larger number of atoms may be
1204: lost without compromising the squeezing, indicating the stronger
1205: sensitivity of spin squeezed states to dissipation for larger amounts of
1206: squeezing.
1207:
1208: We have also presented in detail a scheme based on the interaction of
1209: coherent classical light with an optically dense ensemble of atoms that
1210: leads to an effective coherent spin-changing interaction involving pairs
1211: of atoms. Atoms may be transferred to the same final state leading to spin
1212: squeezing (analogous to squeezing of light by degenerate OPO) or to
1213: different final states in this case leading to quantum correlations
1214: between
1215: different atomic modes (analogous to quantum correlations between
1216: electromagnetic modes by non-degenerate OPO).
1217: We have shown that this process is robust with respect to realistic
1218: decoherence mechanisms and can result in rapid generation of correlated
1219: (spin squeezed) atomic ensembles.
1220: The amount of correlations created by this effective interaction can be
1221: simply expressed in terms of the single photon Rabi frequency $g$, the
1222: atomic spontaneous emission rate $\gamma$ and the cavity decay rate
1223: $\kappa$.
1224: We find that
1225: the generation of spin squeezed states requires $g^2N\sim\kappa\gamma$,
1226: which can easily be achieved in low finesse cavities with e.g. room
1227: temperature atomic vapours. Very strongly correlated states can be
1228: produced when the strong coupling regime $g^2\sim\kappa\gamma$ of cavity
1229: QED is achieved and the generation of Heisenberg limited spin squeezed
1230: states requires $g^2\sim N\kappa\gamma$.
1231: The effective interaction rate $\xi=\Omega_1\Omega_2/\Delta$ which depends
1232: on the Rabi-frequency of two applied classical fields $\Omega_{1,2}$ and a
1233: detuning from an atomic transition $\Delta$ can be fast and is
1234: controllable.
1235: Furthermore, the resulting spin excitations can be easily converted into
1236: photons on demand, which facilitates applications in quantum information
1237: processing.
1238: Possible applications involving high-precision measurements in atomic clocks
1239: can be also foreseen.
1240:
1241: We thank M.Fleischhauer, J.I.Cirac, V.Vuletic, S.Yelin and P.Zoller
1242: for helpful discussions.
1243: This work was supported by the NSF through the grant to the ITAMP.
1244:
1245: %%%%%%%%%%%%%%%%
1246:
1247: \appendix
1248:
1249: \section{Ramsey Spectroscopy}
1250:
1251: In Ramsey spectroscopy \cite{Wineland}, a collection of $N$ two-level atoms
1252: are made to interact
1253: with two separated fields (in time or in space). The lower and upper
1254: states (refered to as ground and excited state) have
1255: an energy difference $\hbar\omega_0$ and atoms will thus acquire a different
1256: dynamical phase $e^{-i E t/\hbar}$ depending on which state they are in.
1257: The effect of properly chosen electromagnetic fields is to perform a
1258: transformation that prepares the atoms
1259: in a superposition of the two states $|g\rangle$ and $|e\rangle$.
1260: The different
1261: parts of the wavefunction of atoms (corresponding to the ground and excited
1262: state) acquire a relative phase due to dynamical evolution and when the
1263: inverse transformation is applied, an interference effect is obtained.
1264: An exact parallel with the Mach-Zender interferometer can be drawn
1265: \cite{Dowling}:
1266: the transformation preparing atoms in a superposition of ground and excited
1267: states
1268: is equivalent to the transformation that lets a photon incident
1269: on a beam splitter
1270: explore the two arms of an interferometer. The relative phase acquired in
1271: the two atomic states during free evolution of duration $T$ is the equivalent
1272: of the relative phase acquired by photons travelling in the arms of the
1273: interferometer.
1274: Finally, the second pulse that performs the inverse
1275: transformation on atoms is the equivalent of the recombination of signals from
1276: the two interferometer arms on a beam splitter.
1277: At the end of this sequence, the
1278: number of atoms in either states, equivalent to the number of photons
1279: from either output of the final beam splitter, is measured.
1280: In this way, the signal
1281: measured depends on the acquired relative phase which can thus be estimated
1282: with some accuracy.
1283:
1284: We will now quantify this more precisely: let the frequency of the applied
1285: electromagnetic pulses be $\omega$,
1286: and the time delay between the two zones of interaction be $T$.
1287: The duration and strength of the applied fields are
1288: chosen so as to lead to $\pi/2$ pulses, i.e. transformation of the
1289: atomic state according to
1290:
1291: \begin{eqnarray}
1292: |e\rangle & \rightarrow & \frac{|e\rangle -i |g\rangle}{\sqrt{2}}
1293: \nonumber \\
1294: |g\rangle & \rightarrow & \frac{|e\rangle +i |g\rangle}{\sqrt{2}}.
1295: \label{pi2pulse}
1296: \end{eqnarray}
1297:
1298: During their free evolution between the two zones, atoms in the ground and
1299: excited states acquire a relative phase $\phi$ which, in a frame rotating
1300: with the frequency of the applied field, is $\phi=(\omega-\omega_0)T$.
1301:
1302: Before entering the first interaction zone, the atoms are prepared in
1303: their lower state $|g\rangle$ and at the exit of the second zone, the
1304: number of atoms in states $|e\rangle$ and $|g\rangle$ is measured.
1305:
1306: For simplicity, we consider the case when the first zone leads to a $\pi/2$
1307: pulse and the second one a $-\pi/2$ pulse.
1308: The picture of angular momentum is particularly well suited to discuss the
1309: Ramsey interferometric scheme and leads to an intuitive pictorial
1310: representation of the scheme.
1311: The Schwinger angular momentum operators are defined as
1312:
1313: \begin{eqnarray}
1314: \hat{J}_x & = & (\hat{\Sigma}_{eg}+\hat{\Sigma}_{ge})/2 \nonumber \\
1315: \hat{J}_y & = & (\hat{\Sigma}_{eg}-\hat{\Sigma}_{ge})/2i \nonumber \\
1316: \hat{J}_z & = & (\hat{\Sigma}_{ee}-\hat{\Sigma}_{gg})/2
1317: \end{eqnarray}
1318:
1319: where $\hat{\Sigma}_{\mu\nu}=\sum_{j=1}^{N}|\mu\rangle_{jj}\langle \nu|$
1320: are collective operators. In terms of these, a single $\pi/2$ pulse
1321: (\ref{pi2pulse}) is
1322: represented by a rotation of the pseudo angular momentum vector around the
1323: x-axis
1324: by an angle $\pi/2$. For a single atom we have the correspondence
1325: $|\uparrow\rangle=|e\rangle$ and $|\downarrow\rangle=|g\rangle$. Under a
1326: $\pi/2$ rotation about the x-axis, the state $\uparrow\rangle$ transforms
1327: to $|J_y=-1/2\rangle=(|\uparrow\rangle-i|\downarrow\rangle)/\sqrt{2}$ as
1328: indicated in (\ref{pi2pulse}). For $N$ atoms, we can think of the $N$
1329: individual spin ${\tiny \frac{1}{2}}$ particles combining to form a pseudo
1330: angular momentum
1331: vector of length $J=N/2$. The state of the collection of $N$ atoms can then be
1332: represented by appropriate superpositions of the states $|J,M\rangle$ where
1333: $-J\leq M\leq J$. Of course, only states within the completely symmetric
1334: subspace of the full $2^N$-dimensional Hilbert space can be represented in
1335: this way, which is justified since the coherent interaction of the
1336: electromagnetic fields with the atoms couple only to this symmetric subspace
1337: (i.e. all atoms couple equally to the fields).
1338:
1339:
1340: Free evolution in the rotating frame corresponds to
1341: rotation of the angular momentum around the z-axis at an angular velocity
1342: $\omega-\omega_0$. The whole Ramsey scheme can then be
1343: represented by the sequence: $\pi/2$ rotation about x-axis, $\phi$ rotation
1344: about the z-axis and $-\pi/2$ rotation about the x-axis. This is the
1345: transformation perfomed by the unitary operator
1346:
1347: \begin{equation}
1348: \hat{U}(\phi)=e^{i \pi/2 \hat{J}_x}e^{-i\phi \hat{J}_z}e^{-i \pi/2
1349: \hat{J}_x}
1350: \end{equation}
1351:
1352: where $\phi=(\omega-\omega_0)T$ as before.
1353: At the end of the scheme, the number of atoms in states $|e\rangle$ and
1354: $|g\rangle$ is measured, or equivalently their difference $\hat{J}_z(\phi)$
1355: where
1356:
1357: \begin{eqnarray}
1358: \hat{J}_z(\phi) &=& \hat{U}(\phi)^\dagger\hat{J}_z\hat{U}(\phi) \nonumber
1359: \\
1360: &=& \hat{J}_z\cos\phi-\hat{J}_x\sin\phi .
1361: \end{eqnarray}
1362:
1363: The Ramsey signal is thus
1364:
1365: \begin{equation}
1366: \langle \hat{J}_z(\phi)\rangle = \langle \hat{J}_z \rangle \cos\phi -
1367: \langle \hat{J}_x \rangle\sin\phi
1368: \label{jzphi}
1369: \end{equation}
1370:
1371: and its variance $\Delta J_z(\phi)$ is
1372:
1373: \begin{eqnarray}
1374: \Delta J_z(\phi) &=& \left[(\Delta J_z)^2\cos^2\phi+(\Delta J_x)^2\sin^2\phi
1375: \right.
1376: \nonumber \\
1377: &-& \left. \cos\phi\sin\phi(\langle \hat{J}_x \hat{J}_z+\hat{J}_z \hat{J}_x \rangle -2 \langle \hat{J}_z \rangle
1378: \langle \hat{J}_x \rangle) \right]^{1/2}
1379: \end{eqnarray}
1380:
1381: where the variance is defined as $(\Delta A)^2=\langle \hat{A}^2 \rangle -
1382: \langle \hat{A} \rangle^2$.
1383: From the signal one wants
1384: to estimate the phase $\phi$ and thus the frequency difference
1385: $\omega-\omega_0$. The phase
1386: accuracy achievable from such a measurement is related to the signal
1387: variance (the ``noise'') by
1388:
1389: \begin{equation}
1390: \delta\phi(\phi)=\frac{\Delta J_z(\phi)}{|\frac{\partial \langle
1391: \hat{J}_z(\phi)
1392: \rangle}{\partial\phi}|} .
1393: \end{equation}
1394:
1395: For states such that $\langle \hat{J}_x \rangle =0$ (all the states we
1396: will consider in this paper are of this type), the sensitivity
1397: $|\partial \langle \hat{J}_z(\phi) \rangle /\partial\phi|$ is maximal for
1398: $\phi=\pm\pi/2$ and the phase accuracy can be expressed as
1399:
1400: \begin{equation}
1401: \delta\phi(\pm\pi/2) = \frac{\Delta J_x}{|\langle \hat{J}_z \rangle |}.
1402: \label{dphimin}
1403: \end{equation}
1404:
1405: Since $\Delta J_x$ and $\langle \hat{J}_z \rangle$ depend on the initial
1406: state,
1407: we see that different initial states lead to different phase accuracies.
1408: Of particular importance is the accuracy achievable when all atoms are prepared
1409: in the same initial state. In this case the state of the atomic ensemble is a
1410: pure state, but it is however an uncorrelated state
1411: of the atomic ensemble (i.e. it can be
1412: factorized $|\Psi\rangle = \prod_{j=1}^{N} |\psi\rangle_j$).
1413:
1414:
1415: Consider the case of uncorrelated atoms for which all atoms have been
1416: prepared in the lower state $|g\rangle$,
1417: sometimes called a Bloch state.
1418: The state of the atomic ensemble can thus be expressed in terms of
1419: eigenstates of the collective angular momentum operators as
1420:
1421: \begin{equation}
1422: \prod_{j=1}^{N}|g\rangle_j = |J=N/2,J_z=-N/2\rangle
1423: \label{uncostate}
1424: \end{equation}
1425:
1426: where $J=N/2$ since there are $N$ $2$-level atoms, equivalent to $N$ spin
1427: ${\tiny \frac{1}{2}}$ particles.
1428: For such a state, the expectation value of the angular momentum operators and
1429: their variances are calculated to be
1430: $\langle \hat{J}_x \rangle = \langle \hat{J}_y \rangle = 0$, $\langle \hat{J}_z \rangle = -J$,
1431: $\Delta J_x = \Delta J_y = \sqrt{J/2}$ and $\Delta J_z = 0$. The signal and
1432: its variance are thus
1433:
1434: \begin{eqnarray}
1435: \langle \hat{J}_z(\phi) \rangle &=& -J\cos\phi \nonumber \\
1436: \Delta J_z(\phi) &=& \sqrt{J/2}\sin\phi .
1437: \end{eqnarray}
1438:
1439: The maximum sensitivity is achieved at $\phi=\pm\pi/2$
1440:
1441: \begin{equation}
1442: \delta\phi(\pm\pi/2) = \frac{1}{\sqrt{2J}} = \frac{1}{\sqrt{N}}
1443: \end{equation}
1444:
1445: which is the standard quantum limit (SQL). Performing the
1446: experiment on $N$ independent atoms all prepared in the same initial
1447: state is thus equivalent to repeating the experiment on one atom $N$ times and
1448: leads to an expected $1/\sqrt{N}$ factor of improvement in accuracy
1449: over the one atom result $\Delta S_x/\langle \hat{S}_z \rangle =1$. This
1450: is the best accuracy achievable with atoms all prepared in the same
1451: initial pure quantum state. The number of atoms detected in the upper state,
1452: given by $\langle \hat{N}_+(\phi) \rangle=N/2+\langle \hat{J}_z(\phi) \rangle$,
1453: and its variance are shown in Fig.1a.
1454:
1455: There is a lower bound on the phase accuracy, set by Heisenberg's
1456: uncertainty principle, $\Delta J_i \Delta J_j \geq {\tiny
1457: \frac{1}{2}}|\langle [\hat{J}_i,\hat{J}_j]\rangle |$ where $i,j=x,y,z$. It
1458: is straightforward to show that
1459:
1460: \begin{equation}
1461: \delta\phi\geq\frac{1}{N}
1462: \end{equation}
1463:
1464: which is known as the Heisenberg limit.
1465:
1466: We now see from (\ref{dphimin})
1467: that in order to surpass the SQL, the atomic ensemble must be
1468: prepared in a state such that
1469: $\Delta J_x/|\langle J_z \rangle | \leq 1/\sqrt{N}$, which is a necessary and
1470: sufficient condition for entanglement of an atomic ensemble
1471: \cite{squ-cold}. It is thus important to have a state for which the
1472: variance $\Delta J_x$ is reduced compared to its value for the uncorrelated
1473: state (\ref{uncostate}) while maintaining a large value for
1474: $\langle J_z \rangle$
1475: so that the amplitude of the signal $\langle \hat{J}_z(\phi) \rangle =
1476: \langle \hat{J}_z \rangle \cos\phi$ is not compromised \cite{Holland}.
1477: Such states which have reduced uncertainty in one observable $\Delta
1478: J_x$ (at the expense of the conjugate observable $\Delta J_y$ having
1479: increased fluctuations) have been called spin-squeezed states \cite{Ueda}.
1480:
1481:
1482: %%%%%%%%%%%%%%%%
1483:
1484: \section{Spin squeezed states - Wigner function representation}
1485:
1486: We now consider the Wigner function representation of the states
1487: $|\psi(a)\rangle$. The Wigner distribution of general angular-momentum states
1488: \cite{Agarwal} is obtained from an expansion of the density operator in terms
1489: of the multipole operators
1490:
1491: \begin{equation}
1492: \hat{\rho}=\sum_{k=0}^{2J}\sum_{q=-k}^{+k}\rho_{kq}\hat{T}_{kq}
1493: \end{equation}
1494:
1495: where the multipole operators are
1496:
1497: \begin{eqnarray}
1498: \hat{T}_{kq} & = & \sum_{m=-J}^{+J}\sum_{m'=-J}^{+J}(-1)^{J-m}\sqrt{2k+1}
1499: \left(\begin{array}{ccc}J & k & J \\ -m & q & m' \end{array}\right)
1500: \nonumber \\
1501: & \times & |J,m\rangle\langle J,m'|
1502: \end{eqnarray}
1503:
1504: and $\left(\begin{array}{ccc}J & k & J \\ -m & q & m' \end{array}\right)$ is
1505: the usual Wigner $3j$ symbol. The wigner distribution is then given by
1506:
1507: \begin{eqnarray}
1508: W(\theta,\phi) &=& \sum_{k=0}^{2J}\sum_{q=-k}^{+k}Y_k^q(\theta,\phi)\rho_{kq}
1509: \end{eqnarray}
1510:
1511: where
1512: $\rho_{kq}=\langle \hat{T}_{kq} \rangle = {\rm Tr}[\hat{\rho}\hat{T}_{kq}]$ and
1513: $Y_k^q(\theta,\phi)$ are the spherical harmonics.
1514: In Fig. 7, the Wigner function for the state $|\psi(-1)\rangle$ clearly
1515: shows
1516: the way in which this state has a large negative expectation value for
1517: $\hat{J}_z$, reduced variance in $\hat{J}_x$ and increased variance in
1518: $\hat{J}_y$.
1519:
1520: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1521:
1522: \begin{figure}[ht]
1523: \begin{center}
1524: \epsfig{file=figB1.eps,width=8.5cm}
1525: \leavevmode
1526: \end{center}
1527: \vspace*{2ex}
1528: \caption{Wigner function representation of the state $|\psi(a)\rangle$ with
1529: $a=-1)$. Plotted is the surface $r(\theta,\phi)=W(\theta,\phi)$, showing the
1530: large and negative value of $\langle \hat{J}_z \rangle$, reduced variance
1531: $\Delta J_x$ and correspondingly increased variance $\Delta J_y$.}
1532: \end{figure}
1533:
1534: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1535:
1536: %%%%%%%%%%%%%%%%%%%
1537:
1538: \section{Adiabatic elimination of excited state in Raman scattering}
1539:
1540: From the Hamitonian (\ref{hamram}), we obtain the equations of motion for
1541: the cavity mode and the ground state coherence $\Sigma_{gb_1}$
1542:
1543: \begin{eqnarray}
1544: \dot{a} &=& -\kappa a -i g_1^* \Sigma_{b_1a_1} -i g_2^*\Sigma_{ga_2}+F_a(t)
1545: \\
1546: \dot{\Sigma}_{gb_1} &=& -(\gamma_0-i\delta_1)\Sigma_{gb_1}+i\Omega_1
1547: \Sigma_{a_1b_1}-i g_1^*a^\dagger \Sigma_{ga_1} \nonumber \\
1548: &+& F_{gb_1}(t)
1549: \label{c2}
1550: \end{eqnarray}
1551:
1552: and the optical polarizations associated with Stokes emission evolve according
1553: to
1554:
1555: \begin{eqnarray}
1556: \dot{\Sigma}_{b_1a_1} &=& -[\gamma-i(\Delta-\delta_1)]\Sigma_{b_1a_1}
1557: -i\Omega_1\Sigma_{b_1g}\nonumber \\
1558: &-& i g_1 a(\Sigma_{b_1b_1}-\Sigma_{a_1a_1})+
1559: F_{b_1a_1}(t) \\
1560: \dot{\Sigma}_{a_1g} &=& -(\gamma+i\Delta)\Sigma_{a_1g}-i\Omega_1^*
1561: (\Sigma_{a_1a_1}-\Sigma_{gg})\nonumber \\
1562: &+& i g_1^*a^\dagger\Sigma_{b_1g}+F_{a_1g}(t)
1563: \end{eqnarray}
1564:
1565: where we assume that population in the excited state $|a_1\rangle$ decays
1566: towards $|b_1\rangle$ at a rate $\gamma_1$, towards $|g\rangle$ at a rate
1567: $\gamma_2$ and we assume a dephasing rate $\gamma_0$ for ground state
1568: coherences ($\gamma=(\gamma_1+\gamma_2)/2$ and $\gamma\gg\gamma_0$).
1569:
1570:
1571: We proceed by adiabatic elimination of optical polarizations associated with
1572: Stokes emission. To this end we assume large single-photon
1573: detuning $\Delta \gg \gamma$ and to first order in
1574: $\hat{a}$ we obtain ($\Sigma_{gg}\sim N$)
1575: \begin{eqnarray}
1576: \Sigma_{b_1a_1} &=& \frac{\Omega_1}{\Delta}(1-i\frac{\gamma}{\Delta})
1577: \Sigma_{b_1g}+i\frac{F_{b_1a_1}(t)}{\Delta} \\
1578: \Sigma_{a_1g} &=& \frac{\Omega_1^*}{\Delta}N(1+i\frac{\gamma}{\Delta})-i
1579: \frac{F_{a_1g}(t)}{\Delta}
1580: \end{eqnarray}
1581:
1582: which we substitute in (\ref{c2}) and obtain for the ground state spin
1583: flip operator
1584: $S_1=\Sigma_{gb_1}/\sqrt{N}$
1585:
1586: \begin{eqnarray}
1587: \dot{S_1} &=& -\left[(\gamma_0+\gamma_L)-i(\delta_1+\delta_L)\right] S_1
1588: -i\frac{g_1^*\sqrt{N}\Omega_1}{\Delta}a^\dagger \nonumber \\
1589: &+& {\bar F}_{S_1}(t)
1590: \end{eqnarray}
1591:
1592: where $\gamma_L=\gamma|\Omega_1|^2/\Delta^2$ is an optical pumping rate,
1593: $\delta_L=|\Omega_1|^2/\Delta$ is the light shift and ${\bar F}_{S_1}(t)$
1594: is a modified noise force. Light shifts can be
1595: incorporated in a redefinition of the energies and we ignore them in the
1596: remainder of this paper. Since the ground state decoherence rate is typically
1597: very small, we also assume $\gamma_L\gg\gamma_0$ and in that limit the new
1598: $\delta$-correlated noise forces have correlations
1599:
1600: \begin{eqnarray}
1601: \langle {\bar F}_{S_1}(t){\bar F}_{S_1}^\dagger(t') \rangle &=& 2\gamma_L
1602: \frac{\gamma_2}{\gamma}\delta(t-t') \\
1603: \langle {\bar F}_{S_1}^\dagger(t){\bar F}_{S_1}(t') \rangle &=& 2\gamma_L
1604: \frac{\gamma_1}{\gamma}\delta(t-t')
1605: \end{eqnarray}
1606:
1607: %%%%%%%%%%%%%%%%%%%%%%%%
1608:
1609: \section{Adiabatic elimination of bright polariton}
1610:
1611: After adiabatic elimination of the excited state $|a_1\rangle$, the
1612: relevant equations of motion are
1613:
1614: \begin{eqnarray}
1615: \dot{S}_1^\dagger &=& -(\gamma_L+i\delta_1)S_1^\dagger+i\chi a +
1616: {\bar F}_{S_1}^\dagger(t)
1617: \nonumber \\
1618: \dot{a} &=& -\kappa a-i\chi^*S_1^\dagger-ig_2^*\Sigma_{ga_2}+F_a(t)
1619: \nonumber \\
1620: \dot{S}_2 &=& -(\gamma_L+i\delta_2)S_2-i\frac{\Omega_2^*}{\sqrt{N}}
1621: \Sigma_{ga_2}+{\bar F}_{S_2}(t)
1622: \nonumber \\
1623: \dot{\Sigma}_{ga_2} &=& -(\gamma+i\delta_2)\Sigma_{ga_2}-i\Omega_2\sqrt{N}S_2
1624: -ig_2N a \nonumber \\
1625: &+& F_{ga_2}(t)
1626: \label{spineqns}
1627: \end{eqnarray}
1628:
1629: where $S_2 = \Sigma_{gb_2}/\sqrt{N}$.
1630:
1631: From (\ref{spineqns}) and (\ref{polaritons}) and in the limit of large
1632: ratio of
1633: speed of light in vacuum to group velocity of Stokes photons
1634: $\eta=|g_2|^2N/|\Omega_2|^2\gg1$, we obtain the equations of motion in terms
1635: of bright and dark polaritons
1636:
1637: \begin{eqnarray}
1638: \dot{S}_1^\dagger &=& -(\gamma_L+i\delta_1)S_1^\dagger+
1639: i\chi(\frac{P_D}{\eta}+P_B) +F_{S_1}^\dagger(t)
1640: \label{eqspinflip}
1641: \\
1642: \dot{P}_D &=& -(\kappa/\eta+\Gamma_2)P_D-i\frac{\chi^*}{\sqrt{\eta}}S_1^\dagger
1643: -\frac{\kappa-\Gamma_2}{\sqrt{\eta}}P_B\nonumber \\
1644: &+& F_D(t)
1645: \label{eqdarkpol}
1646: \\
1647: \dot{P}_B &=& -(\kappa+\Gamma_2/\eta)P_B-i\chi^*S_1^\dagger-ig_2\sqrt{N}
1648: {\tilde \Sigma}_{ga_2}\nonumber \\
1649: &+&F_B(t)
1650: \label{eqpolb} \\
1651: \dot{{\tilde \Sigma}}_{ga_2} &=& -\Gamma{\tilde \Sigma}_{ga_2}-ig_2\sqrt{N}P_B
1652: +{\tilde F}_{ga_2}(t)
1653: \label{eqoptcoh}
1654: \end{eqnarray}
1655:
1656: where $\Gamma_2=\gamma_L+i\delta_2$, $\Gamma=\gamma+i\delta_2$ and
1657: ${\tilde \Sigma}_{ga_2}=\Sigma_{ga_2}/\sqrt{N}$ and noise forces were
1658: modified
1659: appropriately. Note that in the picture of dark and bright polaritons,
1660: only
1661: the bright polariton is coupled to the excited state through the optical
1662: coherence $\Sigma_{ga_2}$.
1663:
1664: Under adiabatic conditions, the bright polariton evolves slowly (on a typical
1665: timescale $T$) and we can solve perturbatively in $1/T$. The equations
1666: (\ref{eqpolb}) and (\ref{eqoptcoh}) are of the form $\dot{{\bf x}}=
1667: -{\bf M}.{\bf x}+{\bf y}$, where ${\bf x}$ is the vector $(P_B,{\tilde
1668: \Sigma}_{ga_2})$, ${\bf M}$ is a $2\times 2$ matrix and ${\bf y}$ is a
1669: source term
1670:
1671: \begin{eqnarray}
1672: \frac{d}{dt}
1673: \left[\begin{array}{c}P_B \\ {\tilde \Sigma}_{ga_2}\end{array}\right]
1674: &=&-\left(\begin{array}{cc}\kappa & ig_2\sqrt{N} \\ ig_2\sqrt{N} &
1675: \Gamma\end{array}\right)
1676: \left[\begin{array}{c}P_B \\ {\tilde \Sigma}_{ga_2}\end{array}\right]
1677: \nonumber \\
1678: &+&
1679: \left[\begin{array}{c}-i\chi^*S_1^\dagger-\frac{\kappa}{\eta}P_D+F_B(t) \\
1680: {\tilde F}_{ga_2}(t)\end{array}\right]
1681: \label{mateqn}
1682: \end{eqnarray}
1683:
1684: where we have used $\kappa\gg\gamma_L$ and where $F_B(t)$ and ${\tilde
1685: F}_{ga_2}(t)$ are appropriate noise forces.
1686: These equations can be solved
1687: easily to first order by ${\bf x}^{(0)}(t)={\bf M}^{-1}.{\bf y}$, higher
1688: order approximations yielding ${\bf x}^{(n)}(t)={\bf M}^{-1}.[{\bf y}-
1689: \dot{{\bf x}}^{(n-1)}(t)]$.
1690:
1691: We can rewrite
1692: \begin{equation}
1693: \frac{|g_2|^2N}{\kappa\gamma} \sim 3\pi\times\left(\frac{N}{V}L\lambda^2\right)
1694: \times{\mathcal F}
1695: \end{equation}
1696:
1697: i.e. the density length product multiplied by the cavity finesse, so
1698: that with densities corresponding to room temperature atomic vapours, optical
1699: wavelengths and finesse of order $100$ this quantity is already of order
1700: $\sim 10^4$.
1701: We can thus assume that $|g_2|^2N/(\kappa\gamma)\gg 1$ and solve in powers
1702: of $\kappa\gamma/(|g_2|^2N)$.
1703:
1704: We see from (\ref{mateqn}) that ${\bf x}^{(n)}(t)$ is of order
1705: $[\kappa\gamma/(|g_2|^2N)]^{(n+1)}$ and thus solving to lowest order we
1706: find
1707:
1708: \begin{eqnarray}
1709: P_B &=& \frac{1}{|g_2|^2N}[-i\Gamma\chi^*S_1^\dagger-
1710: \frac{\kappa\Gamma}{\sqrt{\eta}}P_D+\Gamma F_B(t) \nonumber \\
1711: &-& ig_2F_{ga_2}(t)]
1712: \end{eqnarray}
1713:
1714: so that when $\eta\gg 1$,
1715:
1716: \begin{eqnarray}
1717: a &\simeq& \frac{P_D}{\eta}+P_B \nonumber \\
1718: &\simeq& \frac{P_D}{\eta}+\frac{1}{|g_2|^2N}[-i\Gamma\chi^*S_1^\dagger
1719: +\Gamma F_B(t) \nonumber \\
1720: &-& ig_2F_{ga_2}(t)].
1721: \end{eqnarray}
1722:
1723: The coupled equations of motion for the dark state polariton (\ref{eqdarkpol})
1724: and the spin flip (\ref{eqspinflip}) then become
1725:
1726: \begin{eqnarray}
1727: \dot{S}_1^\dagger &=& (\frac{|g_1|^2}{|g_2|^2}\gamma_L-\gamma_L-i\delta_1)
1728: S_1^\dagger + i\frac{\chi}{\sqrt{\eta}}P_D +{\tilde F}_{S_1}^\dagger(t)
1729: \label{eqfinspin}
1730: \\
1731: \dot{P}_D &=& -(\kappa/\eta+\gamma_L+i\delta_2)P_D-i\frac{\chi^*}{\sqrt{\eta}}
1732: S_1^\dagger + {\tilde F}_D(t)
1733: \label{eqfinpol}
1734: \end{eqnarray}
1735:
1736: where ${\tilde F}_{S_1}^\dagger(t)$ and ${\tilde F}_D(t)$ are modified noise
1737: forces with correlations
1738:
1739: \begin{eqnarray}
1740: \langle {\tilde F}_D(t){\tilde F}_D^\dagger(t') \rangle &=& \frac{\kappa}{\eta}
1741: +2\gamma_L\frac{\gamma_2}{\gamma} \\
1742: \langle {\tilde F}_D^\dagger(t){\tilde F}_D(t') \rangle &=& 2\gamma_L\frac{\gamma_1}{\gamma} \\
1743: \langle {\tilde F}_{S_1}(t){\tilde F}_{S_1}^\dagger(t') \rangle &=& 2\gamma_L\frac{\gamma_2}{\gamma} \\
1744: \langle {\tilde F}_{S_1}^\dagger(t){\tilde F}_{S_1}(t') \rangle &=& 2\gamma_L\frac{\gamma_1}{\gamma} \\
1745: \end{eqnarray}
1746:
1747: and all other correlations can be neglected. The coherent part of the
1748: interaction can thus be obtained from an effective hamiltonian
1749:
1750: \begin{equation}
1751: H_{eff} = \frac{\hbar\chi}{\sqrt{\eta}}S_1P_D+{\rm h.c.}
1752: \end{equation}
1753:
1754: where the interaction rate is $\chi/\sqrt{\eta}=g_1\Omega_1^*|\Omega_2|/(|g_2|
1755: \Delta)$.
1756:
1757: We note (\ref{eqfinpol}) that cavity losses are strongly suppressed in the
1758: limit $\eta \gg 1$.
1759: Indeed, subsequent to the large group velocity reduction \cite{slow-light},
1760: the polariton is almost purely atomic and the excitation leaks very slowly
1761: out of the medium.
1762: The equation of motion for coherence $S_1^+$ (\ref{eqfinspin}) contains
1763: a loss term
1764: (due to isotropic spontaneous emission) and a linear gain term (due to emission
1765: into bright polariton).
1766: The two can compensate each other. However the linear phase-insensitive
1767: amplification is also accompanied by
1768: correspondingly increased fluctuations, represented by new Langevin forces
1769: $\tilde{F}_D(t),\tilde{F}_{S_1}^+(t)$.
1770: In the case that $g_1=g_2$ and when all Rabi frequencies are taken to be
1771: real, we have the interaction rate
1772: $\xi=\chi/\sqrt{\eta}=\Omega_1\Omega_2/\Delta$.
1773:
1774: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1775:
1776: \def\etal{\textit{et al.}}
1777: \begin{thebibliography}{99}
1778: %%%%%%%%%%%%%%%%%%%%%%%%%%%
1779:
1780: \bibitem{Kimble-Houches} H.J. Kimble, in {\it Fundamental systems in
1781: quantum optics}, edited by J. Dalibard, J.-M. Raimond and J. Zinn-Justin
1782: (North-Holland 1990).
1783:
1784: \bibitem{Walls-Milburn} D.F. Walls and G.J. Milburn, {\it Quantum Optics} (Springer, Berlin, 1994).
1785:
1786: \bibitem{Q-teleport}
1787: K. Mattle {\it et al.}, Phys. Rev. Lett. {\bf 76} 4656, (1996);
1788: Z.Y.Ou {\it et. al}, Phys. Rev. Lett. {\bf 68}
1789: 3663 (1992)
1790:
1791: \bibitem{Longdist} L.-M. Duan, M.D. Lukin, J.I. Cirac and P. Zoller,
1792: quant-ph/0105105.
1793:
1794: \bibitem{Ueda} M. Kitagawa and M. Ueda, Phys. Rev. A {\bf 47}, 5138 (1993).
1795:
1796: \bibitem{Bouwmeester} D. Bouwmeester, A.K. Ekert, A. Zeilinger (eds.),
1797: {\it The physics of quantum information}, (Springer , New York, 2000).
1798:
1799: \bibitem{Nielsen-Chuang} M.A. Nielsen and I.L. Chuang., {\it Quantum computation and quantum information}, (Cambridge University Press, New York, 2000).
1800:
1801: \bibitem{Wineland} D.J. Wineland, {\it et al.}, Phys. Rev. A {\bf 46}, R6797
1802: (1992); {\it ibid} {\bf 50}, 67 (1994); J.J. Bollinger {\it et al.}
1803: Phys. Rev. A {\bf 54} R4649 (1996);
1804: S. F. Huelga {\it et al.} Phys. Rev. Lett. {\bf 79}, 3865 (1997); V.
1805: Meyer {\it et al.}, Phys. Rev. Lett. {\bf 86} 5870 (2001).
1806:
1807: \bibitem{Sachdev} S.Sachdev, {\it Quantum phase transitions},
1808: (Cambridge University Press, New York, 1999).
1809:
1810: \bibitem{Wineland2} C.A. Sackett {\it et al.}, Nature {\bf 404}, 256
1811: (2000).
1812:
1813: \bibitem{Polzik} B. Julsgaard, A. Kozhekin and E. Polzik, Nature
1814: {\bf 413} 400 (2001).
1815:
1816: \bibitem{Lloyd} H. Touchette and S. Lloyd, Phys. Rev. Lett. {\bf 84},
1817: 1156 (2000).
1818:
1819: \bibitem{Meystre} P.Meystre, {\it Atom Optics}, (Springer , New York,
1820: 2001).
1821:
1822: \bibitem{Paper1} A. Andr\'e, L.-M. Duan and M.D. Lukin, quant-ph/0107075
1823:
1824: \bibitem{slow-light} L.\ V.\ Hau {\it et al.},
1825: Nature \textbf{397}, 594 (1999); M. Kash {\it et al.}
1826: Phys. Rev. Lett. {\bf 82}, 5229 (1999); D. Budker {\it et al.},
1827: Phys. Rev. Lett. {\bf 83}, 1767 (1999).
1828:
1829: \bibitem{stop-light} C. Liu, Z. Dutton, C.H. Behroozi and L. Hau, Nature {\bf 409}, 490 (2001); D.F. Phillips {\it et al.}, Phys. Rev. Lett. {\bf 86}, 783 (2001).
1830:
1831: \bibitem{darkpolar} M.D. Lukin, S.F. Yelin and M. Fleischhauer, Phys. Rev. Lett. {\bf 84}, 4232 (2000); M. Fleischhauer and M.D. Lukin, Phys. Rev. Lett. {\bf 84}, 5094 (2000).
1832:
1833: \bibitem{Q-mem} J.I. Cirac, P. Zoller, H.J. Kimble and H. Mabuchi, Phys. Rev. Lett. {\bf 78}, 3221 (1997); S.J. Enk, J.I. Cirac and P. Zoller, Science {\bf 279}, 205 (1998).
1834:
1835: \bibitem{Kasevich} P. Bouyer and M.A. Kasevich, Phys. Rev. A {\bf 56}
1836: R1083 (1997).
1837:
1838: \bibitem{Holland} T. Kim {\it et al.} Phys. Rev. A {\bf 57}, 4004 (1998).
1839:
1840: \bibitem{Vardi}J.R. Anglin and A. Vardi, Phys. Rev. A {\bf 64}, 013605
1841: (2001)
1842:
1843: \bibitem{Courtens} F.T. Arecchi {\it et al.}, Phys. Rev. A {\bf 6} 2211 (1972).
1844:
1845: \bibitem{squ-cold} L.-M. Duan, A. S\o rensen, J. I. Cirac and P. Zoller Phys. Rev. Lett. {\bf 85}, 3991 (2000); A. S\o rensen, L.-M. Duan, J. I. Cirac and P. Zoller, Nature {\bf 409},63 (2001).
1846:
1847: \bibitem{squ-Polzik} A.~Kuzmich, K.~M\o lmer, and E.~S.~Polzik,
1848: Phys. Rev. Lett. {\bf 79}, 4782 (1997);
1849: J.~Hald, {\it et al.}, {ibid} 1319 (1999).
1850:
1851: \bibitem{squ-Bigelow}
1852: A. Kuzmich, L. Mandel and N.P. Bigelow, Phys. Rev. Lett. {\bf 85}, 1594 (2000).
1853:
1854: \bibitem{Bouchoule} I.~Bouchoule and K.~M\o lmer, quant-ph/0105144
1855:
1856: \bibitem{brightpolar} M. Fleischhauer and M.D. Lukin, quant-ph/0106066
1857:
1858: \bibitem{Ketterle} J. Stenger {\it et al.}, Nature {\bf 396}, 345
1859: (1998).
1860:
1861: \bibitem{Heinzen} D.J. Heinzen, R. Wynar, P.D. Drummond and K.V. Kheruntsyan, Phys. Rev. Lett. {\bf 84}, 5029 (2000).
1862:
1863: \bibitem{cavityQED} P.W.H. Pinkse {\it et al.}, J. Mod. Opt. {\bf 47},
1864: 2769
1865: (2000); C.J. Hood {\it et al.}, Phys. Rev. A {\bf 64}, 033804 (2001).
1866:
1867: \bibitem{1atom} J. Ye {\it et al.}, Phys. Rev. Lett. {\bf 83}, 4987
1868: (1999); P.W.H. Pinkse {\it et al.}, Nature {\bf 404}, 365 (2000).
1869:
1870: \bibitem{Dowling} J.P. Dowling, Phys. Rev. A {\bf 57} 4736 (1998).
1871:
1872: \bibitem{Agarwal} J.P. Dowling, G.S. Agarwal and W.P. Schleich, Phys. Rev. A
1873: {\bf 49}, 4101 (1994).
1874:
1875:
1876: \end{thebibliography}
1877:
1878:
1879: \end{document}
1880:
1881:
1882:
1883: