quant-ph0611148/mezk.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: \documentclass[pra, showpacs, twocolumn, floatfix]{revtex4}
3: \usepackage{graphicx}
4: \usepackage{amsmath, amsfonts, amssymb, bm}
5: \usepackage[]{psfrag}
6: \psfragscanoff \setlength{\parindent}{0pt}
7: \begin{document}
8: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
9: 
10: \title{Localization of atomic ensembles via superfluorescence}
11: 
12: \author{Mihai \surname{Macovei}}
13: \email{mihai.macovei@mpi-hd.mpg.de}
14: \affiliation{Max-Planck Institute for Nuclear Physics, 
15: Saupfercheckweg 1,D-69117 Heidelberg, Germany}
16: 
17: \author{J\"org \surname{Evers}}
18: \email{joerg.evers@mpi-hd.mpg.de}
19: \affiliation{Max-Planck Institute for Nuclear Physics, 
20: Saupfercheckweg 1,D-69117 Heidelberg, Germany}
21: 
22: \author{Christoph H. \surname{Keitel}}
23: \email{keitel@mpi-hd.mpg.de}
24: \affiliation{Max-Planck Institute for Nuclear Physics, 
25: Saupfercheckweg 1,D-69117 Heidelberg, Germany}
26: 
27: \author{M. Suhail \surname{Zubairy}}
28: \email{zubairy@physics.tamu.edu} 
29: \affiliation{Institute for Quantum Studies and Dept. of Physics,
30: Texas A\&M University, College Station, Texas 77843, USA}
31: \affiliation{Max-Planck Institute for Nuclear Physics, 
32: Saupfercheckweg 1,D-69117 Heidelberg, Germany}
33: 
34: \date{\today}
35: 
36: 
37: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
38: \begin{abstract}
39: The sub-wavelength localization of an ensemble of atoms 
40: concentrated to a small volume in space is investigated.
41: The localization relies on the interaction of the ensemble
42: with a standing wave laser field. The light scattered
43: in the interaction of standing wave field and atom
44: ensemble depends on the position of the ensemble relative
45: to the standing wave nodes. This relation can be described
46: by a fluorescence intensity profile, which depends on the standing
47: wave field parameters, the ensemble properties, and which
48: is modified due to collective effects in the ensemble of nearby 
49: particles. We demonstrate that the intensity profile can
50: be tailored to suit different localization setups.
51: Finally, we apply these results to two localization schemes.
52: First, we  show how to localize an ensemble fixed at a
53: certain position in the standing wave field. Second, we discuss
54: localization of an ensemble passing through the standing wave
55: field.
56: \end{abstract}
57: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
58: \pacs{32.50.+d, 32.80.-t, 42.50.Gy, 42.50.Fx}
59: \maketitle
60: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
61: 
62: 
63: \section{INTRODUCTION}
64: Nano-technology requires an accurate control of the interacting 
65: components, both in terms of detection and preparation.
66: This is a major motivation for the considerable attention that
67: was devoted recently to sub-wavelength localization of single 
68: particles. Several remarkable schemes 
69: were proposed to achieve this goal~\cite{QCW,TW,HL,KDR,LcZ,LPK,AK}. 
70: Of related interest is the problem of localizing and distinguishing two 
71: nearby particles. Optical resolution of two molecules at the 
72: nanometer scale and 
73: manipulation of their degree of entanglement was experimentally 
74: demonstrated in~\cite{He}, and the collective interaction between 
75: pairs of oriented nanostructures was considered in~\cite{Bar}. 
76: Also the measurement of the relative position of two atoms 
77: assisted by spontaneous emission~\cite{Su} or the measurements 
78: of interparticle separations on a scale smaller than 
79: the emission wavelength~\cite{JZ} were investigated in detail. 
80: Somewhat related, considerable effort is devoted to quantum 
81: lithography~\cite{Lt,hemmer}. The above localization schemes, 
82: however, have in common that they apply to the localization
83: of a single particle or to the measurement of relative position
84: of two individual particles. In many cases of interest,
85: however, an ensemble of particles is concentrated to a
86: small region in space such that the ensemble properties
87: become relevant, while the properties of the individual
88: ensemble constituents cannot be resolved or are rapidly 
89: fluctuating in time.
90: 
91: Therefore, here we describe a scheme capable of localizing an ensemble
92: of two-level atoms which are bunched together in a volume much 
93: smaller than an emission wavelength. 
94: Possible realizations include small clusters, few-atom 
95: impurities, or atoms trapped, e.g., in optical lattices.
96: The localization relies
97: on the coherent interaction with a standing-wave electromagnetic 
98: field.  Since the interatomic
99: distances are small, the atoms interact collectively via the 
100: environmental vacuum modes. One consequence of this is the
101: appearance of superfluorescence, i.e., the scattered light
102: intensity scales with the number of atoms $N$ squared 
103: ($I \propto N^{2}$).
104: %
105: We find that the fluorescence light emitted collectively 
106: by the ensemble is a function of the ensemble position 
107: in the standing wave. In particular, for suitable standing
108: wave parameters and for ensemble positions
109: around the nodes of the standing wave field, the emitted
110: fluorescence intensity sharply drops to a minimum over a narrow
111: spatial region. The narrow width of the dip in the spatial
112: intensity profile is a direct consequence of the collectivity.
113: %
114: Since this collective fluorescence intensity profile
115: is our main observable, we discuss the profile in detail
116: in terms of the available free parameters, and show that
117: the profile can be tailored to suit a given localization
118: setup.
119: %
120: Based on these results, we then propose two schemes which exploit 
121: this spatial fluorescence intensity profile to localize an 
122: ensemble of atoms.
123: First, we assume the sample to be fixed within the standing wave 
124: field. In this case, the spatial fluorescence intensity profile can 
125: be scanned along the standing wave axis by changing the relative
126: phase of the laser fields forming the standing wave.
127: A continuous measurement of the intensity of the scattered
128: light throughout this scan reveals the position of the 
129: sample on a sub-wavelength scale. We further show that 
130: this setup also enables one
131: to measure the distance between two samples, the number
132: of atoms in a sample, or the linear dimension of the sample.
133: %
134: Second, we consider an atom cluster flying through the standing
135: wave field. Here a scanning is not possible due 
136: to the short interaction time of ensemble and standing wave
137: field. Rather, the absolute intensity of the scattered
138: light can be used to recover the crossing position of the
139: ensemble.
140: %
141: 
142: The article is organized as follows.
143: In Sec.~\ref{sec-theory}, we present our theoretical model 
144: for the ensemble interaction with the standing wave field
145: and derive the intensity profile as our main observable.
146: Sec.~\ref{sec-results} consists of three parts.
147: In the first part Sec.~\ref{sec-int}, we in detail study
148: the collective fluorescence profile numerically.
149: In Sec.~\ref{sec-sweep}, the results are applied to the
150: localization of an ensemble fixed in the standing wave field.
151: The last part Sec.~\ref{sec-single} discusses the localization
152: of an ensemble flying through the standing wave field.
153: Finally, Sec.~\ref{sec-summary} discusses and summarizes the
154: results.
155: 
156: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
157: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
158: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
159: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
160: \section{\label{sec-theory}THEORY}
161: In the usual mean-field, dipole, and rotating-wave approximations the interaction of such an atomic sample with 
162: an external laser field and the surrounding vacuum modes, in a frame rotating with the laser frequency $\omega_{L}$, 
163: is described by the Hamiltonian $H=H_{0} + H_{L} + H_{I}$ where 
164: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
165: \begin{align} 
166: H_{0}&= \sum_{k}\hbar(\omega_{k}-\omega_{L})a^{\dagger}_{k}a_{k} + \sum^{N}_{j=1}\hbar(\omega_{0j}-\omega_{L})S_{zj}, 
167: \nonumber \allowdisplaybreaks[2] \\
168: %
169: H_{L}&=\sum^{N}_{j=1}\hbar\bigl( \Omega(\vec r_{j})S^{+}_{j} + \Omega^{\ast}(\vec r_{j})S^{-}_{j} \bigr), 
170: \nonumber \allowdisplaybreaks[2]\\
171: %
172: H_{I}&= i\sum_{k}\sum^{N}_{j=1}(\vec g_{k}\cdot \vec d_{j}) a^{\dagger}_{k}S^{-}_{j}e^{-i\vec k\cdot \vec r_{j}}
173:  + \textrm{ H.c.}\,. \label{Hm}
174: \end{align}
175: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
176: Here $S^{\pm}_{j}$ are the raising and lowering operators for the $j$th atom, positioned at $\vec r_{j}$, and obeying the commutation 
177: relations $[S^{+}_{j},S^{-}_{l}]=2S_{zj}\delta_{jl}$, and $[S_{zj},S^{\pm}_{l}]=\pm S^{\pm}_{j}\delta_{jl}$ with $S_{zj}$
178: being the inversion operator. $a^{\dagger}$ and  $a$ are the radiation creation and annihilation operators satisfying 
179: the commutation relations $[a_{k},a^{\dagger}_{k^{'}}]=\delta_{kk^{'}}$, and $[a_{k},a_{k^{'}}]=[a^{\dagger}_{k},a^{\dagger}_{k^{'}}]=0$.
180: 
181: In Eq.~(\ref{Hm}), $H_{0}$ represents the free electromagnetic field (EMF) and free atomic Hamiltonians, respectively. 
182: The second term, i.e. $H_{L}$, describes the interaction of the atomic system with an external standing-wave coherent field. 
183: In general, the Rabi frequencies of the atoms in a standing wave are position-dependent since 
184: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
185: \begin{eqnarray*}
186: \Omega(\vec r_{j})=\Omega_{j}\cos{(\vec k_{L}\cdot \vec r_{j})}, 
187: \end{eqnarray*}
188: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
189: where $\Omega_{j} = (\vec d_{j} \cdot \vec E_{L})/\hbar$ while 
190: $E_{L}=|\vec E_{L}|$ is the amplitude of the electromagnetic field intensity with a wave vector $\vec k_{L}$ and 
191: $\vec d_{j}$ is the dipole moments of the atoms. 
192: %
193: Note that the scheme described here can also be generalized to
194: multiphoton transitions. Then, the Rabi frequency can be written as
195: \begin{equation}
196: \Omega(\vec r_{j})=\Omega_{j}^{(n)}\cos{(n\vec k_{L}\cdot \vec r_{j})}, 
197: \end{equation}
198: %
199: where $\Omega_{j}^{(n)}$ is a multiphoton Rabi-frequency arising
200: from an adiabatic elimination of intermediate states, and
201: $n$ denotes the number of photons involved in the multiphoton
202: process. In this way, the wavelength $\lambda=2\pi/k$ can be 
203: reduced to the effective wavelength $\lambda/n$, thus 
204: increasing the spatial resolution~\cite{hemmer}.
205: %
206: The last term in Eq.~(\ref{Hm}), $H_{I}$, takes into account the interaction 
207: of all atoms with the environmental vacuum modes. 
208: 
209: In the Born-Markov approximations the quantum dynamics of the driven multi-atom sample 
210: (each atom having identical transition frequency $\omega_{0}$) is governed by the master equation \cite{AGb,FS,KL,PR}:
211: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
212: \begin{align}
213: \frac{d}{dt}\rho(t) &+ \frac{i}{\hbar}[\tilde H_{0},\rho] = \nonumber \\
214: &\quad \sum^{N}_{i,j=1}\{\gamma_{ij}(\omega_{0})[S^{+}_{j},S^{-}_{l}\rho] + \textrm{H.c.} \}\,, \label{ME}\\
215: %
216: \tilde H_{0}&=\hbar\sum_{j}[\Delta S_{zj}/2 + \Omega(\vec r_{j})S^{+}_{j} + \textrm{H.c.}]\,.
217: \end{align}
218: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
219: Here, $\Delta = \omega_{0} - \omega_{L}$ is the detuning of atomic levels 
220: from the frequency of the driving field. Further,
221: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
222: \begin{eqnarray}
223: \gamma_{jl}(\omega_{0})=\chi_{jl}(\omega_{0}) + i\Omega_{jl}(\omega_{0}), \label{CP}
224: \end{eqnarray} 
225: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
226: where the collective parameters describing the mutual interactions among any atomic pair in the sample are given, respectively, 
227: by \cite{AGb,FS}
228: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
229: \begin{eqnarray}
230: \chi_{jl}(\omega) &=& \frac{3\gamma}{2}\bigl \{[1-\cos^{2}{\xi_{jl}}]\frac{\sin(\omega r_{jl}/c)}{\omega r_{jl}/c} + [1-3\cos^{2}{\xi_{jl}}] \nonumber \\
231: &\times&\bigl [\frac{\cos(\omega r_{jl}/c)}{(\omega r_{jl}/c)^{2}}- \frac{\sin(\omega r_{jl}/c)}{(\omega r_{jl}/c)^{3}}\bigr ]\bigr \}, \nonumber \\
232: \Omega_{jl}(\omega) &=& \frac{3\gamma}{4}\bigl \{[\cos^{2}{\xi_{jl}}-1]\frac{\cos(\omega r_{jl}/c)}{\omega r_{jl}/c} + [1-3\cos^{2}{\xi_{jl}}] \nonumber \\
233: &\times& \bigl [\frac{\sin(\omega r_{jl}/c)}{(\omega r_{jl}/c)^{2}} + \frac{\cos(\omega r_{jl}/c)}{(\omega r_{jl}/c)^{3}}\bigr ]\bigr \}, 
234: \label{ColP}
235: \end{eqnarray}
236: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
237: with $2\gamma = 4\omega^{3}_{0}d^{2}_{0}/(3\hbar c^{3})$ being the single-atom spontaneous decay rate. 
238: Here, we have assumed 
239: that all the dipole moments are identical and parallel, i.e. $d_{j}=d_{l} \cdots \equiv d_{0}$, and then $\xi_{jl}$ is the angle between the dipole moments $\vec d_{0}$ and $\vec r_{jl} = \vec r_{j} - \vec r_{l}$.
240: 
241: Inspecting the master equation (\ref{ME}) one can easily distinguish the part of it describing the coherent evolution of atoms
242: under the influence of the laser field, i.e. the term containing the Hamiltonian $\tilde H_{0}$, from that characterizing 
243: the collective spontaneous emission due to the vacuum modes, that is the terms proportional to $\rm{Re\{\gamma_{jl}(\omega_{0})\}}$.
244: The dipole-dipole interactions between the two-level atoms are described by the terms proportional to $\rm{Im\{\gamma_{jl}(\omega_{0})\}}$.
245: If the interparticle separations are small enough, that is $\omega r_{jl}/c \equiv k r_{jl} \to 0$ ($j \not = l$), then to second 
246: order in this parameter, Eqs.~(\ref{ColP}) reduce to 
247: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
248: \begin{align}
249: \chi_{jl}(k) &= \gamma\{1 - \frac{1}{5}(k r_{jl})^{2}[1-\frac{1}{2}\cos^{2}\xi_{jl}]\}, \nonumber \\
250: \Omega_{jl}(k) &= 3\gamma\{[\cos^{2}\xi_{jl} -1](2/kr_{jl}-kr_{jl}) + [1- 
251: \nonumber \\
252:  3&\cos^{2}\xi_{jl}]  
253: [(kr_{jl})^{-1} - kr_{jl}/4 + 2(kr_{jl})^{-3}]\}/8. \label{apColP}  
254: \end{align}
255: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
256: It is easy to realize that in this case $\chi_{jl} \to \gamma$ while $\Omega_{jl}$ reduces to the static dipole-dipole
257: interaction potential, i.e. 
258: \begin{eqnarray}
259: \Omega_{jl} = \frac{3\gamma}{4(kr_{jl})^{3}}\{1- 3\cos^{2}\xi_{jl}\}. \label{sdd}
260: \end{eqnarray}
261: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
262: For lower atomic densities the collective parameters $\chi_{jl}$ and $\Omega_{jl}$ vanish because the atoms react independently from 
263: each other in this particular case. Finally, the master equation (\ref{ME}) describes adequately driven atomic samples of any shapes 
264: providing that retardation effects are negligible.
265: 
266: When dealing with smaller atomic systems of an arbitrary irregular shape it is hard to specify the orientation of dipole moments
267: relative to the interparticle separations. That is to say, we do not have a privileged angular distribution of photons as they
268: are emitted equally in all directions. Since there is no information on the dipole orientations, we average over all directions
269: the dipole-dipole interaction potential. Interestingly, the static dipole-dipole interaction given by Eq.~(\ref{sdd}) vanishes in 
270: this case. Then, according to the second term in Eq.~(\ref{apColP}), the averaged dipole-dipole interactions among the two-level 
271: emitters are given by the expression:
272: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
273: \begin{eqnarray}
274: \Omega^{(av)}_{jl} = - \frac{\gamma}{2kr_{jl}}. \label{av_dd}
275: \end{eqnarray}
276: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
277: Thus, the dipole-dipole potential reduces from a short-range to a long-range interaction, although the radiators
278: are close to each others. Moreover, the influence of the dipole-dipole interactions, in a two-atom system, was shown 
279: to be negligible in practice for interparticle separations such that $kr_{jl} \ge \pi/10$ \cite{LMe}. 
280: 
281: In what follows we shall apply Eqs.~(\ref{ME}-\ref{av_dd}) to the localization of a small atomic system within an
282: emission wavelength. Suppose that the linear dimension of the atomic sample is much less than the emission wavelength 
283: (say, for instance, smaller than $0.1\lambda$). Under this assumption the two-level emitters are almost in an equivalent
284: position relative to the driving standing-wave laser and we can omit the atomic indices from the expression characterizing 
285: the Rabi frequency, i.e $\Omega(r_{j}) \approx \Omega(r) \equiv \Omega(x)$. The master equations (\ref{ME}) transforms then 
286: into: 
287: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
288: \begin{eqnarray}
289: \frac{d}{dt}\rho(t) &+& i[\tilde \Delta S_{z} + \Omega(x)(S^{+} + S^{-}) - \Omega_{d}S^{+}S^{-},\rho] \nonumber \\
290: &=& \gamma \{ [S^{+},S^{-}\rho] + [\rho S^{+},S^{-}] \}. 
291: \label{MER}
292: \end{eqnarray}
293: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
294: Here $\tilde \Delta = \Delta + \Omega_{d}$ and $\{ S^{\pm}, S_{z} \}$ are collective atomic operators satisfying 
295: the standard commutation relations of su(2) algebra \cite{AGb,FS,KL,PR} while $\Omega_{d}$ is the dipole-dipole interaction 
296: potential considered identical for all radiators. Note that Eq.~(\ref{MER}) describes a small driven 
297: atomic system that involves symmetrized multiparticle states only.
298: The antisymmetric states are decoupled from the dynamics within our current framework,
299: and in the following we assume that they are not populated initially.
300: The dipole-dipole interaction $\Omega_{d}$ 
301: considerably shifts the symmetric states from the field resonance if the interparticle separations 
302: are very small~\cite{MFK}. Even though
303: the hypothesis of identical dipole-dipole interactions is, in general, not fulfilled, it admits to solve analytically 
304: the above master equation in the long-time limit. Thus, in order to get some insight on how the dipole-dipole 
305: interactions affect the localization processes of a small system as a whole we shall accept the hypothesis of 
306: identical dipole-dipole interactions between any pair in the sample.
307: 
308: The solving procedure of Eq.~(\ref{MER}) was described in \cite{KL,PR} for running wave lasers.
309: Adopting it to the case of a standing wave field, one arrives at the steady-state solution
310: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
311: \begin{eqnarray}
312: \rho_{s} = Z^{-1}\sum^{N}_{n,m=0}C_{nm}(x)\bigl(S^{-}\bigr )^{n}\bigl(S^{+} \bigr)^{m}, \label{SSS}
313: \end{eqnarray}
314: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
315: where 
316: \begin{eqnarray*}
317: C_{nm}(x) &=& (-1)^{n+m}\alpha^{-n}(\alpha^{\ast})^{-m}a_{nm}, \\
318: a_{nm} &=& \frac{\Gamma(1+n+\beta)\Gamma(1+m+\beta^{\ast})}{n!m!\Gamma(1+\beta)\Gamma(1+\beta^{\ast})},
319: \end{eqnarray*}
320: with  $\alpha=i\Omega(x)/(\gamma + i\Omega_{d})$ and $\beta=i\tilde \Delta/(\gamma + i\Omega_{d})$.
321: The normalization constant $Z$ is chosen such that $\rm{Tr\{\rho_{s}\}=1}$, i.e. 
322: \begin{eqnarray*}
323: Z=\sum^{N}_{n,m=0}(-1)^{n+m}\alpha^{-n}(\alpha^{\ast})^{-m}a_{nm}{\rm Tr}\{(S^{-})^{n}(S^{+})^{m}\}.
324: \end{eqnarray*}
325: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
326: 
327: The trace can be performed using the following relations
328: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
329: \begin{eqnarray}
330: S^{+}|s,l \rangle = \sqrt{(s-l)(s+l+1)}|s,l+1\rangle, \nonumber \\ 
331: S^{-}|s,l \rangle = \sqrt{(s+l)(s-l+1)}|s,l-1\rangle, \label{DiO}
332: \end{eqnarray}
333: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
334: where the collective Dicke states $|s,l\rangle$, with $s=N/2$ and $-s \le l \le s$, are the eigenstates for the operator 
335: $S_{z}$ and the operator of the total "spin" $S^{2}$ \cite{PR}:
336: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
337: \begin{eqnarray*}
338: S_{z}|s,l \rangle = l|s,l\rangle, \\ 
339: S^{2}|s,l \rangle = s(s+1)|s,l\rangle.
340: \end{eqnarray*}
341: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
342: Thus, 
343: \begin{eqnarray}
344: &{}&{\rm Tr}\{(S^{-})^{n}(S^{+})^{m}\}=\sum^{s}_{l=-s}\langle l,s|(S^{-})^{n}(S^{+})^{m}|l,s\rangle \nonumber \\
345: &=& \delta_{n,m}\sum^{s}_{l=-s}\langle l,s|(S^{-})^{n}(S^{+})^{n}|l,s\rangle, \label{sp} 
346: \end{eqnarray}
347: and, then 
348: \begin{eqnarray}
349: Z=\sum^{N}_{n=0}a_{nn}|\alpha|^{-2n}\frac{(N+n+1)!(n!)^{2}}{(N-n)!(2n+1)!}. \label{Z}
350: \end{eqnarray}
351: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
352: 
353: The intensity of the collective resonance fluorescence emitted by driving the multiparticle system is calculated 
354: taking into account that this quantity is proportional to the first order atomic correlator, i.e. 
355: $I \propto \langle S^{+}S^{-}\rangle$. Then, using Eq.~(\ref{SSS} - \ref{Z}), one obtains:
356: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
357: \begin{eqnarray}
358: I(x) = Z^{-1}\sum^{N}_{k=1}C_{k-1k-1}(x)\frac{(N+k+1)!(k!)^{2}}{(N-k)!(2k+1)!}. \label{It}
359: \end{eqnarray}
360: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
361: 
362: 
363: \section{\label{sec-results}RESULTS}
364: We now turn to the discussion of Eq.~(\ref{It}).
365: First, we study the dependence of the collective fluorescence intensity
366: on the various external parameters. Second, we show how the
367: collective fluorescence intensity can be used to precisely locate
368: a sample of particles which is fixed in space inside the standing
369: wave field. Third, we discuss the localization of a collection 
370: of particles flying through the cavity field.
371: 
372: 
373: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  
374: \begin{figure}[t]
375: \includegraphics[height=4.5cm]{Fig1.eps}
376: \caption{\label{fig-1} The dependence of the collective steady-state resonance fluorescence intensity 
377: $I/N^{2}$ as function of $kx$. The solid, dashed and dotted curves are for $\Omega/(N\gamma)=100; 50$ and $25$, respectively. 
378: Other parameters are: $\Omega_{d}/\gamma=-10$,  $\Delta/(N\gamma)=0.5$ and $N=100$.}
379: \end{figure}
380: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  
381: 
382: 
383: \subsection{\label{sec-int}Collective fluorescence intensity}
384: 
385: Fig.~\ref{fig-1} shows the collective fluorescence intensity 
386: versus the position of the multiparticle collection in the
387: standing wave field. Note that $kx=\pi/2$ corresponds to
388: a node of the standing wave field, and thus the 
389: fluorescence intensity vanishes for particles located at 
390: this point in space. The parameters in Fig.~\ref{fig-1}
391: are number of atoms $N=100$, dipole-dipole coupling constant
392: $\Omega_{d}/\gamma=-10$, and detuning $\Delta/(N\gamma)=0.5$.
393: The solid line shows a standing wave Rabi frequency
394: $\Omega/(N\gamma)=100$, the dashed line is for 
395: $\Omega/(N\gamma)=50$, and the dotted one is for 
396: $\Omega/(N\gamma)=25$. It can be seen that with decreasing 
397: Rabi frequency, the width of the dip in the fluorescence
398: intensity around the node at $kx=\pi/2$ increases. Thus
399: a strong driving field allows for a narrow region in space
400: that leads to vanishing fluorescence intensity, while a
401: weaker field gives rise to fluorescence intensity over
402: a wider range of positions.
403: 
404: %
405: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  
406: \begin{figure}[t]
407: \includegraphics[width=8cm]{Fig2.eps}
408: \caption{\label{fig-2}(Color online) 
409: The steady-state collective resonance fluorescence intensity $I/N^{2}$ 
410: as function of $kx$  and $\bar \Delta = \Delta/(N\gamma)$. Here 
411: $N=100$, $\Omega/(N\gamma)=50$
412: and $\Omega_{d}/\gamma = -10$.
413: The black line on top of the surface plot indicates the position
414: $\bar \Delta = |\Omega_d|/\gamma = 10$.}
415: \end{figure}
416: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  
417: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  
418: \begin{figure}[t]
419: \includegraphics[width=8cm]{Fig3.eps}
420: \caption{\label{fig-3}(Color online) The steady-state collective resonance fluorescence intensity $I/N^{2}$ as function 
421: of $kx$ 
422:  and $\bar \Delta = \Delta/(N\gamma)$. Here $N=2$, $\Omega/(N\gamma)=50$
423: and $\Omega_{d}/\gamma = -5$. The black line on top of the surface 
424: plot indicates the position
425: $\bar \Delta = 5$ around which the two-atom symmetric collective state
426: is located.}
427: \end{figure}
428: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  
429: 
430: The second free parameter is the detuning $\Delta$ between the
431: driving field frequency and the bare transition frequency 
432: $\omega_0$ of the individual atoms in the sample. 
433: Fig.~\ref{fig-2} shows the dependence of the collective fluorescence
434: intensity versus the position of the sample on this detuning.
435: Note that the y-axis of this figure is a scaled detuning 
436: $\bar\Delta = \Delta/(N\gamma)$. Thus, $\bar\Delta=0$
437: corresponds to the resonance case $\Delta = 0$, whereas
438: $\bar\Delta > 0$ indicates $\Delta>0$ and thus 
439: $\omega_L < \omega_0$. A qualitative understanding of this figure
440: can be gained from the case of a two atom sample,
441: as shown in Fig.~\ref{fig-3}. In a collective state
442: basis, the two-atom sample corresponds to a collective ground state
443: $|g_a, g_b\rangle$ at energy 0, where each of the two atoms 
444: $a,b$ is in its respective ground state,
445: an excited collective state $|e_a, e_b\rangle$
446: at energy $2\hbar \omega_0$, and
447: a symmetric [antisymmetric] collective state at energy 
448: $\hbar(\omega_0 + \Omega_d)$ [$\hbar(\omega_0 - \Omega_d)$]:
449: %
450: \begin{subequations}
451: \begin{eqnarray}
452: |S\rangle = \frac{1}{\sqrt{2}}\left (|g_a, e_b\rangle  + |e_a, g_b\rangle \right ) \,,\\
453: |A\rangle = \frac{1}{\sqrt{2}}\left (|g_a, e_b\rangle  - |e_a, g_b\rangle \right ) \,.
454: \end{eqnarray}
455: \end{subequations}
456: %
457: In the limit
458: of small interparticle distance chosen in our analysis,
459: the asymmetric state decouples from the dynamics,
460: such that we are essentially left with a three-state
461: ladder system. In Figure~\ref{fig-3}, the symmetric state
462: is located at $\bar \Delta = 5$, since the dipole-dipole
463: coupling is chosen as $\Omega_d = -5\gamma$. 
464: %
465: %
466: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  
467: \begin{figure}[t]
468: \includegraphics[width=8cm]{Fig4.eps}
469: \caption{\label{fig-4}Collective fluorescence versus atom ensemble
470: position in the standing wave field for different sizes of the
471: ensemble. The parameters are $\Omega_{d}/\gamma = -5$,
472: $\Delta/\gamma=10$, and $\Omega/\gamma=100$.
473: The solid line is for number of atoms $N = 2$, the dashed shows
474: the case $N = 4$, and the dotted is for $N=8$.}
475: \end{figure}
476: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  
477: %
478: It can be seen that the width of the intensity dip is minimal
479: if the driving field frequency is in resonance with the
480: symmetric state, and becomes wider in moving away from the
481: resonance. 
482: No additional structure is visible around $\bar \Delta = -5$, 
483: where the asymmetric state is located, since it is
484: decoupled.
485: Note that for vanishing dipole-dipole interaction, $\Omega_d=0$, 
486: Figs.~\ref{fig-2} and \ref{fig-3} would exhibit features symmetric with
487: respect to both the planes given by $\bar{\Delta} = 0$
488: and $kx=\pi/2$.
489: 
490: These results from the two-atom case directly carry over to the
491: many-particle sample. The minimum width of the intensity
492: dip is close to the position where symmetric state combinations
493: can be expected, whereas no structure can be found 
494: towards asymmetric collective states.
495: A direct identification of the position of the detuning with
496: minimum dip width is difficult, however, since the collective-state
497: basis of a multi-particle sample includes many symmetric collective
498: states. In the example of Fig.~\ref{fig-2}, in a very small range
499: around the node $kx=\pi/2$ the dip at $\bar\Delta=10$ is narrowest,
500: but its width increases faster in moving away from the node
501: than it does for slightly lower values of $\bar\Delta$.
502: The many-particle
503: case also shows an additional structure at $\bar\Delta = 0$,
504: see Fig.~\ref{fig-2}, which, however, is not of interest for
505: our current localization scheme.
506: 
507: Finally, in Fig.~\ref{fig-4}, we show the dependence of the 
508: collective intensity on the number of particles in the sample.
509: This is different from the previous results, since by changing
510: the number of atoms, both the scaled Rabi frequency
511: $\Omega/(N\gamma)$ and the scaled detuning $\Delta/(N\gamma)$
512: are changed at the same time. It can be seen from Fig.~\ref{fig-4}
513: that for a given standing wave intensity and a given detuning,
514: varying the number of atoms in the ensemble changes the width
515: of the intensity dip at the nodes. This can be understood by
516: noting that a change of the number of atoms effectively shifts
517: the position of the symmetric state resonance. Since the
518: laser field frequencies are kept fixed in Fig.~\ref{fig-4}, 
519: this corresponds to moving along the $\bar\Delta$ axis
520: in Figs.~\ref{fig-2} and \ref{fig-3}. Therefore, different
521: widths of the intensity dip can be observed.
522: The maximum intensity changes with $N$ in Fig.~\ref{fig-4},
523: since in this figure the unscaled Rabi frequency $\Omega$ is 
524: kept fixed.
525: It should be noted that the parameters in Fig.~\ref{fig-4} are such 
526: that for any shown number of atoms, the scaled Rabi frequency
527: $\Omega/(N\gamma)$ dominates the dynamics. A further increase
528: of the number of atoms which leads to $|\Omega_d|/\gamma \gg \Omega/(N\gamma)$
529: shifts the relevant collective states out of the laser field
530: resonance such that the total fluorescence intensity vanishes.
531: 
532: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  
533: \begin{figure}[t]
534: \includegraphics[width=6cm]{Fig5.eps}
535: \caption{\label{fig-sweep}(Color online) 
536: Scanning-dip scheme. The figure depicts
537: a possible experimental implementation of our scheme. An atom ensemble
538: (green dot) is assumed fixed inside the standing wave field. A detector measures
539: the scattered fluorescence light, while the phase of the standing wave
540: field is varied. The black curve symbolizes the fluorescence intensity
541: profile as, e.g., shown in Fig.~\ref{fig-1}. If the intensity dip does
542: not coincide with the ensemble position, then a high intensity of 
543: fluorescence is detected.
544: But if the dip sweeps across the ensemble position, then the measured
545: intensity drops to a minimum over a narrow spatial range, thus providing a 
546: sub-wavelength localization.}
547: \end{figure}
548: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  
549: 
550: 
551: \subsection{\label{sec-sweep}Scanning-dip spectroscopy}
552: After the discussion of the collective fluorescence intensity
553: as our main observable, we now turn to the application of this
554: observable to the localization of a collection of atoms.
555: As the first setup, we consider a collection of atoms which is
556: fixed inside the standing wave field at an unknown position.
557: In order to detect the position of the sample, the total 
558: collective fluorescence intensity is continuously
559: monitored, which may already provide a coarse position measurement. 
560: Then the relative phase of the two counter-propagating
561: fields forming the standing wave is changed, such that the
562: nodal structure of the field shifts along the standing wave
563: propagation axis. Throughout this shift, the detected intensity
564: is modulated in time with the collective fluorescence 
565: intensity profile, as depicted in Fig.~\ref{fig-sweep}. 
566: If the intensity profile is located such that its dip does
567: not coincide with the actual position of the sample, then the 
568: intensity is near its maximum value, see Fig.~\ref{fig-sweep}(a).
569: But if the two positions coincide, then
570: the intensity vanishes, see Fig.~\ref{fig-sweep}(b).
571: Obviously, for this scheme it is desirable to have the intensity
572: dip as narrow as possible in order to achieve a localization
573: well below the usual diffraction limit. According to our results
574: of Sec.~\ref{sec-int}, this can be achieved by using a
575: strong standing wave field and by tuning it close to the
576: symmetric collective state resonance. For instance, the
577: solid curve in Fig.~\ref{fig-1} has a width of about
578: $\Delta(kx) = 0.02\pi$, corresponding to
579: $\Delta x = 0.01 \lambda$. Note, however, that the obtained
580: accuracy is also limited by the spatial size of the ensemble,
581: if the dip width is smaller than the linear dimension of the
582: sample.
583: 
584: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  
585: \begin{figure}[t]
586: \includegraphics[width=6cm]{Fig6.eps}
587: \caption{\label{fig-single}(Color online) 
588: Single-pass localization. (a) shows the schematic setup. An ensemble
589: indicated by the green dot passes through a standing wave field.
590: The intensity of the scattered fluorescence light is measured.  
591: (b) Via the fluorescence intensity profile, the measured intensity
592: can be related to the sub-wavelength particle position. The higher
593: the slope of the intensity profile, the better is the accuracy of
594: the localization. In this figure, the horizontal bars indicate 
595: two measurement outcomes with a certain uncertainty. The vertical
596: bars show the corresponding potential positions. The black curve is
597: a fluorescence intensity profile as, e.g., shown in Fig.~\ref{fig-1}.}
598: \end{figure}
599: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  
600: 
601: The present scheme can also be used to measure the distance 
602: between the center of masses of two ensembles by relating the required phase 
603: shifts between two positions with vanishing intensity to 
604: the measured intensity profile.
605: %
606: Further, since the width of the intensity dip depends on the number
607: of atoms in the ensemble if all other parameters remain fixed 
608: [see Fig.~\ref{fig-4}], this scheme can also be used to obtain
609: the number of atoms in the ensemble located in the standing wave field
610: by measuring the width of the intensity dip via the sweep of the
611: standing wave phase.
612: %
613: Finally, the spatial dimension of the sample can be measured from
614: the width of the intensity dip if the other parameters are known.
615: It should be noted that, in contrast to the precision position detection,
616:  these measurements relying on the
617: determination of the width of the intensity dip are relative
618: measurements in the sense that they do not require a reference 
619: to obtain the absolute standing wave phase. Only the change
620: of the phase is relevant, and thus also knowledge of the 
621: actual position of the sample within the standing wave field 
622: is not required for these relative measurements. 
623: The maximum attenuation of the fluorescence intensity  depends
624: on the  width of the collection relative to the width of the dip in the
625: intensity profile. If the profile dip is narrower than the 
626: sample, then the dip only affects part of the sample. Still,
627: this will result in a sudden reduction of the fluorescence
628: intensity, which is sufficient to determine the onset of the
629: overlap of intensity dip and atom sample. Also, a continuous
630: scan of the standing wave phase results in repeated intensity
631: dips over time which can be used to suppress statistical errors
632: in the measurement. 
633: 
634: \subsection{\label{sec-single}Single-pass localization}
635: In this section, we discuss a different experimental setup.
636: We now assume that a collection of atoms (atomic cluster) 
637: flies through a standing
638: wave field at an unknown position on the standing wave field axis.
639: The aim is to gain as much information on the position as possible
640: by measuring the collective resonance fluorescence. Obviously,
641: the scanning-dip scheme described in Sec.~\ref{sec-sweep} is not
642: suitable for this kind of setup, since the change of the standing
643: wave phase is too slow as compared to the interaction time
644: of field and atom ensemble. In the present scheme, we only
645: have to require that the time of flight $\tau_f$
646: through the standing
647: wave field is much larger than the time $\tau_s$ 
648: needed to evolve into the steady state. 
649: For example, for a thermal beam with velocity $300$~m/s and
650: a standing wave field width of $1$~mm, the flight time is
651: about $\tau_f = 3 \cdot 10^{-6}$~s. The steady-state time $\tau_s$ is of the order 
652: $(N\gamma)^{-1}$. For $\gamma = 10$~MHz and $N=10$, one obtains
653: $\tau_s = 10^{-8}$~s and thus $\tau_s \ll \tau_f$.
654: Note that the preparation of atom clusters has been discussed in~\cite{cluster}.
655: 
656: We now make use of the fact that position information can be gained from 
657: the absolute value of the scattered light intensity during the
658: flight of the ensemble through the field.
659: The schematic setup is shown in Fig.~\ref{fig-single}(a).
660: The measured intensity allows to fix a horizontal section in 
661: a collective fluorescence intensity plot versus ensemble position 
662: as shown in Fig.~\ref{fig-single}(b).
663: Ideally, this section provides a set of few discrete points
664: where the intensity profile crosses the measured intensity.
665: These points correspond to the potential positions of the
666: ensemble. From the two examples in Fig.~\ref{fig-single}(b)
667: it is clear that the localization for a given intensity measurement
668: with a measurement uncertainty becomes better with increasing slope
669: of the intensity profile. A high slope, however, leads to 
670: a pronounced plateau with almost constant intensity in between the
671: dips. This means that, in case of a narrow dip, for large parts of 
672: the single wavelength width only a rather inaccurate localization 
673: is possible. Therefore, in contrast to the sweep scheme in 
674: Sec.~\ref{sec-sweep}, in this setup a wide intensity dip 
675: is desirable in order to achieve sub-wavelength localization for all 
676: possible positions.
677: The reason is that for a wide intensity dip, wide plateaus in the 
678: intensity profile are avoided.
679: A wide intensity dip can be achieved,  for example,
680: by working with weaker standing wave fields or far away
681: from the symmetric collective state resonance.
682: 
683: 
684: 
685: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
686: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
687: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
688: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
689: \section{\label{sec-summary}DISCUSSION AND SUMMARY}
690: We have described a scheme to localize small atomic 
691: samples with sub-wavelength accuracy. The scheme relies on 
692: measuring the super-fluorescence radiation scattered in a 
693: standing wave field. We have demonstrated that external parameters 
694: such as the strength of the applied lasers or the detuning 
695: from the atomic resonance are convenient tools to tailor the 
696: localization region for a given experimental setup.
697: Based on these results, two possible experimental situations have been
698: considered. First, for fixed samples, a scanning-dip spectroscopy
699: was proposed. Here, the standing wave field phase is changed
700: in order to scan the fluorescence intensity profile along the
701: cavity axis in order to reveal the actual position of the sample.
702: This setup also allows for a number of relative measurements,
703: for example, of distance between two collections, of the number
704: of atoms in a sample, or of the linear dimension of the sample.
705: Second, for samples passing through the standing wave
706: field only once, a single-pass scheme was discussed, which relates
707: the maximum intensity measured to the passing position 
708: of the sample. Our scheme can be generalized to the multi-photon
709: case.
710: 
711: 
712: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
713: {\small $^\ast$ On leave from \it{Technical University of Moldova, Physics Department, 
714: \c{S}tefan Cel Mare Av. 168, MD-2004 Chi\c{s}in\u{a}u, Moldova.}}
715: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
716: 
717: 
718: 
719: 
720: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
721: \begin{thebibliography}{99}
722: \bibitem{QCW}
723: R. Quadt, M. Collett, and D. F. Walls, Phys. Rev. Lett. {\bf 74}, 351 (1995).
724: 
725: \bibitem{TW}
726: J. E. Thomas, L. J. Wang, Phys. Rep. {\bf 262}, 311 (1995).
727: 
728: \bibitem{HL}
729: M. Holland, S. Marksteiner, P. Marte, and P. Zoller, Phys. Rev. Lett. {\bf 76}, 3683 (1996).
730: 
731: \bibitem{KDR}
732: S. Kunze, K. Dieckmann and G. Rempe, Phys. Rev. Lett. {\bf 78}, 2038 (1997).
733: 
734: \bibitem{LcZ}
735: S. Qamar, S.-Y. Zhu, and M. S. Zubairy, Phys. Rev. A {\bf 61}, 063806 (2000); 
736: %
737: K. T. Kapale, S. Qamar, and M. S. Zubairy, Phys. Rev. A {\bf 67}, 023805 (2003); 
738: %
739: M. Sahrai, H. Tajalli, K. T. Kapale, and M. S. Zubairy, Phys. Rev. A {\bf 72}, 013820 (2005).
740: 
741: \bibitem{LPK} 
742: E. Paspalakis and P. L. Knight, Phys. Rev. A {\bf 63}, 065802 (2001); 
743: %
744: G. Liu, S. Gong, D. Cheng, X. Fan, and Z. Xu, Phys. Rev. A {\bf 73}, 025801 (2006).
745: 
746: \bibitem{AK} 
747: G. S. Agarwal, K. T. Kapale, J. Phys. B: At. Mol. Opt. Phys. {\bf 39}, 3437 (2006).
748:  
749: \bibitem{He} 
750: C. Hettich, C. Schmitt, S. Zitzmann, S. K\"{u}hn, I. Gerhardt, V. Sandoghder, Science {\bf 298}, 385 (2002).
751: 
752: \bibitem{Bar} 
753: M.~D. Barnes, P.~S. Krstic, P. Kumar, A. Mehta, and J. C. Wells, Phys. Rev. B {\bf 71}, 241303(R) (2005).
754: 
755: \bibitem{Su} 
756: L. Zheng, C. Li, Y. Li, and C. P. Sun, Phys. Rev. A {\bf 71}, 062101 (2005).
757: 
758: \bibitem{JZ} 
759: J.-T. Chang, J. Evers, M. O. Scully, and M. S. Zubairy, Phys. Rev. A {\bf 73}, 031803(R) (2006);
760: J.-T. Chang, J. Evers and M. S. Zubairy, Phys. Rev. A 74, 043820 (2006).
761: 
762: \bibitem{Lt} 
763: A. N. Boto, P. Kok, D.~S. Abrams, S.~L. Braunstein, C.~P. Williams, and J. P. Dowling, Phys. Rev. Lett. {\bf 85}, 2733 (2000); 
764: %
765: M. D'Angelo, M. V. Chekhova and Y. Shih, Phys. Rev. Lett.  {\bf 87}, 013602 (2001); 
766: %
767: J. Xiong, D.~Z. Cao, F. Huang, H.~G. Li, X.~J. Sun, and K. Wang Phys. Rev. Lett.  {\bf 94}, 173601 (2005); 
768: %
769: C. H. Keitel and  S. X. Hu, Appl. Phys. Lett. {\bf 80}, 541 (2002).
770: 
771: \bibitem{hemmer}
772: P. R. Hemmer, A. Muthukrishnan, M. O. Scully, and M. S. Zubairy, Phys. Rev. Lett.  {\bf 96}, 163603 (2006).
773: 
774: \bibitem{AGb} 
775: G. S. Agarwal, {\it Quantum Statistical Theories of Spontaneous Emission and their Relation to 
776: other Approaches} (Springer, Berlin, 1974).
777: 
778: \bibitem{FS} 
779: Z. Ficek, S. Swain, {\it Quantum Interference and Coherence: Theory and Experiments} (Springer, Berlin, 2005).
780: 
781: \bibitem{KL} 
782: S. Ya. Kilin, JETP {\bf 55}, 38 (1982).
783: 
784: \bibitem{PR} 
785: R. R. Puri, {\it Mathematical Methods of Quantum Optics}, Springer, Berlin 2001 (espacially chapter 12 and 
786: references therein).
787: 
788: \bibitem{LMe} 
789: G. Lenz and P. Meystre, Phys. Rev. A {\bf 48}, 3365 (1993). 
790: 
791: \bibitem{MFK} 
792: M. Macovei, Z. Ficek, and C. H. Keitel, Phys. Rev. A {\bf 73}, 063821 (2006).
793: 
794: \bibitem{cluster}
795: G. M. D'Ariano, N. Sterpi, and A. Zucchetti, 
796: Phys. Rev. Lett. {\bf 74}, 900 (1995).
797: 
798: \end{thebibliography}
799: \end{document}
800: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
801:  
802: 
803: 
804: 
805: 
806: 
807: \end{document}
808: 
809: