1: \documentclass[12pt]{article}
2: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: \usepackage[dvips]{graphicx}
4:
5: \newcommand{\beq}{\begin{equation}}
6: \newcommand{\eeq}{\end{equation}}
7: \newcommand{\bqa}{\begin{eqnarray}}
8: \newcommand{\eqa}{\end{eqnarray}}
9: \newcommand{\nn}{\nonumber}
10: \newcommand{\nl}[1]{\nn \\ && {#1}\,}
11: \newcommand{\erf}[1]{Eq.~(\ref{#1})}
12: \newcommand{\dg}{^\dagger}
13: \newcommand{\rt}[1]{\sqrt{#1}\,}
14: \newcommand{\smallfrac}[2]{\mbox{$\frac{#1}{#2}$}}
15: \newcommand{\bra}[1]{\langle{#1}|}
16: \newcommand{\ket}[1]{|{#1}\rangle}
17: \newcommand{\ip}[1]{\langle{#1}\rangle}
18: \newcommand{\bl}{{\bigl(}}
19: \newcommand{\br}{{\bigr)}}
20: \newcommand{\dbd}[1]{\frac{\partial}{\partial {#1}}}
21: %\newcommand{\cal}[1]{\it{#1}}
22:
23:
24: %\begin{document}
25:
26: \addtolength{\topmargin}{-30pt}
27: \addtolength{\textheight}{90pt}
28:
29:
30:
31: \title{Entangled states and collective nonclassical effects
32: in two-atom systems}
33: \author{Z. Ficek$^*$ and R. Tana\'s$^{\dagger}$}
34: \date{$^*$Department of Physics, School of Physical Sciences,
35: The University of Queensland, Brisbane, QLD 4072, Australia\\
36: \vspace{0.5cm}
37: $^{\dagger}$Nonlinear Optics Division, Institute of Physics,
38: Adam Mickiewicz University, Umultowska 85, 61-614 Pozna\'n, Poland}
39:
40: \begin{document}
41:
42: \maketitle
43:
44: \begin{abstract}
45: We propose a review of recent developments on entanglement and
46: non-classical effects in collective two-atom systems and present a
47: uniform physical picture of the many predicted
48: phenomena. The collective effects have brought into sharp focus some of
49: the most basic features of quantum theory, such as nonclassical states
50: of light and entangled states of multiatom systems. The entangled states
51: are linear superpositions of the internal states of the system which
52: cannot be separated into product states of the individual atoms.
53: This property is recognized as entirely quantum-mechanical effect and
54: have played a crucial role in many discussions of the nature of quantum
55: measurements and, in particular, in the developments of quantum
56: communications. Much of the fundamental interest in entangled states is
57: connected with its practical application ranging from quantum computation,
58: information processing, cryptography, and interferometry to atomic
59: spectroscopy.
60: \end{abstract}
61:
62: \newpage
63:
64: \section{Introduction}\label{ftsec1}
65:
66:
67: A central topic in the current studies of collective effects in multi-atom
68: systems are the theoretical investigations and experimental implementation
69: of entangled states to quantum computation and quantum information
70: processing~\cite{pv}. The term entanglement, one of the most intriguing
71: properties of multiparticle systems, was introduced by
72: Schr\"{o}dinger~\cite{schr} in his discussions of the foundations of
73: quantum mechanics. It describes a multiparticle system
74: which has the astonishing property that the results of a measurement on one
75: particle cannot be specified independently of the results of measurements on
76: the other particles. In recent years, entanglement has become of interest not
77: only for the basic understanding of quantum mechanics, but also because it
78: lies at the heart of many new applications ranging from quantum
79: information~\cite{beke,pok}, cryptography~\cite{ek} and quantum
80: computation~\cite{bar,gro} to atomic and molecular
81: spectroscopy~\cite{boll1,boll2}. These practical implementations all stem
82: from the
83: realization that we may control and manipulate quantum systems at the
84: level of single atoms and photons to store and transfer information in
85: a controlled way and with high fidelity.
86:
87: All the implementations of entangled atoms must contend with the conflict
88: inherent to open systems. Entangling operations on atoms must provide
89: strong coherent coupling between the atoms, while shielding the atoms
90: from the environment in order to make the effect of decoherence and
91: dissipation negligible. The difficulty of isolating the atoms from the
92: environment is the main obstacle inhibiting practical applications
93: of entangled states. The environment consists of a continuum of
94: electromagnetic
95: field modes surrounding the atoms. This gives rise to decoherence that
96: leads to the loss of information stored in the system. However, it
97: has been recognised that the collective properties of multi-atom systems
98: can alter spontaneous emission compared with the single atom case.
99: As it was first pointed out by Dicke~\cite{dic}, the interaction
100: between the
101: atomic dipoles could cause the multiatom system to decay with two
102: significantly different, one enhanced and the other reduced,
103: spontaneous emission rates. The presence of the reduced spontaneous
104: emission rate induces a reduction of the linewidth of the spectrum of
105: spontaneous emission~\cite{ftk87,buz}. This reduced (subradiant)
106: spontaneous emission implies that the multi-atom system can decohere
107: slower compared with the decoherence of individual atoms.
108:
109: Several physical realisations of entangled atoms have been proposed
110: involving trapping and cooling of a small number of ions or neutral
111: atoms~\cite{eich,deb,ber,tos}. This is the case with the lifetime of
112: the superradiant and subradiant states that have been demonstrated
113: experimentally with two barium ions confined in a spherical Paul
114: trap~\cite{eich,deb}. The reason for using cold trapped atoms or ions
115: is twofold. On the one hand, it has been realised that the trapped atoms
116: are essentially motionless and lie at a known and controllable distance
117: from one another, permitting qualitatively new studies of interatomic
118: interactions not accessible in a gas cell or an atomic beam~\cite{ag70}.
119: The advantage of the
120: trapped atoms is that it allows to separate collective effects,
121: arising from the correlations between the atoms, from the single-atom
122: effects. On the other hand it was discovered that cold trapped atoms
123: can be prepared in maximally entangled states that are isolated from
124: its environment~\cite{bbtk,afs,bcjd,bdj,tbk}.
125:
126: An example of maximally entangled states in a two-atom system are the
127: superradiant and subradiant states, which correspond to the symmetric
128: and antisymmetric combinations of the atomic dipole moments,
129: respectively. These states are created by the interaction between the
130: atoms and are characterized by different spontaneous decay rates that
131: the symmetric state decays with an enhanced, whereas the antisymmetric
132: state decays with a reduced spontaneous emission rate. The reduced
133: spontaneous emission rate of the antisymmetric state implies that the
134: state is weakly coupled to the environment. For the case of the atoms
135: confined into the region much smaller than the optical wavelength
136: (Dicke model), the antisymmetric state is completely decoupled from
137: the environment, and therefore can be regarded as a decoherence-free
138: state.
139:
140: Another particularly interesting entangled states of a two-atom
141: system are two-photon entangled states that are superpositions of
142: only those states of the two-atom system in which both or neither of
143: the atoms is excited. These states have been known for a long time as
144: pairwise atomic states or multi-atom squeezed
145: states~\cite{mas,pas1,pas2,pas3,pas4,pas5}.
146: The two-photon entangled states cannot be generated
147: by a coherent laser field coupled to the atomic dipole moments. The
148: states can be created by a two-photon excitation process with
149: nonclassical correlations that can transfer the population from the
150: two-atom ground state to the upper state without populating the
151: intermediate one-photon states. An obvious candidate for the creation
152: of the two-photon entangled states is a broadband squeezed vacuum
153: field which is characterised by strong nonclassical two-photon
154: correlations~\cite{park,fd1,fd2}.
155:
156: One of the fundamental interests in collective atomic effects is to
157: demonstrate creation of entanglement on systems containing only two atoms.
158: A significant body of work on preparation of a two-atom system in an
159: entangled state has accumulated, and two-atom entangled states have
160: already been demonstrated experimentally using ultra cold trapped
161: ions in free space~\cite{deb,tur} and cavity quantum electrodynamics
162: (QED) schemes~\cite{hag,rbh}. In the free space situation, the collective
163: effects arise from the interaction between the atoms through the vacuum
164: field that the electromagnetic field produced by one of the atoms
165: influences the dipole moment of the another atom. This leads to an
166: additional damping and a shift of the atomic levels that both depend
167: on the interatomic separation. In the cavity QED
168: scheme, the atoms interact through the cavity mode and in a good cavity
169: limit, photons emitted by one of the atoms are almost immediately
170: absorbed by the another atom. In this case, the system behaves like
171: the Dicke model. Moreover, the strong coupling of the atoms to
172: the cavity mode prevents the atoms to emit photons to the vacuum
173: modes different from the cavity mode that reduces decoherence.
174:
175: Recently, the preparation of correlated superposition
176: states in multi-atom system has been performed using a quantum nondemolition
177: (QND) measurement technique~\cite{kmb}. Osnaghi {\it et al.}~\cite{osna}
178: have demonstrated coherent control of two Rydberg
179: atoms in a non-resonant cavity environment. By adjusting the atom-cavity
180: detuning, the final entangled state could be controlled, opening the door
181: to complex entanglement manipulations~\cite{ymbrh}. Several proposals have
182: also been made for entangling atoms trapped in distant
183: cavities~\cite{mb1,mb2,mb3,mb4,mb5,prsg}, or in a Bose-Einstein
184: condensate~\cite{fg1,fg2}. In a very important experiment,
185: Schlosser {\it et al.}~\cite{schl} succeeded in confining single atoms in
186: microscopic traps, thus enhancing the possibility of further progress in
187: entanglement and quantum engineering.
188:
189: This review is concerned primarily with two-atom systems, since
190: it is generally believed that entanglement of only two microscopic
191: quantum systems (two qubits) is essential to implement quantum
192: protocols such as quantum computation.
193: Some description of the theoretical tools required for prediction of
194: entanglement in atomic systems is appropriate. Thus, we propose to begin
195: the review with an overview of the mathematical apparatus necessary
196: for describing the interaction of atoms with the electromagnetic field.
197: We will present the master equation technique and, in addition, we also
198: describe a more general formalism based on the
199: quantum jump approach. We review theoretical and experimental schemes
200: proposed for the preparation of two two-level atoms in an entangled state.
201: We will also relate the atomic entanglement to nonclassical effects
202: such as photon antibunching, squeezing and sub-Poissonian photon
203: statistics. In particular, we consider different schemes of generation
204: of entangled and nonclassical states of two identical as well as
205: nonidentical atoms. The cases of maximally and non-maximally entangled
206: states will be considered and methods of detecting of particular entangled
207: and nonclassical state of two-atom systems are discussed. Next, we will
208: examine methods of preparation of a two-atom system in two-photon
209: entangled states. Finally, we will discuss methods of mapping of the
210: entanglement of light on atoms involving collective atomic
211: interactions and squeezing of the atomic dipole fluctuations.
212:
213:
214: \section{Time evolution of a collective atomic system}\label{ftsec2}
215:
216:
217: The standard formalism for the calculations of the time evolution and
218: correlation properties of a collective system of atoms is the master
219: equation method. In this approach, the dynamics are studied in terms of
220: the reduced density operator $\hat{\rho}_{A}$ of the atomic system
221: interacting with the quantized
222: electromagnetic (EM) field regarded as a reservoir~\cite{leh,lui,ag74}.
223: There are many possible realizations of reservoirs. The typical reservoir
224: to which atomic systems are coupled is the quantized three-dimensional
225: multimode field. The reservoir can be modelled as a vacuum field whose
226: the modes are in ordinary vacuum states, or in thermal states, or even
227: in squeezed vacuum states. The major advantage of the master equation is
228: that it allows us to consider the evolution of the atoms plus field
229: system entirely in terms of average values of atomic operators. We can
230: derive equations of motion for expectation values of an arbitrary
231: combination of the atomic operators, and solve these equations for
232: time-dependent averages or the steady-state. Another method is the
233: quantum jump approach. This is based on the theory of quantum
234: trajectories~\cite{car93}, which is equivalent to the Monte Carlo
235: wave-function approach~\cite{dcm92,pk98}, and allows to predict all
236: possible trajectories of a single quantum system which stochastically
237: emits photons. Both methods, the master equation and quantum jumps
238: approaches lead to the same final results of the dynamics of an atomic
239: system, and are widely used in quantum optics.
240:
241:
242: \subsection{Master equation approach}\label{ftsec21}
243:
244:
245: We first give an outline of the derivation of the master equation of
246: a system of $N$ non-identical nonoverlapping atoms coupled to the
247: quantized three-dimensional EM field. This derivation is a generalisation
248: of the master equation technique, introduced by Lehmberg~\cite{leh}, to
249: the case of non-identical atoms interacting with a squeezed vacuum field.
250: Useful references on the derivation of
251: the master equation of an atomic system coupled to an ordinary vacuum
252: are the books of Louisell~\cite{lui} and Agarwal~\cite{ag74}.
253: The atoms are modelled as two-level systems, with excited state
254: $\ket {e_{i}}$, ground state $\ket {g_{i}}$, transition frequency
255: $\omega_{i}$, and transition dipole moments $\vec{\mu}_{i}$. We assume
256: that the atoms are located at different points $\vec{r}_{1},\dots
257: \vec{r}_{N}$, have different transition frequencies $\omega_{1}\neq
258: \omega_{2}\neq \dots \neq \omega_{N}$, and different transition dipole
259: moments $\vec{\mu}_{1}\neq \vec{\mu}_{2}\neq \dots \neq \vec{\mu}_{N}$.
260:
261: In the electric dipole approximation, the total Hamiltonian of the
262: combined system, the atoms plus the EM field, is given by
263: \begin{eqnarray}
264: \hat{H} &=& \sum_{i=1}^{N}\hbar \omega_{i}S_{i}^{z} +
265: \sum_{\vec{k}s} \hbar \omega_{k}\left(\hat {a}_{\vec{k}s}^{\dagger}
266: \hat {a}_{\vec{k}s} +\frac{1}{2}\right) \nonumber \\
267: &-& i\hbar\sum_{\vec{k}s}\sum_{i=1}^{N}\left[ \vec{\mu}_{i}\cdot
268: \vec{g}_{\vec{k}s}\left(\vec{r}_{i}\right) \left(
269: S_{i}^{+}+S_{i}^{-}\right)\hat{a}_{\vec{k}s}
270: -{\rm H.c.}\right] \ ,\label{t1}
271: \end{eqnarray}
272: where $S_{i}^{+} = \ket {e_{i}}\bra {g_{i}}$ and $S_{i}^{-} =
273: \ket {g_{i}}\bra {e_{i}}$ are the dipole raising and lowering
274: operators, $S_{i}^{z} =\left(\ket {e_{i}}
275: \bra {e_{i}} -\ket {g_{i}}\bra {g_{i}}\right)/2$ is the energy operator
276: of the $i$th atom, $\hat {a}_{\vec{k}s}$ and
277: $\hat {a}_{\vec{k}s}^{\dagger}$ are the annihilation and creation
278: operators of the field mode~$\vec{ks}$, which has wave vector $\vec{k}$,
279: frequency $\omega_{k}$ and the index of polarization~$s$. The coupling
280: constant
281: \begin{equation}
282: \vec{g}_{\vec{k}s}\left( \vec{r}_{i}\right) =\left( \frac{\omega_{k}}{2
283: \epsilon_{0}\hbar V}\right) ^{\frac{1}{2}}
284: \bar{e}_{\vec{k}s}e^{i\vec{k}\cdot \vec{r}_{i}} \ , \label{t2}
285: \end{equation}
286: is the mode function of the three-dimensional vacuum field,
287: evaluated at the position $\vec{r}_{i}$ of the $i$th atom,
288: $V$ is the normalization volume, and $\bar{e}_{\vec{k}s}$ is the unit
289: polarization vector of the field.
290:
291: The atomic dipole operators, appearing in Eq.~(\ref{t1}), satisfy
292: the well-known commutation and anticommutation relations
293: \begin{equation}
294: \left[S_{i}^{+}, S_{j}^{-}\right] =2S_{i}^{z}\delta_{ij} \ ,\quad
295: \left[S_{i}^{z}, S_{j}^{\pm}\right] =\pm S_{i}^{\pm}\delta_{ij} \
296: ,\quad \left[S_{i}^{+}, S_{j}^{-}\right]_{+} =\delta_{ij}
297: \ ,\label{t3}
298: \end{equation}
299: with $\left(S_{i}^{\pm}\right)^{2}\equiv 0$.
300:
301: While this is straightforward, it is often the case that it is simpler
302: to work in the interaction picture in which the
303: Hamiltonian~(\ref{t1}) evolves in time according to the interaction
304: with the vacuum field. Therefore, we write the total Hamiltonian~(\ref{t1})
305: as
306: \begin{eqnarray}
307: \hat{H} &=& \hat{H}_{0} +\hat{H}_{I} \ , \label{t4}
308: \end{eqnarray}
309: where
310: \begin{eqnarray}
311: \hat{H}_{0} &=& \sum_{i=1}^{N}\hbar \omega_{i}S_{i}^{z}
312: +\sum_{\vec{k}s} \hbar \omega_{k}\left(\hat {a}_{\vec{k}s}^{\dagger}
313: \hat {a}_{\vec{k}s} +\frac{1}{2}\right) \ ,\label{t5}
314: \end{eqnarray}
315: is the Hamiltonian of the non-interacting atoms and the EM field, and
316: \begin{eqnarray}
317: \hat{H}_{I} &=&
318: -i\hbar\sum_{\vec{k}s}\sum_{i=1}^{N}\left[ \vec{\mu}_{i}\cdot
319: \vec{g}_{\vec{k}s}\left(\vec{r}_{i}\right)
320: \left(S_{i}^{+}+S_{i}^{-}\right)\hat{a}_{\vec{k}s}
321: -{\rm H.c.}\right] \ ,\label{t6}
322: \end{eqnarray}
323: is the interaction Hamiltonian between the atoms and the EM field.
324:
325: The Hamiltonian $\hat{H}_{0}$ transforms the total Hamiltonian~(\ref{t1}) into
326: \begin{eqnarray}
327: \hat{H}\left(t\right) &=& e^{i\hat{H}_{0}t/\hbar}
328: \left(\hat{H}-\hat{H}_{0}\right)e^{-i\hat{H}_{0}t/\hbar} =
329: \hat{V}\left(t\right) \ ,\label{t7}
330: \end{eqnarray}
331: where
332: \begin{eqnarray}
333: \hat{V}\left(t\right) &=& -i\hbar\sum_{\vec{k}s}\sum_{i=1}^{N}
334: \left\{ \vec{\mu}_{i}\cdot
335: \vec{g}_{\vec{k}s}\left(\vec{r}_{i}\right)
336: S_{i}^{+}\hat{a}_{\vec{k}s}e^{-i\left(\omega_{k}-\omega_{i}\right)t}\right.
337: \nonumber \\
338: &&\left. +\vec{\mu}_{i}\cdot
339: \vec{g}_{\vec{k}s}\left(\vec{r}_{i}\right)S_{i}^{-}\hat{a}_{\vec{k}s}
340: e^{-i\left(\omega_{k}+\omega_{i}\right)t}
341: -{\rm H.c.}\right\} \ .\label{t8}
342: \end{eqnarray}
343:
344: We will consider the time evolution of the collection of atoms interacting
345: with the vacuum field in terms of the density operator $\hat{\rho}_{AF}$
346: characterizing the statistical state of the combined system of the
347: atoms and the vacuum field. The time evolution of the density operator
348: of the combined system obeys the equation
349: \begin{equation}
350: \frac{\partial}{\partial t}\hat{\rho}_{AF} =
351: \frac{1}{i\hbar}\left[\hat{H},\hat{\rho}_{AF}\right] \
352: .\label{t9}
353: \end{equation}
354:
355: Transforming Eq.~(\ref{t9}) into the interaction picture with
356: \begin{equation}
357: \tilde{\hat{\rho}}_{AF}\left(t\right) = e^{i\hat{H}_{0}t/\hbar}
358: \hat{\rho}_{AF}e^{-i\hat{H}_{0}t/\hbar} \ ,\label{t10}
359: \end{equation}
360: we find that the transformed density operator satisfies the equation
361: \begin{equation}
362: \frac{\partial}{\partial t}\tilde{\hat{\rho}}_{AF}\left(t\right) =
363: \frac{1}{i\hbar}\left[ \hat{V}\left(t\right),
364: \tilde{\hat{\rho}}_{AF}\left(t\right)\right] \ , \label{t11}
365: \end{equation}
366: where the interaction Hamiltonian $\hat{V}\left(t\right)$ is given
367: in Eq.~(\ref{t8}).
368:
369: Equation~(\ref{t11}) is a simple differential equation which can be
370: solved by the iteration method. For the initial time $t=0$, the
371: integration of Eq.~(\ref{t11}) leads to the following first-order
372: solution in $\hat{V}\left(t\right)$:
373: \begin{eqnarray}
374: \tilde{\hat{\rho}}_{AF}\left(t\right) =
375: \tilde{\hat{\rho}}_{AF}\left(0\right)
376: + \frac{1}{i\hbar}\int_{0}^{t}dt^{\prime}\left[
377: \hat{V}\left(t^{\prime}\right),
378: \tilde{\hat{\rho}}_{AF}\left(t^{\prime}\right)\right] \ . \label{t12}
379: \end{eqnarray}
380: Substituting Eq.~(\ref{t12}) into the right side of Eq.~(\ref{t11})
381: and taking the trace over the vacuum field variables, we find that to
382: the second order in $\hat{V}\left(t\right)$ the reduced density
383: operator of the atomic system $\hat{\rho}_{A}\left(t\right)={\rm
384: Tr}_{F}\tilde{\hat{\rho}}_{AF}\left(t\right)$ satisfies the
385: integro-differential equation
386: \begin{eqnarray}
387: \frac{\partial}{\partial t}\hat{\rho}_{A}\left(t\right) &=&
388: \frac{1}{i\hbar}{\rm Tr}_{F}\left[\hat{V}\left(t\right),
389: \tilde{\hat{\rho}}_{AF}\left(0\right)\right] \nonumber \\
390: &-& \frac{1}{\hbar^{2}}\int_{0}^{t} dt^{\prime} {\rm Tr}_{F}\left\{\left[
391: \hat{V}\left(t\right),\left[\hat{V}\left(t^{\prime} \right),
392: \tilde{\hat{\rho}}_{AF}\left(t^{\prime} \right)\right]\right]\right\}
393: \ .\label{t13}
394: \end{eqnarray}
395: We choose an initial state with no correlations between the atomic
396: system and the vacuum field, which allows us to factorize the initial
397: density operator of the combined system as
398: \begin{eqnarray}
399: \tilde{\hat{\rho}}_{AF}\left(0\right)
400: =\hat{\rho}_{A}\left(0\right)\hat{\rho}_{F}\left(0\right) \
401: ,\label{t14}
402: \end{eqnarray}
403: where $\hat{\rho}_{F}$ is the density operator of the vacuum field.
404:
405: We now employ the Born approximation~\cite{lui}, in which the interaction
406: between the atomic system and the field is supposed to be weak, and there
407: is no the back reaction effect of the atoms on the field. In this
408: approximation the state of the vacuum field does not change in time,
409: and we can write the density operator
410: $\tilde{\hat{\rho}}_{AF}\left(t^{\prime}\right)$, appearing in
411: Eq.~(\ref{t13}), as
412: \begin{eqnarray}
413: \tilde{\hat{\rho}}_{AF}\left(t^{\prime}\right)
414: =\hat{\rho}_{A}\left(t^{\prime}\right)\hat{\rho}_{F}\left(0\right)
415: \ .\label{t15}
416: \end{eqnarray}
417: Under this approximation, and after changing the time variable to
418: $t^{\prime}=t -\tau$, Eq.~(\ref{t13}) simplifies to
419: \begin{eqnarray}
420: \frac{\partial}{\partial t}\hat{\rho}\left(t\right) &=&
421: \frac{1}{i\hbar}{\rm Tr}_{F}\left[\hat{V}\left(t\right),
422: \hat{\rho}\left(0\right)\hat{\rho}_{F}\left(0\right)\right] \nonumber \\
423: &-& \frac{1}{\hbar^{2}}\int_{0}^{t} d\tau {\rm Tr}_{F}\left\{\left[
424: \hat{V}\left(t\right),\left[\hat{V}\left(t -\tau \right),
425: \hat{\rho}\left(t-\tau \right)\hat{\rho}_{F}\left(0\right)
426: \right]\right]\right\}
427: \ ,\label{t16}
428: \end{eqnarray}
429: where we use a shorter notation $\hat{\rho}= \hat{\rho}_{A}$.
430:
431: Substituting the explicit form of
432: $\hat{V}\left(t\right)$ into Eq.~(\ref{t16}), we find that the
433: evolution of the density operator depends on the first and second
434: order correlation functions of the vacuum field operators. We assume
435: that a part of the vacuum modes is in a squeezed vacuum state for
436: which the correlation functions are given by~\cite{park,fd1,fd2}
437: \begin{eqnarray}
438: {\rm Tr}_{F}\left[\hat{\rho}_{F}\left(0\right){\hat a}_{\vec{k}s}
439: \right] &=& {\rm Tr}_{F}\left[\hat{\rho}_{F}\left(0\right)
440: {\hat a}^{\dagger}_{\vec{k}s}\right] = 0 \ , \nonumber \\
441: {\rm Tr}_{F}\left[\hat{\rho}_{F}\left(0\right){\hat a}_{\vec{k}s}
442: {\hat a}^{\dagger}_{\vec{k}^{\prime}s^{\prime}}\right] &=&
443: \left[\left|D\left(\omega_{k}\right)\right|^{2}
444: N\left(\omega_{k}\right)+1\right]
445: \delta^{3}\left(\vec{k}-\vec{k^{\prime}}\right)\delta_{ss^{\prime}}
446: \ ,\nonumber \\
447: {\rm Tr}_{F}\left[\hat{\rho}_{F}\left(0\right){\hat a}^{\dagger}_{\vec{k}s}
448: {\hat a}_{\vec{k}^{\prime}s^{\prime}}\right] &=&
449: \left|D\left(\omega_{k}\right)\right|^{2} N\left(\omega_{k}\right)
450: \delta^{3}\left(\vec{k}-\vec{k^{\prime}}\right)\delta_{ss^{\prime}}
451: \ ,\nonumber \\
452: {\rm Tr}_{F}\left[\hat{\rho}_{F}\left(0\right){\hat a}_{\vec{k}s}
453: {\hat a}_{\vec{k}^{\prime}s^{\prime}}\right] &=&
454: D^{2}\left(\omega_{k}\right) M\left(\omega_{k}\right)
455: \delta^{3}\left(2\vec{k}_{s} -\vec{k}-\vec{k}^{\prime}\right)
456: \delta_{ss^{\prime}} \ ,\nonumber \\
457: {\rm Tr}_{F}\left[\hat{\rho}_{F}\left(0\right){\hat a}^{\dagger}_{\vec{k}s}
458: {\hat a}^{\dagger}_{\vec{k}^{\prime}s^{\prime}}\right] &=&
459: D^{2}\left(\omega_{k}\right) M^{\ast}\left(\omega_{k}\right)
460: \delta^{3}\left(2\vec{k}_{s} -\vec{k}-\vec{k}^{\prime}\right)
461: \delta_{ss^{\prime}} \ ,\label{t17}
462: \end{eqnarray}
463: where the parameters $N\left(\omega_{k}\right)$ and
464: $M\left(\omega_{k}\right)$ characterize squeezing in the vacuum field,
465: such that $N\left(\omega_{k}\right)$ is the number of photons in the
466: mode $\vec{k}$, $M\left(\omega_{k}\right)=
467: |M\left(\omega_{k}\right)|{\rm exp}(i\phi_{s})$ is the magnitude of
468: two-photon correlations between the vacuum modes, and
469: $\phi_{s}$ is the phase of the squeezed field. The two-photon
470: correlations are symmetric about the squeezing carrier frequency
471: $2\omega_{s}$, i.e.
472: $M\left(\omega_{k}\right) = M\left(2\omega_{s}-\omega_{k}\right)$, and
473: are related by the inequality
474: \begin{equation}
475: \left|M\left(\omega_{k}\right)\right|^{2} \leq
476: N\left(\omega_{k}\right)\left(N\left(2\omega_{s}-\omega_{k}\right)
477: +1\right) ,\label{t18}
478: \end{equation}
479: where the term $+1$ on the right-hand side arises from the quantum
480: nature of the squeezed field~\cite{fd1,fd2}. Such a field is often
481: called a quantum squeezed field. For a classical analogue
482: of squeezed field the two-photon correlations are given by the
483: inequality $\left|M\left(\omega_{k}\right)\right| \leq
484: N\left(\omega_{k}\right)$. Thus, two-photon correlations with
485: $0<\left|M\left(\omega_{k}\right)\right| \leq
486: N\left(\omega_{k}\right)$ may be generated by a classical field,
487: whereas correlations with $N\left(\omega_{k}\right)<
488: \left|M\left(\omega_{k}\right)\right| \leq
489: \sqrt{N\left(\omega_{k}\right)\left(N\left(2\omega_{s}
490: -\omega_{k}\right)+1\right)}$ can only be generated by a quantum
491: field which has no classical analog.
492:
493: The parameter $D\left(\omega_{k}\right)$, appearing in Eq.~(\ref{t17}),
494: determines the matching of the squeezed modes to the three-dimensional
495: vacuum modes surrounding the atoms, and contains both the amplitude and
496: phase coupling. The explicit form of $D\left(\omega_{k}\right)$ depends
497: on the method of propagation and focusing the squeezed
498: field~\cite{pas5,pg89}. For perfect matching,
499: $\left| D\left(\omega_{k}\right)\right|^{2} =1$, whereas $\left|
500: D\left( \omega_{k}\right)\right|^{2} <1$ for an imperfect matching.
501: The perfect matching is an idealization as it is practically impossible
502: to achieve perfect matching in present experiments~\cite{kimb1,kimb2}.
503: In order to avoid the experimental difficulties, cavity situations
504: have been suggested. In this case, the parameter
505: $D\left(\omega_{k}\right)$ is identified as the cavity transfer function,
506: the absolute value square of which is the Airy function of the
507: cavity~\cite{bw,yar}. The function $|D\left(\omega_{k}\right)|^{2}$
508: exhibits a sharp peak centred at the cavity axis and all
509: the cavity modes are contained in a small solid angle around this
510: central mode. By squeezing of these modes we can achieve perfect
511: matching between the squeezed field and the atoms. In a realistic
512: experimental situation the input squeezed modes have a Gaussian profile
513: for which the parameter $D\left(\omega_{k}\right)$ is given by
514: \cite{yar,fde,fd91}
515: \begin{equation}
516: D\left(\omega_{k}\right) =\exp\left[- W_{0}\sin^{2}\theta_{k}-ikz_{f}\cos
517: \theta _{k}\right] \ ,\label{t19}
518: \end{equation}
519: where $\theta_{k}$ is an angle over which the squeezed mode $\vec{k}$
520: is propagated, and $W_{0} $ is the beam spot size at the focal point
521: $z_{f}$. Thus, even in the cavity situation, perfect matching could be
522: difficult to achieve in present experiments.
523:
524: Before returning to the derivation of the master equation, we should
525: remark that in realistic experimental situations, the squeezed modes
526: cover only a small portion of the modes surrounding the atoms. The
527: squeezing modes lie inside a cone of angle $\theta_{k}<\pi $, and the
528: modes outside the cone are in their ordinary vacuum state. In fact, the
529: modes are in a finite temperature black-body state, which means that
530: inside the cone the modes are in mixed squeezed vacuum and black-body
531: states. However, this is not a serious practical problem as experiments are
532: usually performed at low temperatures where the black-body radiation is
533: negligible. In principle, we can include the black-body radiation effect
534: (thermal noise) to the problem replacing
535: $\left|D\left(\omega_{k}\right)\right|^{2} N\left( \omega_{k}\right)$
536: in Eq.~(\ref{t17})
537: by $\left|D\left(\omega_{k}\right)\right|^{2} N\left( \omega_{k}\right)
538: +\bar{N},$ where $\bar{N}$ is proportional to the photon number in
539: the black-body radiation.
540:
541: We now return to the derivation of the master equation for the density
542: operator of the atomic system coupled to a squeezed vacuum field.
543: First, we change the sum over $\vec{k}s$ into an integral
544: \begin{eqnarray}
545: \sum_{\vec{k}s}\longrightarrow \frac{V}{\left(2\pi
546: c\right)^{3}}\sum_{s=1}^{2}\int_{0}^{\infty}d\omega_{k}\omega_{k}^{2}
547: \int d\Omega_{k} \ .\label{t20}
548: \end{eqnarray}
549: Next, with the correlation functions~(\ref{t17}) and after the
550: rotating-wave approximation (RWA)~\cite{ae}, in which we ignore all
551: terms oscillating at higher frequencies,
552: $2\omega_{i}, \omega_{i}+\omega_{j}$, the general master
553: equation~(\ref{t16}) can be written as
554: \begin{eqnarray}
555: \frac{\partial}{\partial t} \hat{\rho}\left( t\right) =
556: &\sum_{i,j=1}^{N}& \left\{
557: \left[S_{j}^{-}\hat{X}_{ij}\left( t,\tau \right),S_{i}^{+}\right]
558: +\left[S_{j}^{-},\hat{X}_{ji}^{\dagger}\left( t,\tau
559: \right)S_{i}^{+}\right]\right. \nonumber \\
560: &&+ \left. \left[S_{j}^{+}\hat{Y}_{ij}\left( t,\tau
561: \right),S_{i}^{-}\right]
562: +\left[S_{j}^{+},\hat{Y}_{ji}^{\dagger}\left( t,\tau
563: \right)S_{i}^{-}\right]\right. \nonumber \\
564: &&+ \left. \left[S_{i}^{+}\hat{K}_{ij}\left( t,\tau
565: \right),S_{j}^{+}\right]
566: +\left[S_{i}^{+},\hat{K}_{ij}\left( t,\tau
567: \right)S_{j}^{+}\right]\right. \nonumber \\
568: &&+ \left. \left[S_{i}^{-}\hat{K}^{\dagger}_{ij}\left( t,\tau
569: \right),S_{j}^{-}\right]
570: +\left[S_{i}^{-},\hat{K}^{\dagger}_{ij}\left( t,\tau
571: \right)S_{j}^{-}\right]\right\} \ ,\label{t21}
572: \end{eqnarray}
573: where the two-time operators are
574: \begin{eqnarray}
575: \hat{X}_{ij}\left( t,\tau \right) &=& \frac{V}{(2\pi c)^{3}}\int
576: d\omega_{k}\omega_{k}^{2} e^{-i\left( \omega _{i}-\omega_{j}\right) t}
577: \int d\Omega _{k}\sum_{s=1}^{2}\chi_{ij}^{(-)}\left( t,\tau \right)
578: \ , \nonumber \\
579: \hat{Y}_{ij}\left( t,\tau \right) &=&\frac{V}{(2\pi c)^{3}}
580: \int d\omega_{k}\omega_{k}^{2}e^{i\left( \omega _{i}-\omega_{j}\right) t}
581: \int d\Omega _{k}\sum_{s=1}^{2}\chi_{ij}^{(+)}\left( t,\tau \right)
582: \ , \label{t22} \\
583: \hat{K}_{ij}\left( t,\tau \right) &=&\frac{V}{(2\pi c)^{3}}\int
584: d\omega_{k}\omega_{k}\left( 2\omega_{s}-\omega_{k}\right)
585: e^{-i\left(2\omega_{s}- \omega _{i}-\omega_{j}\right) t} \nonumber \\
586: &\times&
587: \int_{\Omega_{s}} d\Omega _{k}\sum_{s=1}^{2}\chi_{ij}^{(M)}
588: \left( t,\tau \right) \ ,\nonumber
589: \end{eqnarray}
590: with
591: \begin{eqnarray}
592: \chi_{ij}^{(\pm)}\left( t,\tau \right) &=&
593: \left[\left|D\left(\omega_{k}\right)\right|^{2}
594: N\left(\omega_{k}\right)+1\right]
595: \left[\vec{\mu}_{i}\cdot \vec{g}_{\vec{k}s}\left(\vec{r}_{i}\right) \right]
596: \left[\vec{\mu}_{j}^{\ast}\cdot \vec{g}_{\vec{k}s}^{\ast}
597: \left(\vec{r}_{j}\right) \right] \nonumber \\
598: &\times& \int_{0}^{t}d\tau \hat{\rho}\left(t-\tau
599: \right)e^{-i\left(\omega_{k}\pm \omega_{j}\right)\tau} \nonumber \\
600: &+&\left|D\left(\omega_{k}\right)\right|^{2} N\left(\omega_{k}\right)
601: \left[\vec{\mu}_{i}^{\ast}\cdot
602: \vec{g}_{\vec{k}s}^{\ast}\left(\vec{r}_{i}\right) \right]
603: \left[\vec{\mu}_{j}\cdot \vec{g}_{\vec{k}s}
604: \left(\vec{r}_{j}\right) \right] \nonumber \\
605: &\times& \int_{0}^{t}d\tau \hat{\rho}\left(t-\tau
606: \right)e^{i\left(\omega_{k}\mp \omega_{j}\right)\tau} \ ,\nonumber \\
607: \chi_{ij}^{\left( M\right) }\left(t\right) &=& M\left( \omega_{k}\right)
608: D^{2}\left(\omega_{k}\right)
609: \left[\vec{\mu}_{i}\cdot \vec{g}_{\vec{k}s}\left(
610: \vec{r}_{i}\right) \right] \left[\vec{\mu}_{j}\cdot
611: \vec{g}_{\vec{k}s}\left(\vec{r}_{j}\right) \right] \nonumber \\
612: &\times& \int_{0}^{t}d\tau \hat{\rho}\left(t-\tau
613: \right)e^{i\left(2\omega_{s}-\omega_{k}- \omega_{j}\right)\tau} \
614: ,\label{t23}
615: \end{eqnarray}
616: and $\Omega_{s}$ is the solid angle over which the squeezed vacuum
617: field is propagated.
618:
619: The master equation (\ref{t21}) with parameters (\ref{t22}) and
620: (\ref{t23}) is
621: quite general in terms of the matching of the squeezed modes to the vacuum
622: modes and the bandwidth of the squeezed field relative to the atomic
623: linewidths. The master equation is in the form of an integro-differential
624: equation, and can be simplified by employing the Markov
625: approximation~\cite{lui}.
626: In this approximation the integral over the time delay~$\tau $ contains
627: functions which decay to zero over a short correlation time~$\tau_{c}.$
628: This correlation time is of the order of the inverse bandwidth of the
629: squeezed field, and the short correlation time approximation is formally
630: equivalent to assume that squeezing bandwidths are much larger
631: than the atomic linewidths. Over this short time-scale the density operator
632: would
633: hardly have changed from $\hat{\rho} \left( t\right) ,$ thus we can replace
634: $\hat{\rho} \left( t-\tau \right)$ by $\hat{\rho }\left( t\right) $ in
635: Eq.~(\ref{t23}) and extend the integral to infinity. Under these
636: conditions, we can perform the integration over $\tau$ and
637: obtain~\cite{ae}
638: \begin{equation}
639: \lim_{t \to\infty}\int_{0}^{t}d\tau \hat{\rho} \left( t-\tau \right)
640: e^{i x \tau } \approx \hat{\rho}
641: \left( t\right) \left[ \pi \delta \left( x\right)
642: +i\frac{{\cal P}}{x}\right] \ ,\label{t24}
643: \end{equation}
644: where ${\cal P}$ indicates the principal value of the integral. Moreover,
645: for squeezing bandwidth much larger than the atomic linewidths, we
646: can approximate the squeezing parameters and the mode function
647: evaluated at $\omega_{k}$ by their maximal values evaluated at
648: $\omega_{s}$, i.e., we can take
649: $N\left(\omega_{k}\right) =N\left(\omega_{s}\right)$,
650: $M\left( \omega_{k}\right)= M\left( \omega_{s}\right)$, and
651: $D\left(\omega_{k}\right)= D\left(\omega_{s}\right)$.
652:
653: Finally, to carry out the polarization sums and integrals over
654: $d\Omega _{k}$ in Eq.~(\ref{t22}), we assume that the dipole moments
655: of the atoms are parallel and use the spherical representation for the
656: propagation vector $\vec{k}$. The integral over $d\Omega_{k}$ contains
657: integrals over the spherical angular coordinates $\theta$ and $\phi$.
658: The angle $\theta$ is formed by $\vec{r}_{ij}$ and $\vec{k}$ directions,
659: so we can write
660: \begin{eqnarray}
661: \vec{k} = \left|\vec{k}\right|\left[\sin\theta \cos \phi
662: \ ,\sin\theta \sin \phi \ ,\cos\theta \right] \ .\label{t25}
663: \end{eqnarray}
664: In this representation, the unit polarization vectors $\bar{e}_{\vec{k}1}$
665: and $\bar{e}_{\vec{k}2}$ may be chosen as~\cite{lui}
666: \begin{eqnarray}
667: \bar{e}_{\vec{k}1} &=& \left[ -\cos\theta \cos\phi \ ,-\cos\theta
668: \sin\phi \ ,\sin\theta \right] \ , \nonumber \\
669: \bar{e}_{\vec{k}2} &=& \left[ \sin\phi \ ,-\cos\phi \ , 0 \right] \ ,
670: \label{t26}
671: \end{eqnarray}
672: and the orientation of the atomic dipole moments can be taken in
673: the $x$ direction
674: \begin{eqnarray}
675: \vec{\mu}_{i} &=& \left|\vec{\mu}_{i}\right|\left[1 \ ,0 \ ,0\right] \
676: ,\nonumber \\
677: \vec{\mu}_{j} &=&
678: \left|\vec{\mu}_{j}\right|\left[1 \ ,0 \ ,0\right] \ .\label{t27}
679: \end{eqnarray}
680: With this choice of the polarization vectors and the orientation of
681: the dipole moments, we obtain
682: \begin{eqnarray}
683: \hat{X}_{ij}\left( t,\tau \right) &=& \left\{
684: \left[ 1+\tilde{N}\left(\omega_{s}\right)\right]\left(\frac{1}{2}\Gamma_{ij}
685: -i\Omega_{ij}^{(-)}\right) +i\tilde{N}\left(\omega_{s}\right)
686: \Omega_{ij}^{(+)}\right\}
687: \hat{\rho} \left(t\right) e^{-i\left( \omega_{i}-\omega _{j}\right) t}
688: \ ,\nonumber \\
689: \hat{Y}_{ij}\left( t,\tau \right) &=& \left\{
690: \tilde{N}\left(\omega_{s}\right)\left(\frac{1}{2}\Gamma_{ij}
691: +i\Omega_{ij}^{(-)}\right) -i\left[ 1+\tilde{N}\left(\omega_{s}
692: \right)\right]\Omega_{ij}^{(+)}\right\}
693: \hat{\rho} \left(t\right) e^{i\left( \omega_{i}-\omega _{j}\right) t}
694: \ ,\nonumber \\
695: \hat{K}_{ij}\left( t,\tau \right) &=& \tilde{M}\left(\omega_{s}\right)
696: \left(\frac{1}{2}\Gamma_{ij} +i\Omega_{ij}^{(M)}\right)
697: \hat{\rho} \left( t\right) e^{-i\left( 2\omega
698: _{s}-\omega _{i}-\omega _{j}\right) t} \ , \label{t28}
699: \end{eqnarray}
700: where
701: \begin{eqnarray}
702: \tilde{N}\left(\omega_{s}\right) &=& N\left(\omega_{s}\right)
703: \left|D\left(\omega_{s}\right)\right|^{2} v\left(\theta
704: _{s}\right) \ ,\nonumber \\
705: \tilde{M}\left(\omega_{s}\right) &=& M\left(\omega_{s}\right)
706: \left|D\left(\omega_{s}\right)\right|^{2} v\left(\theta
707: _{s}\right) \ ,\label{t29}
708: \end{eqnarray}
709: with
710: \begin{eqnarray}
711: v\left( \theta _{s}\right) =\frac{1}{2}\left[ 1-\frac{1}{4}\left( 3+\cos
712: ^{2}\theta _{s}\right) \cos \theta _{s}\right] \ , \label{t30}
713: \end{eqnarray}
714: and $\theta _{s}$ is the angle over which the squeezed vacuum is
715: propagated.
716:
717: The parameters $\Gamma_{ij}$, which appear in Eq.~(\ref{t28}),
718: are spontaneous emission rates, such that
719: \begin{eqnarray}
720: \Gamma_{i}\equiv \Gamma_{ii} =
721: \frac{\omega_{i}^{3}\mu_{i}^{2}}{3\pi \varepsilon_{o}\hbar
722: c^{3}} \label{t31}
723: \end{eqnarray}
724: is the spontaneous emission rate of the $i$th atom, equal to the
725: Einstein $A$ coefficient for spontaneous emission, and
726: \begin{eqnarray}
727: \Gamma_{ij}=\Gamma_{ji}=
728: \sqrt{\Gamma_{i}\Gamma_{j}}F\left(k_{0}r_{ij}\right) \qquad (i\neq j)
729: \ ,\label{t32}
730: \end{eqnarray}
731: where
732: \begin{eqnarray}
733: F\left(k_{0}r_{ij}\right) &=&\frac{3}{2}
734: \left\{ \left[1 -\left( \bar{\mu}\cdot \bar{r}
735: _{ij}\right)^{2} \right] \frac{\sin \left( k_{0}r_{ij}\right)
736: }{k_{0}r_{ij}}\right. \nonumber \\
737: &&\left. +\left[ 1 -3\left( \bar{\mu}\cdot
738: \bar{r}_{ij}\right)^{2} \right] \left[ \frac{\cos \left(
739: k_{0}r_{ij}\right) }{\left( k_{0}r_{ij}\right) ^{2}}-\frac{\sin \left(
740: k_{0}r_{ij}\right) }{\left( k_{0}r_{ij}\right) ^{3}}\right] \right\} \ ,
741: \label{t33}
742: \end{eqnarray}
743: are collective spontaneous emission rates arising from the coupling
744: between the atoms through the vacuum
745: field~\cite{ftk87,leh,ag74,mk74,mk75}. In the
746: expression~(\ref{t33}), $\bar{\mu}=\bar{\mu}_{i}=\bar{\mu}_{j}$ and
747: $\bar{r}_{ij}$ are unit vectors along the atomic transition dipole
748: moments and the vector $\vec{r}_{ij}=\vec{r}_{j}-\vec{r}_{i}$,
749: respectively. Moreover, $k_{0}=\omega_{0}/c$, where
750: $\omega_{0}=(\omega_{i}+\omega_{j})/2$, and we have assumed that
751: $(\omega_{i}-\omega_{j})\ll \omega_{0}$.
752:
753: The remaining parameters $\Omega_{ij}^{(\pm)}$ and $\Omega_{ij}^{(M)}$,
754: that appear in Eq.~(\ref{t28}), will contribute to the shifts of the
755: atomic levels, and are given by
756: \begin{eqnarray}
757: \Omega_{ij}^{(\pm)} &=& P\frac{\sqrt{\Gamma _{i}\Gamma _{j}}} {2\pi
758: \omega_{0}^{3}}\int_{0}^{\infty}
759: \frac{\omega_{k}^{3}F\left(\omega_{k}r_{ij}/c\right)}
760: {\omega_{k}\pm \omega_{j}}d\omega_{k} \ ,\label{t34}
761: \end{eqnarray}
762: and
763: \begin{eqnarray}
764: \Omega_{ij}^{(M)} &=& P\frac{\sqrt{\Gamma _{i}\Gamma _{j}}} {2\pi
765: \omega_{0}^{3}}\int_{0}^{\infty}
766: \frac{\omega_{k}^{2}\left(2\omega_{s}-\omega_{k}\right)
767: F\left(\omega_{k}r_{0}/c\right)}
768: {2\omega_{s}-\omega_{k}- \omega_{j}}d\omega_{k} \ ,\label{t35}
769: \end{eqnarray}
770: where $F\left(\omega_{k}r_{0}/c\right)$ is given in Eq.~(\ref{t33}) with
771: $k_{0}$ replaced by $\omega_{k}/c$, and $r_{ij}$ replaced by
772: $r_{0}=r_{i}+r_{j}$.
773:
774: With the parameters (\ref{t28}), the master equation of the system of
775: non-identical atoms in a broadband squeezed vacuum, written in the
776: Schr\"{o}dinger picture, reads
777: \begin{eqnarray}
778: \frac{\partial \hat{\rho} }{\partial t} &=&
779: -\frac{1}{2}\sum_{i,j=1}^{N}\Gamma _{ij}
780: \left[ 1+\tilde{N}\left(\omega_{s}\right) \right] \left( \hat{\rho}
781: S_{i}^{+}S_{j}^{-}+S_{i}^{+}S_{j}^{-}\hat{\rho}
782: -2S_{j}^{-}\hat{\rho} S_{i}^{+}\right) \nonumber \\
783: &&-\frac{1}{2}\sum_{i,j=1}^{N}\Gamma _{ij}\tilde{N}\left(\omega_{s}\right)
784: \left( \hat{\rho} S_{i}^{-}S_{j}^{+}+S_{i}^{-}S_{j}^{+}\hat{\rho}
785: -2S_{j}^{+}\hat{\rho} S_{i}^{-}\right) \nonumber \\
786: &&+\frac{1}{2}\sum_{i,j=1}^{N}\left(\Gamma _{ij}
787: +i\Omega_{ij}^{(M)}\right)\tilde{M}\left( \omega_{s}\right)
788: \left( \hat{\rho} S_{i}^{+}S_{j}^{+}+S_{i}^{+}S_{j}^{+}\hat{\rho}
789: -2S_{j}^{+}\hat{\rho} S_{i}^{+}\right) \nonumber \\
790: &&+\frac{1}{2}\sum_{i,j=1}^{N}\left(\Gamma _{ij}-i\Omega_{ij}^{(M)}\right)
791: \tilde{M}^{\ast }\left(\omega_{s}\right)
792: \left( \hat{\rho} S_{i}^{-}S_{j}^{-}+S_{i}^{-}S_{j}^{-}\hat{\rho}
793: -2S_{j}^{-}\hat{\rho} S_{i}^{-}\right) \nonumber \\
794: &&-i\sum_{i=1}^{N}\left(\omega_{i}+\delta_{i}\right)\left[
795: S^{z}_{i},\hat{\rho} \right]
796: -i\sum_{i\neq j}^{N}\Omega_{ij}\left[ S^{+}_{i}S^{-}_{j},
797: \hat{\rho} \right] \ , \label{t36}
798: \end{eqnarray}
799: where
800: \begin{eqnarray}
801: \delta_{i} &=& \left[ 2\tilde{N}\left(\omega_{s}\right)
802: +1\right]\left( \Omega_{ii}^{(+)} -\Omega_{ii}^{(-)}\right)
803: \label{t37}
804: \end{eqnarray}
805: represent a part of the intensity dependent Lamb shift of
806: the atomic levels, while
807: \begin{eqnarray}
808: \Omega_{ij} = -\left( \Omega_{ij}^{(+)} +\Omega_{ij}^{(-)}\right) \
809: ,\qquad (i\neq j)
810: \label{t38}
811: \end{eqnarray}
812: represents the vacuum induced coherent (dipole-dipole) interaction
813: between the atoms. It is well known that to obtain a complete
814: calculation of the Lamb shift, it is necessary to extend the calculations
815: to a second-order multilevel Hamiltonian including electron mass
816: renormalisation~\cite{sz97}.
817:
818: The parameters $\delta _{i}$ are usually absorbed into the atomic
819: frequencies $\omega _{i}$, by redefining the frequencies
820: $\tilde{\omega}_{i}=\omega _{i}+\delta_{i}$ and are not often
821: explicitly included in the master equations. The other parameters,
822: $\Omega_{ij}^{(M)}$ and $\Omega_{ij}$, do not appear as a shift of the
823: atomic levels. One can show by the calculation of the integral appearing in
824: Eq.~(\ref{t35}) that the parameter $\Omega_{ij}^{(M)}$ is
825: negligibly small when the carrier frequency of the squeezed field is tuned
826: close to the atomic frequencies~\cite{fd91,mil,oc,pk89}. On the other hand,
827: the parameter $\Omega_{ij}$ is independent of the squeezing parameters
828: $\tilde{N}\left(\omega_{s}\right)$
829: and $\tilde{M}\left(\omega_{s}\right)$, and arises from the interaction
830: between the atoms through the vacuum field. It can be seen that
831: $\Omega_{ij}$ plays a role of a coherent (dipole-dipole) coupling between
832: the atoms. Thus, the collective interactions between the atoms
833: give rise not only to the modified dissipative spontaneous
834: emission but also lead to a coherent coupling between the atoms.
835:
836: Using the contours integration method, we find from Eq.~(\ref{t38})
837: the explicit form of $\Omega_{ij}$ as~\cite{ftk87,leh,ag74,st64,hh64}
838: \begin{eqnarray}
839: \Omega_{ij} &=&\frac{3}{4}\sqrt{\Gamma _{i}\Gamma _{j}}\left\{ -\left[
840: 1-\left( \bar{\mu}\cdot \bar{r}%
841: _{ij}\right)^{2} \right] \frac{\cos \left( k_{0}r_{ij}\right)
842: }{k_{0}r_{ij}}\right. \nonumber \\
843: &&\left. +\left[ 1 -3\left( \bar{\mu}\cdot
844: \bar{{r}}_{ij}\right)^{2} \right] \left[ \frac{\sin \left(
845: k_{0}r_{ij}\right) }{\left( k_{0}r_{ij}\right) ^{2}}+\frac{\cos \left(
846: k_{0}r_{ij}\right) }{\left( k_{0}r_{ij}\right) ^{3}}\right] \right\} \ .
847: \label{t39}
848: \end{eqnarray}
849:
850: \begin{figure}[t]
851: \begin{center}
852: \includegraphics[width=13cm]{ftfig1.eps}
853: \end{center}
854: \caption{(a) Collective damping
855: $\Gamma_{ij}/\sqrt{\Gamma_{i}\Gamma_{j}}$ and (b) the
856: dipole-dipole interaction
857: $\Omega_{ij}/\sqrt{\Gamma_{i}\Gamma_{j}}$ as a function of
858: $r_{ij}/\lambda$ for $\bar{\mu} \perp
859: \bar{r}_{ij}$ (solid line) and $\bar{\mu} \parallel
860: \bar{r}_{ij}$ (dashed line).}
861: \label{ftfig1}
862: \end{figure}
863:
864: The collective parameters $\Gamma_{ij}$ and $\Omega_{ij}$, which both depend
865: on the interatomic separation, determine the collective properties of the
866: multiatom system. In Fig.~\ref{ftfig1}, we
867: plot $\Gamma_{ij}/\sqrt{\Gamma_{i}\Gamma_{j}}$ and
868: $\Omega_{ij}/\sqrt{\Gamma_{i}\Gamma_{j}}$ as a function of
869: $r_{ij}/\lambda$, where $\lambda$ is the resonant wavelength.
870: For large separations $(r_{ij}\gg \lambda)$ the parameters are
871: very small $(\Gamma _{ij}=\Omega_{ij}\approx 0)$, and become
872: important for $r_{ij}<\lambda/2$. For atomic separations much smaller than
873: the resonant wavelength (the small sample model), the parameters attain their
874: maximal values
875: \begin{equation}
876: \Gamma _{ij}=\sqrt{\Gamma _{i}\Gamma _{j}} \ ,\label{t40}
877: \end{equation}
878: and
879: \begin{equation}
880: \Omega_{ij}\approx \frac{3\sqrt{\Gamma _{i}\Gamma _{j}}}{4\left(
881: k_{0}r_{ij}\right) ^{3}}\left[ 1 -3\left(\bar{\mu}
882: \cdot \bar{r}_{ij}\right)^{2}\right] \ . \label{t41}
883: \end{equation}
884: In this small sample model $\Omega_{ij}$ corresponds to the quasistatic
885: dipole-dipole interaction potential.
886:
887: Equation (\ref{t36}) is the final form of the master equation that gives
888: us an elegant description of the physics involved in the dynamics of
889: interacting atoms. The collective parameters $\Gamma_{ij}$ and
890: $\Omega_{ij}$, which arise from the mutual interaction between the atoms,
891: significantly modify the master equation of a two-atom system. The parameter
892: $\Gamma_{ij}$ introduces a coupling between the atoms through the vacuum
893: field that the spontaneous emission from one of the atoms influences the
894: spontaneous emission from the other. The dipole-dipole interaction
895: term $\Omega_{ij}$ introduces a coherent coupling between the atoms. Owing
896: to the dipole-dipole interaction, the population is coherently
897: transferred back and forth from one atom to the other. Here, the
898: dipole-dipole interaction parameter $\Omega_{ij}$ plays a role
899: similar to that of the Rabi frequency in the atom-field interaction.
900:
901: For the next few sections, we restrict ourselves to the interaction of
902: the atoms with the ordinary vacuum, $\tilde{M}(\omega_{s})
903: =\tilde{N}(\omega_{s}) =0$, and driven by an external coherent laser
904: field. In this case, the master equation~(\ref{t36}) can be written as
905: \begin{eqnarray}
906: \frac{\partial \hat{\rho} }{\partial t} &=&
907: -\frac{i}{\hbar}\left[\hat{H}_{s},\hat{\rho}\right]
908: -\frac{1}{2}\sum_{i,j=1}^{N}\Gamma _{ij}\left( \hat{\rho}
909: S_{i}^{+}S_{j}^{-}+S_{i}^{+}S_{j}^{-}\hat{\rho}
910: -2S_{j}^{-}\hat{\rho} S_{i}^{+}\right) \ , \label{t42}
911: \end{eqnarray}
912: where
913: \begin{eqnarray}
914: \hat{H}_{s} &=& \hbar \sum_{i=1}^{N}\left(\omega_{i}+\delta_{i}\right)
915: S^{z}_{i}
916: +\hbar \sum_{i\neq j}^{N}\Omega_{ij}S^{+}_{i}S^{-}_{j}
917: +\hat{H}_{L} \ ,\label{t43}
918: \end{eqnarray}
919: and
920: \begin{eqnarray}
921: \hat{H}_{L} &=&
922: -\frac{1}{2}\hbar \sum_{i=1}^{N}\left[\Omega
923: \left(\vec{r}_{i}\right)S_{i}^{+}
924: e^{i\left(\omega_{L}t+\phi_{L}\right)} + \rm{H.c.}\right] \
925: ,\label{t44}
926: \end{eqnarray}
927: is the interaction Hamiltonian of the atoms with a classical
928: coherent laser field of the Rabi frequency $\Omega \left(\vec{r}_{i}\right)$,
929: the angular frequency $\omega_{L}$ and phase $\phi_{L}$.
930:
931: Note that the Rabi frequencies of the driving field are evaluated at
932: the positions of the atoms and are defined as~\cite{ae}
933: \begin{eqnarray}
934: \Omega \left(\vec{r}_{i}\right)\equiv \Omega_{i} =\vec{\mu}_{i}\cdot
935: \vec{E}_{L}e^{i\vec{k}_{L}
936: \cdot \vec{r}_{i}}/\hbar \ , \label{t45}
937: \end{eqnarray}
938: where $\vec{E}_{L}$ is the amplitude and $\vec{k}_{L}$ is the wave
939: vector of the driving field, respectively. The Rabi
940: frequencies depend on the positions of the atoms and can be
941: different for the atoms located at different points. For example,
942: if the dipole moments of the atoms are parallel, the Rabi
943: frequencies $\Omega_{i}$ and $\Omega_{j}$ of two arbitrary atoms
944: separated by a distance $r_{ij}$ are related by
945: \begin{eqnarray}
946: \Omega_{j} = \Omega_{i}
947: \frac{\left|\vec{\mu}_{j}\right|}
948: {\left|\vec{\mu}_{i} \right|}e^{i\vec{k}_{L}\cdot \vec{r}_{ij}}
949: \ ,\label{t46}
950: \end{eqnarray}
951: where $\vec{r}_{ij}$ is the vector in the
952: direction of the interatomic axis and $|\vec{r}_{ij}| =r_{ij}$ is the
953: distance between the atoms. Thus, for two identical atoms
954: $(|\vec{\mu}_{i}|= |\vec{\mu}_{j}|)$, the Rabi frequencies
955: differ by the phase factor exp$(i\vec{k}_{L}\cdot \vec{r}_{ij})$
956: arising from different position coordinates of the atoms. However,
957: the phase factor depends on the orientation
958: of the interatomic axis in respect to the direction of propagation of the
959: driving field, and therefore exp$(i\vec{k}_{L}\cdot \vec{r}_{ij})$ can be
960: equal to one, even for large interatomic separations $r_{ij}$. This happens
961: when the direction of propagation of the driving field is
962: perpendicular to the interatomic axis, $\vec{k}_{L}\cdot \vec{r}_{ij}=0$.
963: For directions different from perpendicular,
964: $\vec{k}_{L}\cdot \vec{r}_{ij}\neq 0$, and then the atoms are in
965: nonequivalent positions in the driving field, with
966: different Rabi frequencies~$(\Omega_{i}\neq \Omega_{j})$. For a very
967: special geometrical configuration of the atoms that are confined to a
968: volume with linear dimensions that are much smaller compared to the
969: laser wavelength, the phase factor exp$(i\vec{k}_{L}\cdot
970: \vec{r}_{ij}) \approx 1$, and then the Rabi frequencies are
971: independent of the atomic positions. This specific configuration of
972: the atoms is known as the small sample model or the Dicke model,
973: and do not correspond in general to the experimentally realised atomic
974: systems such as atomic beams or trapped atoms.
975:
976: The formalism presented here for the derivation of the master
977: equation can be easily extended to the case of $N$ multi-level
978: atoms~\cite{kbr87,bgz00,bg02,ryf} and atoms interacting with colour (frequency
979: dependent) reservoirs~\cite{lm88,kky,fwd97,mtf99} or photonic band-gap
980: materials~\cite{lambro,dkw}. Freedhoff~\cite{hel87} has extended the master
981: equation formalism to electric quadrupole transitions in atoms.
982: In the following sections, we will apply the
983: master equations~(\ref{t36}) and (\ref{t42}) to a wide variety of cases
984: ranging from two identical as well as nonidentical atoms interacting with
985: the ordinary vacuum to atoms driven by a laser field and finally to atoms
986: interacting with a squeezed vacuum field.
987:
988:
989: \subsection{Quantum jump approach}\label{ftsec22}
990:
991:
992: The master equation is a very powerful tool for calculations of the
993: dynamics of Markovian systems which assume that the bandwidth of the
994: vacuum field is broadband. The Markovian master equation leads to
995: linear differential equations for the density matrix elements that
996: can be solved numerically or analytically by the direct integration.
997:
998: An alternative to the master equation technique is quantum jump
999: approach. This technique is based on quantum
1000: trajectories~\cite{car93} that are equivalent to the Monte Carlo
1001: wave-function approach~\cite{dcm92,pk98}, and has been developed
1002: largely in connection with problems involving prediction of all
1003: possible evolution trajectories of a given system. This approach can
1004: be used to predict all evolution trajectories of a single quantum
1005: system which stochastically emits photons. Our review of this approach
1006: will concentrate on the example considered by Beige and
1007: Hegerfeldt~\cite{bh98} of two identical two-level atoms interacting
1008: with the three-dimensional EM field whose the modes are in the
1009: ordinary vacuum states.
1010:
1011: In the quantum jump approach it is assumed that the probability
1012: density for a photon emission is known for all times $t$, and
1013: therefore the state of the atoms changes abruptly. After one photon
1014: emission the system jumps into another state, which can be determined
1015: with the help of the so called reset operator. The continuous time
1016: evolution of the system between two successive photon emissions is
1017: determined by the conditional Hamiltonian~$\hat{H}_{c}$. Suppose that
1018: at time $t_{0}$ the state of the combined system of the atoms and EM
1019: field is given by
1020: \begin{eqnarray}
1021: \left|\Psi \right\rangle \left\langle \Psi \right|=
1022: \left|0 \right\rangle \hat{\rho}
1023: \left\langle 0\right| \ ,\label{t47}
1024: \end{eqnarray}
1025: where $\hat{\rho}$ is the density operator of the atoms and
1026: $\left|0 \right\rangle$ is the vacuum state of the field. After a
1027: time $\Delta t$ a photon is detected and then the state of the system
1028: changes to
1029: \begin{eqnarray}
1030: {\cal{P}}\hat{U}_{I}\left(t_{0}+\Delta t,t_{0}\right)
1031: \left|0 \right\rangle \hat{\rho} \left\langle 0\right|
1032: \hat{U}_{I}^{\dagger}\left(t_{0}+\Delta t,t_{0}\right)
1033: {\cal{P}} \ ,\label{t48}
1034: \end{eqnarray}
1035: where ${\cal{P}} =1-\left|0 \right\rangle \left\langle 0\right|$ is
1036: the projection onto the one photon space, and
1037: \begin{eqnarray}
1038: \hat{U}_{I}\left(t,t_{0}\right)=
1039: e^{-\frac{i}{\hbar}\hat{V}(t)(t-t_{0})} \label{t49}
1040: \end{eqnarray}
1041: is the evolution operator with the Hamiltonian $\hat{V}(t)$ given
1042: in Eq.~(\ref{t8}).
1043:
1044: The non-normalised state of the atomic system, denoted as
1045: $R(\hat{\rho})\Delta t$, is obtained by taking trace of
1046: Eq.~(\ref{t48}) over the field states
1047: \begin{eqnarray}
1048: R(\hat{\rho})\Delta t = {\rm Tr}_{F}\left(
1049: {\cal{P}}\hat{U}_{I}\left(t_{0}+\Delta t,t_{0}\right)
1050: \left|0 \right\rangle \hat{\rho} \left\langle 0\right|
1051: \hat{U}_{I}^{\dagger}\left(t_{0}+\Delta t,t_{0}\right)
1052: {\cal{P}}\right) \ ,\label{t50}
1053: \end{eqnarray}
1054: where $R(\hat{\rho})$ is called the non-normalised reset state and
1055: the corresponding operator $\hat{R}$ is called the reset
1056: operator.
1057:
1058: Using the perturbation theory and Eq.~(\ref{t8}), we find the
1059: explicit form of $\hat{R}(\hat{\rho})$ for the two-atom system as
1060: \begin{eqnarray}
1061: \hat{R}(\hat{\rho}) &=&
1062: \frac{1}{2}\left(C_{12}^{\ast}+C_{21}\right)S_{1}^{-}\hat{\rho}
1063: S_{2}^{+} +\frac{1}{2}\left(C_{12}+C_{21}^{\ast}\right)S_{2}^{-}
1064: \hat{\rho} S_{1}^{+} \nonumber \\
1065: && +\Gamma\left(S_{1}^{-}\hat{\rho} S_{1}^{+}
1066: +S_{2}^{-}\hat{\rho} S_{2}^{+}\right) \ ,\label{t51}
1067: \end{eqnarray}
1068: where
1069: \begin{eqnarray}
1070: C_{ij} &=& -\frac{3}{2}i\Gamma e^{ik_{0}r_{ij}}\left\{
1071: \left[ 1 -\left( \bar{\mu}\cdot \bar{{r}}_{ij}\right)^{2}
1072: \right] \frac{1}{k_{0}r_{ij}}\right. \nonumber \\
1073: &&\left. +\left[ 1 -3\left( \bar{\mu}\cdot
1074: \bar{{r}}_{ij}\right)^{2} \right]\left(\frac{i}{(k_{0}r_{ij})^{2}}
1075: -\frac{1}{(k_{0}r_{ij})^{3}}\right)\right\} \ .\label{t52}
1076: \end{eqnarray}
1077: Note that ${\rm Re}C_{ij} =\Gamma_{ij}$ and ${\rm Im}C_{ij}
1078: =2\Omega_{ij}$, where $\Gamma_{ij}$ and $\Omega_{ij}$ are the
1079: collective atomic parameters, given in Eqs.~(\ref{t32}) and (\ref{t39}),
1080: respectively.
1081:
1082: The time evolution of the system under the condition that no photon is
1083: emitted is described by the conditional Hamiltonian $\hat{H}_{c}$,
1084: which is found from the relation
1085: \begin{eqnarray}
1086: 1 -\frac{i}{\hbar}\hat{H}_{c}\Delta t &=& \left\langle 0\right|
1087: \hat{U}_{I}\left(t_{0}+\Delta t,t_{0}\right)\left|0
1088: \right\rangle \ ,\label{t53}
1089: \end{eqnarray}
1090: where $\Delta t$ is a short evolution time such that $\Delta t <
1091: 1/\Gamma$. Using second order perturbation theory, we find from
1092: Eq.~(\ref{t53}) that the conditional Hamiltonian for the two-atom
1093: system is of the form
1094: \begin{eqnarray}
1095: \hat{H}_{c} &=& \frac{\hbar}{2i}\left[\Gamma
1096: \left(S_{1}^{+}S_{1}^{-}+S_{2}^{+}S_{2}^{-}\right)
1097: +C_{12}S_{1}^{+}S_{2}^{-} +C_{21}S_{2}^{+}S_{1}^{-}\right] \
1098: .\label{t54}
1099: \end{eqnarray}
1100: Hence, between photon emissions the time evolution of the system is
1101: given by an operator
1102: \begin{eqnarray}
1103: \hat{U}_{c}\left(t_{0}+\Delta t,t_{0}\right)=
1104: e^{-\frac{i}{\hbar}\hat{H}_{c}(t-t_{0})} \ ,\label{t55}
1105: \end{eqnarray}
1106: which is nonunitary since $\hat{H}_{c}$ is non-Hermitian, and the
1107: state vector of the system is
1108: \begin{eqnarray}
1109: \left|\Psi_{\Delta t}\right\rangle &=&
1110: \hat{U}_{c}\left(t_{0}+\Delta t,t_{0}\right)
1111: \left|\Psi_{0}\right\rangle \ .\label{t56}
1112: \end{eqnarray}
1113: Then, the probability to detect no photon until time $t$ is given by
1114: \begin{eqnarray}
1115: P\left(t;\left|\Psi_{0}\right\rangle \right) &=& \left|
1116: \hat{U}_{c}\left(t,t_{0}\right)
1117: \left|\Psi_{0}\right\rangle \right|^{2} \ .\label{t57}
1118: \end{eqnarray}
1119: The probability density $w_{1}\left(t;\left|\Psi_{0}\right\rangle
1120: \right)$ of detecting a photon at time $t$ is defined as
1121: \begin{eqnarray}
1122: w_{1}\left(t;\left|\Psi_{0}\right\rangle \right)= -\frac{d}{dt}
1123: P\left(t;\left|\Psi_{0}\right\rangle \right) ,\label{t58}
1124: \end{eqnarray}
1125: and is often called the waiting time distribution.
1126:
1127: The results (\ref{t57}) and (\ref{t58}) show that in the quantum jump
1128: method one calculates the times of the photon
1129: detection stochastically. Starting at $t=t_{0}$ with a pure state, the
1130: state develops according to $\hat{U}_{c}$ until the first emission at
1131: some time $t_{1}$, determined from the waiting time $w_{1}$. Then the
1132: state is reset, according to Eq.~(\ref{t51}), to a new density matrix
1133: and the system evolves again according to $\hat{U}_{c}$ until the
1134: second emission appearing at some time $t_{2}$, and the procedure
1135: repeats until the final time $t_{n}$. In this way, we obtain a set of
1136: trajectories of the atomic evolution. The ensemble of such
1137: trajectories yields to equations of motion which are solved using the
1138: standard analytical or numerical methods. As a practical matter,
1139: individual trajectories are generally not observed. The ensemble average
1140: over all possible trajectories leads to equations of motion which are
1141: equivalent to the equations of motion derived from the master
1142: equation of the system. Thus, the quantum jump approach is consistent
1143: with the master equation method. However, the advantage of the
1144: quantum jump approach over the master equation method is that it
1145: allows to predict all possible trajectories of a single system. Using
1146: this approach, it has been demonstrated that environment induced
1147: measurements can assist in the realization of universal gates for
1148: quantum computing~\cite{bbtk}. Cabrillo {\it et al.}~\cite{cabr} have
1149: applied the method to demonstrate entangling between distant atoms by
1150: interference. Sch\"{o}n and Beige~\cite{sb} have demonstrated the
1151: advantage of the method in the analysis of a two-atom double-slit
1152: experiment.
1153:
1154:
1155: \section{Entangled atomic states}\label{ftsec3}
1156:
1157:
1158: The modification of spontaneous emission by the collective damping
1159: and in particular the presence of the dipole-dipole interaction
1160: between the atoms suggest that the bare atomic states are no longer
1161: the eigenstates of the atomic system. We will illustrate this on a
1162: system of two identical as well as nonidentical atoms, and present
1163: a general formalism for diagonalization of the Hamiltonian of the
1164: atoms in respect to the dipole-dipole interaction.
1165:
1166: In the absence of the dipole-dipole interaction and the driving laser
1167: field, the space of the two-atom system is spanned by four
1168: product states
1169: \begin{eqnarray}
1170: \ket {g_{1}}\ket {g_{2}} \ ,\quad \ket {e_{1}}\ket {g_{2}} \ ,\quad
1171: \ket {g_{1}}\ket
1172: {e_{2}} \ ,\quad \ket {e_{1}}\ket {e_{2}} \ ,\label{t59}
1173: \end{eqnarray}
1174: with corresponding energies
1175: \begin{eqnarray}
1176: E_{gg}=-\hbar \omega_{0} \ ,\quad E_{eg}=-\hbar \Delta
1177: \ ,\quad E_{ge}=\hbar \Delta \ ,\quad E_{ee}=\hbar \omega_{0}
1178: \ ,\label{t60}
1179: \end{eqnarray}
1180: where $\omega_{0}=\frac{1}{2}\left(\omega_{1}+\omega_{2}\right)$ and
1181: $\Delta =\frac{1}{2}\left(\omega_{2}-\omega_{1}\right)$.
1182:
1183: The product states $\ket {e_{1}}\ket {g_{2}}$ and $\ket {g_{1}}\ket
1184: {e_{2}}$ form a pair of nearly degenerated states. When we include the
1185: dipole-dipole interaction between the atoms, the product states
1186: combine into two linear superpositions (entangled states), with
1187: their energies shifted from $\pm \hbar \Delta$ by the
1188: dipole-dipole interaction energy. To see this, we begin with
1189: the Hamiltonian of two atoms including the dipole-dipole interaction
1190: \begin{eqnarray}
1191: \hat{H}_{aa} = \sum_{i=1}^{2}\hbar \omega_{i} S_{i}^{z}
1192: + \hbar \sum_{i\neq j}\Omega_{ij}S_{i}^{+}S_{j}^{-} \ .\label{t61}
1193: \end{eqnarray}
1194: In the basis of the product states~(\ref{t59}), the
1195: Hamiltonian~(\ref{t61}) can be written in a matrix form as
1196: \begin{eqnarray}
1197: \hat{H}_{aa} &=& \hbar \left(
1198: \begin{array}{cccc}
1199: -\omega_{0} & 0 & 0 & 0 \\
1200: 0 & -\Delta & \Omega_{12} & 0 \\
1201: 0 & \Omega_{12} & \Delta & 0 \\
1202: 0 & 0 & 0 & \omega_{0}
1203: \end{array}
1204: \right) \ . \label{t62}
1205: \end{eqnarray}
1206:
1207: Evidently, in the presence of the dipole-dipole interaction
1208: the matrix~(\ref{t62}) is not diagonal, which indicates that the product
1209: states~(\ref{t59}) are not the eigenstates of the two-atom
1210: system. We will diagonalize the matrix~(\ref{t62}) separately for the
1211: case of identical $(\Delta =0)$ and nonidentical $(\Delta \neq 0)$
1212: atoms to find eigenstates of the systems and their energies.
1213:
1214:
1215: \subsection{Entangled states of two identical atoms}\label{ftsec31}
1216:
1217:
1218: Consider first a system of two identical atoms $(\Delta =0)$.
1219: In order to find energies and corresponding eigenstates of the system,
1220: we have to diagonalize the matrix~(\ref{t62}). The resulting
1221: energies and corresponding eigenstates of the system are~\cite{dic,leh}
1222: \begin{eqnarray}
1223: && E_{g} =-\hbar \omega_{0} \ ,\qquad \ket g = \ket {g_{1}}\ket {g_{2}}
1224: \ ,\nonumber \\
1225: && E_{s} = \hbar \Omega_{12} \ ,\qquad
1226: \ket s = \frac{1}{\sqrt{2}}\left( \ket {e_{1}}\ket {g_{2}}
1227: +\ket {g_{1}}\ket {e_{2}}\right) \ ,\nonumber \\
1228: && E_{a} = -\hbar \Omega_{12} \ ,\qquad
1229: \ket a = \frac{1}{\sqrt{2}}\left( \ket {e_{1}}\ket {g_{2}}
1230: -\ket {g_{1}}\ket {e_{2}}\right) \ ,\nonumber \\
1231: && E_{e} = \hbar \omega_{0} \ ,\qquad
1232: \ket e = \ket {e_{1}}\ket {e_{2}} \ .\label{t63}
1233: \end{eqnarray}
1234:
1235: The eigenstates~(\ref{t63}), first introduced by Dicke~\cite{dic}, are
1236: known as the collective states of two interacting
1237: atoms. The ground state $\ket g$ and the upper state $\ket e$ are not
1238: affected by the dipole-dipole interaction, whereas the states $\ket s$
1239: and $\ket a$ are shifted from their unperturbed energies by the amount
1240: $\pm \Omega_{12}$, the dipole-dipole energy.
1241: The most important property of the collective states $\ket s$ and
1242: $\ket a$ is that they are an example of maximally entangled states
1243: of the two-atom system. The states are linear superpositions of the
1244: product states which cannot be separated into product states of
1245: the individual atoms.
1246: %
1247: \begin{figure}[t]
1248: \begin{center}
1249: \includegraphics[width=10cm]{ftfig2.eps}
1250: \end{center}
1251: \caption{Collective states of two identical atoms. The energies of
1252: the symmetric and antisymmetric states are shifted by the dipole-dipole
1253: interaction $\Omega_{12}$. The arrows indicate possible one-photon
1254: transitions.}
1255: \label{ftfig2}
1256: \end{figure}
1257:
1258: We show the collective states of two identical atoms in Fig.~\ref{ftfig2}.
1259: It is seen that in the collective states representation, the two-atom system
1260: behaves as a single four-level system, with the ground state $\ket g$, the
1261: upper state $\ket e$, and two intermediate states: the symmetric
1262: state $\ket s$ and the antisymmetric state $\ket a$. The energies of
1263: the intermediate states depend on the dipole-dipole interaction and
1264: these states suffer a large shift when the interatomic separation is
1265: small. There are two transition channels $\ket e \rightarrow \ket s
1266: \rightarrow \ket g$ and $\ket e \rightarrow \ket a \rightarrow \ket
1267: g$, each with two cascade nondegenerate transitions. For two
1268: identical atoms, these two channels are uncorrelated, but the
1269: transitions in these channels are damped with significantly different
1270: rates. To illustrate these features, we transform the master
1271: equation~(\ref{t42})
1272: into the basis of the collective states~(\ref{t63}). We define
1273: collective operators $A_{ij}=\ket i\bra j$, where $i,j=e,a,s,g$, that
1274: represent the energies $(i=j)$ of the collective
1275: states and coherences $(i\neq j)$. Using Eq.~(\ref{t63}), we find that
1276: the collective operators are related to the atomic operators
1277: $S_{i}^{\pm}$ through the following identities
1278: \begin{eqnarray}
1279: S_{1}^{+} &=& \frac{1}{\sqrt{2}}\left(A_{es}-A_{ea}+A_{sg}
1280: +A_{ag}\right) \ ,\nonumber \\
1281: S_{2}^{+} &=& \frac{1}{\sqrt{2}}\left(A_{es}+A_{ea}+A_{sg}
1282: -A_{ag}\right) \ .\label{t64}
1283: \end{eqnarray}
1284:
1285: Substituting the transformation identities into Eq.~(\ref{t42}),
1286: we find that in the basis of the collective states the master equation
1287: of the system can be written as
1288: \begin{eqnarray}
1289: \frac{\partial}{\partial t}\hat{\rho} &=&
1290: -\frac{i}{\hbar}\left[\hat{H}_{as}, \hat{\rho}\right]
1291: +\left(\frac{\partial}{\partial t}\hat{\rho}\right)_{s}
1292: +\left(\frac{\partial}{\partial t}\hat{\rho}\right)_{a} \
1293: ,\label{t65}
1294: \end{eqnarray}
1295: where
1296: \begin{eqnarray}
1297: \hat{H}_{as} &=&
1298: \hbar\left[\omega_{0}\left(A_{ee}-A_{gg}\right)
1299: +\Omega_{12}\left(A_{ss}-A_{aa}\right)\right] \nonumber \\
1300: &-& \frac{\hbar}{2\sqrt{2}}
1301: \left(\Omega_{1} +\Omega_{2}\right)\left[\left(A_{es}+A_{sg}\right)
1302: e^{i\left(\omega_{L}t+\phi_{L}\right)} +{\rm H.c.} \right] \nonumber \\
1303: &-& \frac{\hbar}{2\sqrt{2}} \left(\Omega_{2}
1304: -\Omega_{1}\right)\left[\left(A_{ea}-A_{ag}\right)
1305: e^{i\left(\omega_{L}t+\phi_{L}\right)} +{\rm
1306: H.c.} \right] \ ,\label{t66}
1307: \end{eqnarray}
1308: is the Hamiltonian of the interacting atoms and the driving laser
1309: field,
1310: \begin{eqnarray}
1311: \left(\frac{\partial}{\partial t}\hat{\rho}\right)_{s} &=&
1312: - \frac{1}{2}\left(\Gamma +\Gamma_{12}\right)
1313: \left\{\left(A_{ee}+A_{ss}\right)\hat{\rho} +\hat{\rho} \left(A_{ee}
1314: +A_{ss}\right)\right. \nonumber \\
1315: &&- \left. 2\left(A_{se}+A_{gs}\right)\hat{\rho} \left(A_{es}
1316: +A_{sg}\right)\right\} \ ,\label{t67}
1317: \end{eqnarray}
1318: describes dissipation through the cascade $\ket e \rightarrow
1319: \ket s \rightarrow \ket g$ channel involving the symmetric state
1320: $\ket s$, and
1321: \begin{eqnarray}
1322: \left(\frac{\partial}{\partial t}\hat{\rho}\right)_{a} &=&
1323: - \frac{1}{2}\left(\Gamma -\Gamma_{12}\right)
1324: \left\{\left(A_{ee}+A_{aa}\right)\hat{\rho} +\hat{\rho} \left(A_{ee}
1325: +A_{aa}\right)\right. \nonumber \\
1326: &&-\left. 2\left(A_{ae}-A_{ga}\right)\hat{\rho} \left(A_{ea}
1327: -A_{ag}\right)\right\} \ ,\label{t68}
1328: \end{eqnarray}
1329: describes dissipation through the cascade
1330: $\ket e \rightarrow \ket a \rightarrow \ket g$ channel involving the
1331: antisymmetric state $\ket a$.
1332:
1333: We will call the two cascade channels $\ket e \rightarrow \ket s
1334: \rightarrow \ket g$ and $\ket e \rightarrow \ket a \rightarrow
1335: \ket g$ as symmetric and antisymmetric transitions, respectively.
1336: The first term in
1337: $\hat{H}_{as}$ is the energy of the collective states, while the
1338: second and third terms are the interactions of the laser field with
1339: the symmetric and antisymmetric transitions, respectively.
1340: One can see from Eqs.~(\ref{t65})-(\ref{t68}) that the symmetric and
1341: antisymmetric transitions are uncorrelated and decay with different
1342: rates; the symmetric transitions decay with an enhanced (superradiant)
1343: rate $(\Gamma +\Gamma_{12})$, whereas the antisymmetric transitions
1344: decay with a reduced (subradiant) rate $(\Gamma -\Gamma_{12})$. For
1345: $\Gamma =\Gamma_{12}$, which appears when the interatomic separation
1346: is much smaller than the resonant wavelength, the antisymmetric
1347: transitions decouple from the driving field and does not decay. In this
1348: case, the antisymmetric state is completely decoupled from the remaining
1349: states and the system decays only through the symmetric channel.
1350: Hence, for $\Gamma_{12}=\Gamma$ the system reduces to a three-level
1351: cascade system, referred to as the small-sample model or two-atom Dicke
1352: model~\cite{dic,leh,ag74}. The model assumes that the atoms are close
1353: enough that we can ignore any effects resulting from different spatial
1354: positions of the atoms. In other words, the phase factors $\exp
1355: (i\vec{k}\cdot \vec{r}_{i})$ are assumed to have the same value for
1356: all the atoms, and are set equal to one. This assumption may prove
1357: difficult in experimental realization as the present atom trapping and
1358: cooling techniques can trap two atoms at distances of the order of a
1359: resonant wavelength~\cite{eich,deb,ber,tos}. At these distances the
1360: collective damping parameter $\Gamma_{12}$ differs significantly from
1361: $\Gamma$ (see Fig.~\ref{ftfig1}), and we cannot ignore the transitions
1362: to and from the antisymmetric state. We can, however,
1363: employ the Dicke model to spatially extended atomic systems. This
1364: could be achieved assuming that the observation time of the atomic
1365: dynamics is shorter than $\Gamma^{-1}$. The antisymmetric state $\ket
1366: a$ decays on a time scale $\sim (\Gamma -\Gamma_{12})^{-1}$, which for
1367: $\Gamma_{12}\approx \Gamma$ is much longer than $\Gamma^{-1}$. On the
1368: other hand, the symmetric state decays on a time scale $\sim
1369: (\Gamma +\Gamma_{12})^{-1}$, which is shorter than $\Gamma^{-1}$.
1370: Clearly, if we consider short observation times, the antisymmetric
1371: state does not participate in the dynamics and the system can be
1372: considered as evolving only between the Dicke states.
1373:
1374: Although the symmetric and antisymmetric transitions of the collective
1375: system are uncorrelated, the dynamics of the four-level system
1376: may be significantly different from the three-level Dicke model.
1377: As an example, consider the total intensity of the fluorescence field
1378: emitted from a two-atom system driven by a resonant coherent laser
1379: field $(\omega_{L}=\omega_{0})$.
1380: We make two simplifying assumptions in order to obtain a simple
1381: analytical solution: Firstly, we limit our calculations to the
1382: steady-state intensity. Secondly, we take $\vec{k}_{L}\cdot
1383: \vec{r}_{12}=0$ that corresponds to the direction of propagation of
1384: the driving field perpendicular to the interatomic axis. We emphasize
1385: that these assumptions do not limit qualitatively the physics of the
1386: system, as experiments are usually performed in the steady-state,
1387: and with $\vec{k}_{L}\cdot \vec{r}_{12}=0$ the interatomic separation
1388: $r_{12}$ may still be any size relative to the resonant wavelength.
1389:
1390: We consider the radiation intensity $I(\vec{R},t)$ detected at a
1391: point $\vec{R}$ at the moment of time $t$. If the detection point
1392: $\vec{R}$ is in the far-field zone of the radiation emitted by the atomic
1393: system, then the intensity can be expressed in terms of the first-order
1394: correlation functions of the atomic dipole operators as~\cite{leh,ag74}
1395: \begin{eqnarray}
1396: I\left(\vec{R},t\right) = u(\vec{R})\sum_{i,j=1}^{2}\langle
1397: S_{i}^{+}\left(t-R/c\right)
1398: S_{j}^{-}\left(t-R/c\right)\rangle e^{ik\bar{R}\cdot \vec{r}_{ij}}
1399: \ ,\label{t69}
1400: \end{eqnarray}
1401: where
1402: \begin{equation}
1403: u(\vec{R}) = \left(\omega_{0}^{4}\mu^{2}/2R^{2}c^{4}\pi
1404: \varepsilon_{0}\right)\sin^{2}\varphi \label{t70}
1405: \end{equation}
1406: is a constant which depends on the
1407: geometry of the system, $\varphi$ the angle between the
1408: observation direction $\vec{R}=R\bar{R}$ and the atomic dipole
1409: moment $\vec{\mu}$.
1410:
1411: On integrating over all directions, Eq.~(\ref{t69}) yields the total
1412: radiation intensity given in photons per second as
1413: \begin{eqnarray}
1414: I\left(t\right) = \sum_{i,j=1}^{2}\Gamma_{ij}\langle
1415: S_{i}^{+}\left(t-R/c\right)
1416: S_{j}^{-}\left(t-R/c\right)\rangle \ .\label{t71}
1417: \end{eqnarray}
1418: The atomic correlation functions, appearing in Eq.~(\ref{t71}),
1419: are found from the master equation~(\ref{t42}). There are,
1420: however, two different steady-state solutions of the master
1421: equation~(\ref{t42}) depending on whether the collective damping
1422: rates $\Gamma_{12}=\Gamma$ or $\Gamma_{12}\neq
1423: \Gamma$~\cite{ftk81,ftk83,hsf82}.
1424:
1425: For $\Gamma_{12}\neq \Gamma$ and $\vec{k}_{L}\cdot \vec{r}_{12}=0$,
1426: the steady-state solutions for the atomic correlation functions are
1427: \begin{eqnarray}
1428: \left\langle S_{1}^{+}S_{1}^{-}\right\rangle =
1429: \left\langle S_{2}^{+}S_{2}^{-}\right\rangle &=&
1430: \frac{2\Omega^{4} +\Gamma^{2}\Omega^{2}}{4D}
1431: \ ,\nonumber \\
1432: \left\langle S_{1}^{+}S_{2}^{-}\right\rangle =
1433: \left\langle S_{2}^{+}S_{1}^{-}\right\rangle &=&
1434: \frac{\Gamma^{2}\Omega^{2}}{4D} \ ,\label{t72}
1435: \end{eqnarray}
1436: where
1437: \begin{eqnarray}
1438: D = \Omega^{4}+\left(\Omega^{2}
1439: +\Omega_{12}^{2}\right)\Gamma^{2} +\frac{1}{4}\Gamma^{2}\left(\Gamma
1440: +\Gamma_{12}\right)^{2} \ .\label{t73}
1441: \end{eqnarray}
1442: If we take $\Gamma_{12} =\Gamma$ and $\Omega_{12}=0$, that corresponds
1443: to the two-atom Dicke model, the steady-state solutions for the atomic
1444: correlation functions are of the following form
1445: \begin{eqnarray}
1446: \left\langle S_{1}^{+}S_{1}^{-}\right\rangle =
1447: \left\langle S_{2}^{+}S_{2}^{-}\right\rangle &=&
1448: \frac{3\Omega^{4} +2\Omega^{2}\Gamma^{2}}{2D^{\prime}}
1449: \ ,\nonumber \\
1450: \left\langle S_{1}^{+}S_{2}^{-}\right\rangle =
1451: \left\langle S_{2}^{+}S_{1}^{-}\right\rangle &=&
1452: \frac{\Omega^{4}+2\Omega^{2}\Gamma^{2}}{2D^{\prime}} \ ,\label{t74}
1453: \end{eqnarray}
1454: where
1455: \begin{eqnarray}
1456: D^{\prime} = 3\Omega^{4}+4\Gamma^{2}\Omega^{2}+4\Gamma^{4}
1457: \ .\label{t75}
1458: \end{eqnarray}
1459:
1460: In the limit of a strong driving field, $\Omega \gg \Gamma$, the
1461: steady-state total radiation intensity from the two-atom Dicke model
1462: is equal to $4\Gamma/3$. However, for the spatially separated atoms
1463: $I_{ss}=\lim_{t\rightarrow \infty}I\left(t\right) =\Gamma$, which is
1464: twice of the intensity from a single atom~\cite{mw}. There is no
1465: additional enhancement of the intensity.
1466:
1467: Note that in the limit of $r_{12}\rightarrow 0$, the steady-state
1468: solution~(\ref{t72}) does not reduce to that of the Dicke model,
1469: given in Eq.~(\ref{t74}). This fact is connected with conservation of
1470: the total spin $S^{2}$, that $S^{2}$ is a constant of motion for the
1471: Dicke model and $S^{2}$ not being a constant of motion for a spatially
1472: extended system of atoms~\cite{ftk81,ftk83}. We can explain it by
1473: expressing the square of the total spin of the two-atom system in
1474: terms of the density matrix elements of the collective system as
1475: \begin{eqnarray}
1476: S^{2}\left(t\right) = 2 -2\rho_{aa}\left(t\right) \ .\label{t76}
1477: \end{eqnarray}
1478: It is clear from Eq.~(\ref{t76}) that $S^{2}$ is conserved only in
1479: the Dicke model, in which the antisymmetric state is ignored. For a
1480: spatially extended system the antisymmetric state participates fully
1481: in the dynamics and $S^{2}$ is not conserved. The Dicke model reaches
1482: steady state between the triplet states $\ket e$, $\ket s$, and $\ket
1483: g$, while the spatially extended two-atom system reaches steady state
1484: between the triplet and the antisymmetric states.
1485:
1486: Amin and Cordes~\cite{ac78} calculated the total radiation intensity
1487: from an $N$-atom Dicke model and showed the intensity is $N(N+2)/3$
1488: times that for a single atom, which they called "scaling factor". The
1489: above calculations show that the scaling factor is characteristic of
1490: the small sample model and does not exist in spatially extended atomic
1491: systems. Thus, in physical systems the antisymmetric state plays
1492: important role and as we have shown its presence affects the
1493: steady-state fluorescence intensity. The antisymmetric state can also
1494: affect other phenomena, for example, photon
1495: antibunching~\cite{ftk81a}, and purity of two-photon entangled
1496: states, that is discussed in Sec.~\ref{ftsec10}.
1497:
1498:
1499:
1500: \subsection{Collective states of two nonidentical atoms}\label{ftsec32}
1501:
1502:
1503:
1504: For two identical atoms, the dipole-dipole interaction leads to the
1505: maximally entangled symmetric and antisymmetric states that decay
1506: independently with different damping rates. Furthermore, in the case
1507: of the small sample model of two atoms the antisymmetric state decouples
1508: from the external coherent field and the environment, and consequently
1509: does not decay. The decoupling of the antisymmetric state from the
1510: coherent field prevents the state from the external coherent interactions.
1511: This is not, however, an useful property from the point of view of quantum
1512: computation where it is required to prepare entangled states which are
1513: decoupled from the external environment and simultaneously should be
1514: accessible by coherent processes. This requirement can be achieved if
1515: the atoms are not identical, and we will discuss here some consequences
1516: of the fact that the atoms could have different transition frequencies
1517: or different spontaneous emission rates. To make our discussion more
1518: transparent, we will concentrate on two specific cases:
1519: (1) $\Delta \neq 0$ and $\Gamma_{1}=\Gamma_{2}$, and (2)
1520: $\Delta =0$ and $\Gamma_{1}\neq \Gamma_{2}$.
1521:
1522:
1523: \subsubsection{ The case $\Delta \neq 0$ and
1524: $\Gamma_{1}=\Gamma_{2}$}\label{ftsec321}
1525:
1526:
1527: When the atoms are nonidentical with different
1528: transition frequencies, the states~(\ref{t63}) are no longer the
1529: eigenstates of the Hamiltonian~(\ref{t60}). The diagonalization of the
1530: matrix~(\ref{t62}) with $\Delta \neq 0$ leads to the following
1531: energies and corresponding eigenstates~\cite{ftk86}
1532: \begin{eqnarray}
1533: && E_{g} =-\hbar \omega_{0} \ ,\qquad
1534: \ket g = \ket {g_{1}}\ket {g_{2}} \ ,\nonumber \\
1535: && E_{s^{\prime}} = \hbar w \ ,\qquad
1536: \ket {s^{\prime}} = \beta \ket {e_{1}}\ket {g_{2}}
1537: +\alpha \ket {g_{1}}\ket {e_{2}} \ ,\nonumber \\
1538: && E_{a^{\prime}} = -\hbar w \ ,\qquad
1539: \ket {a^{\prime}} = \alpha \ket {e_{1}}\ket {g_{2}}
1540: -\beta \ket {g_{1}}\ket {e_{2}} \ ,\nonumber \\
1541: && E_{e} = \hbar \omega_{0} \ ,\qquad
1542: \ket e = \ket {e_{1}}\ket {e_{2}} \ ,\label{t77}
1543: \end{eqnarray}
1544: where
1545: \begin{equation}
1546: \alpha = \frac{d}{\sqrt{d^{2}+\Omega_{12}^{2}}} \ ,\quad
1547: \beta = \frac{\Omega_{12}}{\sqrt{d^{2}+\Omega_{12}^{2}}} \ ,\quad
1548: w= \sqrt{\Omega_{12}^{2}+\Delta^{2}} \ ,\label{t78}
1549: \end{equation}
1550: and $d = \Delta
1551: +\sqrt{\Omega_{12}^{2}+\Delta^{2}}$.
1552:
1553: The energy level structure of the collective system of two nonidentical
1554: atoms is similar to that of the
1555: identical atoms, with the ground state $\ket g$, the upper state
1556: $\ket e$, and two intermediate states $\ket {s^{\prime}}$ and
1557: $\ket {a^{\prime}}$. The effect of the frequency
1558: difference $\Delta$ on the collective atomic states
1559: is to increase the splitting between the intermediate levels, which
1560: now is equal to $w=\sqrt{\Omega_{12}^{2}+\Delta^{2}}$.
1561: However, the most dramatic effect of the detuning $\Delta$ is on the
1562: degree of entanglement of the intermediate
1563: states $\ket {s^{\prime}}$ and $\ket {a^{\prime}}$ that in the case
1564: of nonidentical atoms the states are no longer maximally entangled states.
1565: For $\Delta =0$ the states $\ket {s^{\prime}}$
1566: and $\ket {a^{\prime}}$ reduce to the maximally entangled states $\ket
1567: s$ and $\ket a$, whereas for $\Delta \gg \Omega_{12}$ the entangled
1568: states reduce to the
1569: product states $\ket {e_{1}}\ket {g_{2}}$ and $-\ket {g_{1}}\ket
1570: {e_{2}}$, respectively.
1571:
1572: Using the same procedure as for the case of identical atoms, we
1573: rewrite the master equation~(\ref{t42}) in terms of the collective
1574: operators $A_{ij}=\ket i\bra j$, where now the collective states $\ket
1575: i$ are given in Eq.~(\ref{t77}). First, we find that in the case of
1576: nonidentical atoms the atomic dipole operators can be written in
1577: terms of the linear combinations of the collective operators as
1578: \begin{eqnarray}
1579: S_{1}^{+} &=& \alpha A_{es^{\prime}}-\beta A_{ea^{\prime}}+\beta
1580: A_{s^{\prime}g}+\alpha A_{a^{\prime}g} \ ,\nonumber \\
1581: S_{2}^{+} &=& \beta A_{es^{\prime}}+\alpha A_{ea^{\prime}}+\alpha
1582: A_{s^{\prime}g}-\beta A_{a^{\prime}g} \ .\label{t79}
1583: \end{eqnarray}
1584: Hence, in terms of the collective operators $A_{ij}$, the master equation
1585: takes the form
1586: \begin{equation}
1587: \frac{\partial}{\partial t}\hat{\rho}
1588: =-\frac{i}{\hbar}\left[\hat{H}_{s^{\prime}},\hat{\rho}
1589: \right] +{\cal{L}}\hat{\rho} \ ,\label{t80}
1590: \end{equation}
1591: where
1592: \begin{eqnarray}
1593: \hat{H}_{s^{\prime}} &=& \hbar\left[\omega_{0}\left(A_{ee}
1594: -A_{gg}\right)
1595: +w\left(A_{s^{\prime}s^{\prime}}- A_{a^{\prime}a^{\prime}}\right)\right]
1596: \nonumber \\
1597: &-& \frac{\hbar}{2} \left\{\left[\left(\alpha \Omega_{1} +\beta
1598: \Omega_{2}\right)A_{es^{\prime}}+
1599: \left(\beta \Omega_{1} +\alpha
1600: \Omega_{2}\right)A_{s^{\prime}g}\right]
1601: e^{i\left(\omega_{L}t+\phi_{L}\right)}\right. \nonumber \\
1602: &+& \left. \left[\left(\alpha \Omega_{2} -\beta
1603: \Omega_{1}\right)A_{ea^{\prime}}
1604: -\left(\beta \Omega_{2} -\alpha
1605: \Omega_{1}\right)A_{a^{\prime}g}\right]
1606: e^{i\left(\omega_{L}t+\phi_{L}\right)}+{\rm H.c.}\right\}
1607: \label{t81}
1608: \end{eqnarray}
1609: is the Hamiltonian of the system in the collective states basis, and
1610: the Liouville operator
1611: ${\cal{L}}\hat{\rho}$ describes the dissipative part of the
1612: evolution. The dissipative part is composed of three terms
1613: \begin{eqnarray}
1614: {\cal{L}}\hat{\rho} &=&
1615: \left(\frac{\partial}{\partial t}\hat{\rho}\right)_{s}
1616: +\left(\frac{\partial}{\partial t}\hat{\rho}\right)_{a}
1617: +\left(\frac{\partial}{\partial t}\hat{\rho}\right)_{I} \
1618: ,\label{t82}
1619: \end{eqnarray}
1620: where
1621: \begin{eqnarray}
1622: \left(\frac{\partial}{\partial t}\hat{\rho}\right)_{s} &=&
1623: -\Gamma_{s^{\prime}}\left\{\left(A_{ee}
1624: +A_{s^{\prime}s^{\prime}}\right)\hat{\rho} +\hat{\rho}\left(
1625: A_{ee}+A_{s^{\prime}s^{\prime}}\right)\right. \nonumber \\
1626: &&\left. -2\left(A_{s^{\prime}e}\hat{\rho} A_{es^{\prime}}
1627: +A_{gs^{\prime}}\hat{\rho} A_{s^{\prime}g}\right)\right\} \nonumber \\
1628: && -\left(\alpha \beta
1629: \Gamma +\Gamma_{12}\right)
1630: \left(A_{s^{\prime}e}\rho A_{s^{\prime}g} + A_{gs^{\prime}}\rho
1631: A_{es^{\prime}}\right) \ ,\label{t83}\\
1632: \left(\frac{\partial}{\partial t}\hat{\rho}\right)_{a} &=&
1633: -\Gamma_{a^{\prime}}\left\{\left(A_{ee}
1634: +A_{a^{\prime}a^{\prime}}\right)\hat{\rho} +\hat{\rho}\left(
1635: A_{ee}+A_{a^{\prime}a^{\prime}}\right)\right. \nonumber \\
1636: &&\left. -2\left(A_{a^{\prime}e}\hat{\rho} A_{ea^{\prime}}
1637: +A_{ga^{\prime}}\hat{\rho} A_{a^{\prime}g}\right)\right\} \nonumber \\
1638: && -\left(\alpha \beta
1639: \Gamma -\Gamma_{12}\right)
1640: \left(A_{a^{\prime}e}\hat{\rho} A_{a^{\prime}g}
1641: + A_{ga^{\prime}}\hat{\rho}A_{ea^{\prime}}\right) \ ,\label{t84}
1642: \end{eqnarray}
1643: and
1644: \begin{eqnarray}
1645: \left(\frac{\partial}{\partial t}\hat{\rho}\right)_{I} &=&
1646: -\Gamma_{a^{\prime}s^{\prime}}\left\{\left(
1647: A_{a^{\prime}s^{\prime}}+A_{s^{\prime}a^{\prime}}\right)\hat{\rho} +
1648: \hat{\rho} \left(A_{a^{\prime}s^{\prime}}+A_{s^{\prime}a^{\prime}}
1649: \right)\right. \nonumber \\
1650: &&-\left. 2\left(A_{ga^{\prime}}\hat{\rho} A_{s^{\prime}g}
1651: +A_{gs^{\prime}}\hat{\rho} A_{a^{\prime}g}+A_{s^{\prime}e}\hat{\rho}
1652: A_{ea^{\prime}} + A_{a^{\prime}e}\hat{\rho}
1653: A_{es^{\prime}}\right) \right\} \nonumber \\
1654: &&+ \left(\alpha^{2}-\beta^{2}\right)\Gamma \left\{
1655: A_{a^{\prime}e}\hat{\rho} A_{s^{\prime}g} + A_{gs^{\prime}}\hat{\rho}
1656: A_{ea^{\prime}}\right. \nonumber \\
1657: &&\left. +A_{s^{\prime}e}\hat{\rho} A_{a^{\prime}g} + A_{ga^{\prime}}
1658: \hat{\rho} A_{es^{\prime}}\right\} \ ,\label{t85}
1659: \end{eqnarray}
1660: with the damping coefficients
1661: \begin{eqnarray}
1662: \Gamma_{s^{\prime}} &=& \frac{1}{2}\left(\Gamma
1663: +2\alpha \beta \Gamma_{12}\right) \ ,\qquad
1664: \Gamma_{a^{\prime}} = \frac{1}{2}\left(\Gamma
1665: -2\alpha \beta \Gamma_{12}\right) \ ,\nonumber \\
1666: \Gamma_{a^{\prime}s^{\prime}} &=&
1667: \frac{1}{2}\left(\alpha^{2}-\beta^{2}\right)\Gamma_{12} \
1668: .\label{t86}
1669: \end{eqnarray}
1670:
1671: The dissipative part of the master equation is very extensive and
1672: unlike the case of identical atoms, contains the interference term
1673: between the symmetric and antisymmetric transitions.
1674: The terms~(\ref{t83}) and (\ref{t84}) describes spontaneous
1675: transitions in the symmetric and antisymmetric channels, respectively.
1676: The coefficients $\Gamma_{s^{\prime}}$, and $\Gamma_{a^{\prime}}$
1677: are the spontaneous emission rates of the transitions. The
1678: interference term~(\ref{t85}) results from spontaneously induced
1679: coherences between the symmetric and antisymmetric transitions.
1680: This term appears only in systems of atoms with different transition
1681: frequencies $(\Delta \neq 0)$, and reflects the fact that, as the
1682: system decays from the state $\ket {s^{\prime}}$, it drives the
1683: antisymmetric state, and vice versa. Thus, in contrast to the case of
1684: identical atoms, the symmetric and antisymmetric transitions are no longer
1685: independent and are correlated due to the presence of the detuning $\Delta$.
1686: Moreover, for nonidentical atoms the damping rate of the antisymmetric state
1687: cannot be reduced to zero. In the case of interatomic separations much
1688: smaller than the optical wavelength (the small sample model), the damping
1689: rate reduces to
1690: \begin{equation}
1691: \Gamma_{a^{\prime}} = \frac{1}{2}\Gamma \left(\alpha -\beta
1692: \right)^{2} \ ,\label{t87}
1693: \end{equation}
1694: that is different from zero, unless $\Delta =0$.
1695: %
1696: \begin{figure}[t]
1697: \begin{center}
1698: \includegraphics[width=10cm]{ftfig3.eps}
1699: \end{center}
1700: \caption{The spontaneous emission damping rate $\Gamma_{a^{\prime}}$
1701: as a function of $\Delta$ for $\vec{\mu}\perp \bar{r}_{12}$, and
1702: different interatomic separations: $r_{12}/\lambda =0.05$ (solid line),
1703: $r_{12}/\lambda =0.1$ (dashed line), $r_{12}/\lambda =0.5$
1704: (dashed-dotted line).}
1705: \label{ftfig3}
1706: \end{figure}
1707:
1708: In Fig.~\ref{ftfig3}, we plot the damping rate $\Gamma_{a^{\prime}}$
1709: as a function
1710: of $\Delta$ for different interatomic separations. The damping rate
1711: vanishes for $\Delta =0$ independent of the interatomic separation,
1712: but for small interatomic separations there is a significant range of
1713: $\Delta$ for which $\Gamma_{a^{\prime}}\ll \Gamma$.
1714:
1715:
1716:
1717: \subsubsection{The case $\Delta =0$ and $\Gamma_{1}\neq
1718: \Gamma_{2}$}\label{sec322}
1719:
1720:
1721: The choice of the collective states~(\ref{t77}) as a basis leads to a
1722: complicated dissipative part of the master equation. A different
1723: choice of collective states is proposed here, which allows to
1724: obtain a simple master equation of the system with only the uncorrelated
1725: dissipative parts of the symmetric and antisymmetric
1726: transitions~\cite{afs}. Moreover, we will show that it is possible to
1727: create an entangled state in the system of two nonidentical atoms which
1728: can be decoupled from the external environment and, at the same time,
1729: the state exhibits a strong coherent coupling with the remaining states.
1730:
1731: To illustrate this, we introduce superposition operator
1732: $S_{s}^{\pm}$ and $S_{a}^{\pm}$ which are linear combinations of the
1733: atomic dipole operators $S_{1}^{\pm}$ and $S_{2}^{\pm}$ as
1734: \begin{eqnarray}
1735: S_{s}^{+} &=& uS_{1}^{+}+vS_{2}^{+} \ ,\quad S_{s}^{-} =
1736: u^{\ast}S_{1}^{-} + v^{\ast}S_{2}^{-} \ ,\nonumber \\
1737: S_{a}^{+} &=& vS_{1}^{+} -u S_{2}^{+} \ ,\quad S_{a}^{-} =
1738: v^{\ast}S_{1}^{-} - u^{\ast}S_{2}^{-} \ ,\label{t88}
1739: \end{eqnarray}
1740: where $u$ and $v$ are the transformation coefficients which are in
1741: general complex numbers. The coefficients satisfy the condition
1742: \begin{equation}
1743: \left|u\right|^{2} +\left|v\right|^{2} =1 \ .\label{t89}
1744: \end{equation}
1745:
1746: The operators $S_{s}^{\pm}$ and $S_{a}^{\pm}$ represent,
1747: respectively, symmetric and antisymmetric superpositions of the
1748: atomic dipole operators.
1749: In terms of the superposition operators, the dissipative
1750: part of the master equation~(\ref{t42}) can be written as
1751: \begin{eqnarray}
1752: {\cal {L}}\hat{\rho} &=& -\Gamma_{ss}\left(S^{+}_{s}S^{-}_{s}\hat{\rho}
1753: +\hat{\rho}
1754: S^{+}_{s}S^{-}_{s}-2S^{-}_{s}\hat{\rho} S^{+}_{s} \right) \nonumber \\
1755: &&- \Gamma_{aa}\left(S^{+}_{a}S^{-}_{a}\hat{\rho} +\hat{\rho}
1756: S^{+}_{a}S^{-}_{a}-2S^{-}_{a}\hat{\rho} S^{+}_{a} \right) \nonumber \\
1757: && -\Gamma_{sa}\left(S^{+}_{s}S^{-}_{a}\hat{\rho} +\hat{\rho}
1758: S^{+}_{s}S^{-}_{a} -2S^{-}_{a}\hat{\rho} S^{+}_{s} \right) \nonumber \\
1759: && -\Gamma_{as}\left(S^{+}_{a}S^{-}_{s}\hat{\rho} +\hat{\rho}
1760: S^{+}_{a}S^{-}_{s} -2S^{-}_{s}\hat{\rho} S^{+}_{a} \right) \ ,
1761: \label{t90}
1762: \end{eqnarray}
1763: where the coefficients $\Gamma_{mn}$ are
1764: \begin{eqnarray}
1765: \Gamma_{ss} &=& \left|u\right|^{2}\Gamma_{1}
1766: +\left|v\right|^{2}\Gamma_{2} +\left(uv^{\ast}
1767: +u^{\ast}v\right)\Gamma_{12} ,\nonumber \\
1768: \Gamma_{aa} &=& \left|v\right|^{2}\Gamma_{1}
1769: +\left|u\right|^{2}\Gamma_{2} -\left(uv^{\ast}
1770: +u^{\ast}v\right)\Gamma_{12} ,\nonumber \\
1771: \Gamma_{as} &=& uv^{\ast}\Gamma_{1} -u^{\ast}v\Gamma_{2}
1772: -\left(\left|u\right|^{2} -\left|v\right|^{2}\right)\Gamma_{12}
1773: ,\nonumber \\
1774: \Gamma_{sa} &=& u^{\ast}v\Gamma_{1} -uv^{\ast}\Gamma_{2}
1775: -\left(\left|u\right|^{2} -\left|v\right|^{2}\right)\Gamma_{12}
1776: .\label{t91}
1777: \end{eqnarray}
1778: The first two terms in Eq.~(\ref{t90}) are familiar spontaneous
1779: emission terms of the symmetric and antisymmetric transitions, and
1780: the parameters $\Gamma_{ss}$ and $\Gamma_{aa}$ are spontaneous emission
1781: rates of the transitions,
1782: respectively. The last two terms are due to coherence between the
1783: superposition states and the parameters
1784: $\Gamma_{as}$ and $\Gamma_{sa}$ describes cross-damping rates between
1785: the superpositions.
1786:
1787: If we make the identification
1788: \begin{equation}
1789: u = \sqrt{\frac{\Gamma_{1}}{\Gamma_{1}+\Gamma_{2}}} \ ,\quad
1790: v = \sqrt{\frac{\Gamma_{2}}{\Gamma_{1}+\Gamma_{2}}} \ ,\label{t92}
1791: \end{equation}
1792: then the damping coefficients~(\ref{t91}) simplify to
1793: \begin{eqnarray}
1794: \Gamma_{ss} &=& \frac{1}{2}\left(\Gamma_{1}+\Gamma_{2}\right)
1795: +\frac{\sqrt{\Gamma_{1}\Gamma_{2}}\left(\Gamma_{12}
1796: -\sqrt{\Gamma_{1}\Gamma_{2}}\right)}{\Gamma_{1}+\Gamma_{2}} \ , \nonumber \\
1797: \Gamma_{aa} &=& \frac{\left(\sqrt{\Gamma_{1}\Gamma_{2}}-\Gamma_{12}\right)
1798: \sqrt{\Gamma_{1}\Gamma_{2}}}{\Gamma_{1}+\Gamma_{2}} \ , \nonumber \\
1799: \Gamma_{sa} &=& \Gamma_{as} = \frac{1}{2} \frac{\left(\Gamma_{1}
1800: -\Gamma_{2}\right)\left(\sqrt{\Gamma_{1}\Gamma_{2}} -\Gamma_{12}\right)}
1801: {\Gamma_{1}+\Gamma_{2}} \ .\label{t93}
1802: \end{eqnarray}
1803: When the damping rates of the atoms are equal $(\Gamma_{1}=\Gamma_{2})$,
1804: the cross-damping terms $\Gamma_{as}$ and $\Gamma_{sa}$ vanish. Furthermore,
1805: if $\Gamma_{12}=\sqrt{\Gamma_{1}\Gamma_{2}}$ then the spontaneous emission
1806: rates $\Gamma_{aa}$, $\Gamma_{as}$ and $\Gamma_{sa}$ vanish regardless of
1807: the ratio between the $\Gamma_{1}$ and $\Gamma_{2}$. In this case,
1808: which corresponds to interatomic separations much
1809: smaller than the optical wavelength, the antisymmetric superposition
1810: does not decay and also decouples from the symmetric superposition.
1811:
1812: An interesting question arises as to whether the nondecaying
1813: antisymmetric superposition can still be coupled to the symmetric
1814: superposition through the coherent interactions $\Omega_{12}$ and
1815: $\Omega$ contained in the Hamiltonian $\hat{H}_{s}$. These interactions
1816: can coherently transfer population between the superpositions.
1817: To check it, we first transform the
1818: Hamiltonian~(\ref{t43}) into the interaction picture and next rewrite
1819: the transformed Hamiltonian in terms of the $S_{s}^{\pm}$ and $S_{a}^{\pm}$
1820: operators as
1821: \begin{eqnarray}
1822: \hat{H}_{s} &=&-\hbar \Delta_{L}\left[\left(
1823: S_{s}^{+}S_{s}^{-}+S_{a}^{+}S_{a}^{-}\right)
1824: +\left(v^{\ast}u-vu^{\ast}\right)\left(
1825: S_{s}^{+}S_{a}^{-}-S_{a}^{+}S_{s}^{-}\right)\right] \nonumber \\
1826: &&+\hbar \Omega_{12}\left\{\left(vu^{\ast}+v^{\ast}u\right)
1827: \left( S_{s}^{+}S_{s}^{-}-S_{a}^{+}S_{a}^{-}
1828: \right)\right. \nonumber \\
1829: &&\left. +\left(|v|^{2}-|u|^{2}\right) \left(
1830: S_{s}^{+}S_{a}^{-}+S_{a}^{+}S_{s}^{-}\right) \right\} \nonumber \\
1831: &&-\frac{1}{2}\hbar \left[ \left( u\Omega _{1}+v\Omega _{2}\right)
1832: S_{s}^{+}
1833: +\left( v\Omega _{1}-u\Omega _{2}\right)S_{a}^{+}
1834: +{\rm H.c.}\right] \ ,\label{t94}
1835: \end{eqnarray}
1836: where $\Delta_{L}=\omega_{L}-\omega_{0}$.
1837:
1838: In the above equation, the first term arises from the atomic Hamiltonian and
1839: shows that in the absence of the interatomic interactions the symmetric and
1840: antisymmetric states have the same energy.
1841: The second term in Eq.~(\ref{t94}), proportional to the dipole-dipole
1842: interaction between the atoms, has two
1843: effects on the dynamics of the symmetric and antisymmetric
1844: superpositions. The first is a shift of the energies and the second is the
1845: coherent interaction between the superpositions. It is seen from
1846: Eq.~(\ref{t94}) that the contribution of $\Omega_{12}$ to the coherent
1847: interaction between the superpositions vanishes for $\Gamma_{1}=\Gamma_{2}$
1848: and then the effect of $\Omega_{12}$ is only the shift of the energies from
1849: their unperturbed values. Note that the dipole-dipole interaction
1850: $\Omega_{12}$ shifts the energies in the opposite directions.
1851: The third term in Eq.~(\ref{t94}) represents the interaction of the
1852: superpositions with the driving laser field. We see that the symmetric
1853: superposition couples to the laser field with an effective Rabi
1854: frequency proportional to $u\Omega _{1}+v\Omega _{2}$, whereas the Rabi
1855: frequency of the antisymmetric superposition is proportional to $v\Omega
1856: _{1}-u\Omega _{2}$ and vanishes for $v\Omega _{1}=u\Omega _{2}$.
1857:
1858: Alternatively, we may write the Hamiltonian (\ref{t94}) in a more
1859: transparent form which shows explicitly the presence of the coherent
1860: coupling between the symmetric and antisymmetric states
1861: \begin{eqnarray}
1862: \hat{H}_{s} &=& -\hbar \left[\left(\Delta_{L}-\Delta^{\prime}\right)
1863: S_{s}^{+}S_{s}^{-} + \left(\Delta_{L}+\Delta^{\prime}
1864: \right)S_{a}^{+}S_{a}^{-} +\Delta_{c}S_{s}^{+}S_{a}^{-}
1865: +\Delta_{c}^{\ast}S_{a}^{+}S_{s}^{-}\right] \nonumber \\
1866: &&- \frac{1}{2}\hbar \left[\left(u\Omega_{1}+v\Omega_{2}\right)
1867: S^{+}_{s}
1868: + \left(v\Omega_{1}-u\Omega_{2}\right) S^{+}_{a}
1869: +{\rm H.c.}\right] \ , \label{t95}
1870: \end{eqnarray}
1871: where $\Delta^{\prime}$ and $\Delta_{c}$ are given by
1872: \begin{eqnarray}
1873: \Delta^{\prime} = \left(vu^{\ast}+v^{\ast}u\right)\Omega_{12} \ ,\quad
1874: \Delta_{c} = \left(|u|^{2}-|v|^{2}\right)\Omega_{12}
1875: + \left(v^{\ast}u-vu^{\ast}\right)\Delta_{L} \ .\label{t96}
1876: \end{eqnarray}
1877:
1878: The parameters $\Delta^{\prime}$ and $\Delta_{c}$ allow us to gain
1879: physical insight into how the dipole-dipole interaction $\Omega_{12}$
1880: and the unequal damping rates $\Gamma_{1}\neq \Gamma_{2}$ can modify the
1881: dynamics of the
1882: two-atom system. The parameter $\Delta^{\prime}$ appears as a shift of the
1883: energies of the superposition systems, while
1884: $\Delta_{c}$ determines the magnitude of the coherent interaction between the
1885: superpositions. For identical atoms the shift $%
1886: \Delta^{\prime}$ reduces to $\Omega_{12}$ that is the dipole-dipole
1887: interaction shift of the energy levels.
1888: In contrast to the shift $\Delta^{\prime}$, which is different from zero for
1889: identical as well as nonidentical atoms, the coherent coupling $\Delta_{c}$
1890: can be different from zero only for nonidentical atoms.
1891:
1892: Thus, the condition $\Gamma_{12}=\sqrt{\Gamma_{1}\Gamma_{2}}$ for
1893: suppression of spontaneous emission from the antisymmetric state is valid
1894: for identical as well as non-identical atoms, whereas the coherent interaction
1895: between the superpositions appears only for nonidentical atoms with different
1896: spontaneous damping rates.
1897:
1898: It should be noted that this treatment is valid with only a minor
1899: modification for a number of other schemes of two-atom systems. For
1900: example, it can be applied to the case of two identical atoms that
1901: experience different intensities and phases of the driving
1902: field~\cite{fs,rfd,lm93}.
1903:
1904: In what follows, we will illustrate how the interference term in the
1905: master equation of two nonidentical atoms results in quantum beats
1906: and transfers of the population to the antisymmetric state even if
1907: the antisymmetric state does not decay. Of particular interest is the
1908: temporal dependence of the total radiation intensity of the
1909: fluorescence field emitted by two interacting atoms.
1910:
1911:
1912: \section{Quantum beats}\label{sec4}
1913:
1914:
1915: The objective of this section is to give an account of interference
1916: effects resulting from the direct correlations between the symmetric
1917: and antisymmetric states. We will first analyse the simplest model
1918: of spontaneous emission from two nonidentical atoms and consider the
1919: time dependence of the total radiation intensity. After this, we will
1920: consider the time evolution of the fluorescence intensity emitted by
1921: two identical atoms that are not in the equivalent positions in the
1922: driving field.
1923:
1924:
1925: \subsection{Quantum beats in spontaneous emission from two
1926: nonidentical atoms}\label{sec41}
1927:
1928: For two nonidentical atoms the master equation~(\ref{t42}), in the
1929: absence of the driving field $(\Omega_{i}=0)$, leads to a closed set
1930: of five equations of motion for the expectation values of the atomic
1931: dipole operators~\cite{ftk86}. This set of equations can be written
1932: in a matrix form as
1933: \begin{eqnarray}
1934: \frac{d}{dt}\vec{X}\left(t\right) = A\vec{X}\left(t\right) \
1935: ,\label{t97}
1936: \end{eqnarray}
1937: where $\vec{X}\left(t\right)$ is a column vector with components
1938: \begin{eqnarray}
1939: X_{1} &=& \langle S_{1}^{+}(t)S_{1}^{-}(t)\rangle \ ,\quad
1940: X_{2} = \langle S_{2}^{+}(t)S_{2}^{-}(t)\rangle \ ,\nonumber \\
1941: X_{3} &=& \langle S_{1}^{+}(t)S_{2}^{-}(t)\rangle \ ,\quad
1942: X_{4} = \langle S_{2}^{+}(t)S_{1}^{-}(t)\rangle \ ,\nonumber \\
1943: X_{5} &=& \langle S_{1}^{+}(t)S_{2}^{+}(t)S_{1}^{-}(t)
1944: S_{2}^{-}(t)\rangle \ ,\label{t98}
1945: \end{eqnarray}
1946: and $A$ is the $5\times 5$ matrix
1947: \begin{eqnarray}
1948: A &=& \left(
1949: \begin{array}{ccccc}
1950: -\Gamma_{1} & 0 & \kappa &
1951: \kappa^{\ast} & 0\\
1952: 0 & -\Gamma_{2} & \kappa^{\ast} &
1953: \kappa & 0 \\
1954: \kappa & \kappa^{\ast} & -\frac{1}{2}(\Gamma_{T}-4i\Delta)
1955: & 0 & 2\Gamma_{12}\\
1956: \kappa^{\ast} &
1957: \kappa & 0 &
1958: -\frac{1}{2}(\Gamma_{T}+4i\Delta) & 2\Gamma_{12}\\
1959: 0 & 0 & 0 & 0 & -\Gamma_{T}
1960: \end{array}
1961: \right) \ , \label{t99}
1962: \end{eqnarray}
1963: with $\kappa =-\frac{1}{2}(\Gamma_{12}+i\Omega_{12})$ and
1964: $\Gamma_{T}=\Gamma_{1}+\Gamma_{2}$.
1965:
1966: It is seen from Eq.~(\ref{t99}) that the equation of motion for the
1967: second-order correlation function $\langle S_{1}^{+}(t)S_{2}^{+}(t)
1968: S_{1}^{-}(t)S_{2}^{-}(t)\rangle$ is decoupled from the remaining four
1969: equations. This allows for an exact solution of the set of
1970: equations~(\ref{t96}). The exact solution is given in
1971: Ref.~\cite{ftk86}. Here, we will focus on two special cases of
1972: $\Delta \neq 0, \Gamma_{1}=\Gamma_{2}$ and $\Delta =0, \Gamma_{1}\neq
1973: \Gamma_{2}$, and calculate the time evolution of the total
1974: fluorescence intensity, defined in Eq.~(\ref{t71}). We will assume
1975: that initially $(t=0)$ atom "1" was in its excited state $\ket {e_{1}}$
1976: and atom "2" was in its ground state $\ket {g_{2}}$.
1977:
1978:
1979: \subsubsection{The case $\Delta \neq 0$, $\Gamma_{1}=\Gamma_{2}=\Gamma$
1980: and $\Omega_{12}\gg \Delta$}\label{sec411}
1981:
1982:
1983: In this case the atoms have the same spontaneous damping rates but
1984: different transition frequencies that, for simplicity, are taken much
1985: smaller than the dipole-dipole interaction potential. In this limit,
1986: the approximate solution of Eq.~(\ref{t97}) leads to the following
1987: total radiation intensity
1988: \begin{eqnarray}
1989: I\left(t\right) = e^{-\Gamma
1990: t}\left[\frac{\Delta}{2\Omega_{12}}\Gamma_{12}\cos 2wt
1991: +\Gamma \cosh \Gamma_{12}t -\Gamma_{12}\sinh \Gamma_{12}t
1992: \right] \ ,\label{t100}
1993: \end{eqnarray}
1994: where $w=\sqrt{\Omega_{12}^{2}+\Delta^{2}}$.
1995:
1996: The total radiation intensity exhibits sinusoidal modulation (beats)
1997: superimposed on exponential decay with the damping rates $\Gamma \pm
1998: \Gamma_{12}$. The amplitude of the oscillations is proportional to
1999: $\Delta$ and vanishes for identical atoms. The damping rate $\Gamma
2000: +\Gamma_{12}$ describes the spontaneous decay from the state
2001: $\ket {s^{\prime}}$ to the ground state $\ket g$, while $\Gamma -
2002: \Gamma_{12}$ is the decay rate of the
2003: $\ket {a^{\prime}} \rightarrow \ket g$ transition. The frequency
2004: $2w$ of the oscillations is equal to the frequency difference between
2005: the $\ket {s^{\prime}}$ and $\ket {a^{\prime}}$ states. The
2006: oscillations reflect the spontaneously induced correlations between
2007: the $\ket {s^{\prime}} \rightarrow \ket g$ and
2008: $\ket {a^{\prime}} \rightarrow \ket g$ transitions. According to
2009: Eq.~(\ref{t86}) the amplitude of the spontaneously induced
2010: correlations is equal to $\Gamma_{a^{\prime}s^{\prime}}$, which in
2011: the limit of $\Omega_{12}\gg \Delta$ reduces to
2012: $\Gamma_{a^{\prime}s^{\prime}}= \Delta \Gamma_{12}/(2\Omega_{12})$.
2013: Hence, the amplitude of the oscillations appearing in Eq.~(\ref{t100})
2014: is exactly equal to the amplitude of the spontaneously induced
2015: correlations. Fig.~\ref{ftfig4} shows the temporal dependence of the
2016: total radiation intensity
2017: for interatomic separation $r_{12}=\lambda/12, \Gamma_{1}=\Gamma_{2},
2018: \bar{\mu}\perp \bar{r}$, and different $\Delta$. As predicted by
2019: Eq.~(\ref{t100}), the intensity exhibits quantum beats whose the
2020: amplitude increases with increasing $\Delta$. Moreover, at short
2021: times, the intensity can become greater than its initial value
2022: $I\left(0\right)$. This effect is known as a superradiant
2023: behavior and is absent in the case of two identical atoms.
2024: Thus, the spontaneously induced correlations between the
2025: $\ket {s^{\prime}} \rightarrow \ket g$ and $\ket {a^{\prime}}
2026: \rightarrow \ket g$ transitions can induce quantum beats and
2027: superradiant effect in the intensity of the emitted field.
2028:
2029: The superradiant effect is characteristic of a large number of
2030: atoms~\cite{bl,gh,er}, and it is quite surprising to obtain this
2031: effect in the system of two atoms. Coffey and Friedberg~\cite{cf78}
2032: and Richter~\cite{rich81} have shown that the superradiant effect can
2033: be observed in some special cases of the atomic configuration of a
2034: three-atom system. Blank {\it et al.}~\cite{bla} have shown that this
2035: effect, for atoms located in an equidistant linear chain, appears for
2036: at least six atoms. Recently, DeAngelis {\it et al.}~\cite{dea} have
2037: experimentally observed the superradiant effect in the radiation
2038: from two identical dipoles located inside a planar symmetrical
2039: microcavity.
2040: %
2041: \begin{figure}[t]
2042: \begin{center}
2043: \includegraphics[width=10cm]{ftfig4.eps}
2044: \end{center}
2045: \caption{Time evolution of the total radiation intensity for
2046: $r_{12}=\lambda/12, \Gamma_{1}=\Gamma_{2},
2047: \bar{\mu}\perp \bar{r}$, and different $\Delta$: $\Delta =0$ (solid
2048: line), $\Delta = -2\Gamma$ (dashed line), $\Delta =-3\Gamma$
2049: (dashed-dotted line).}
2050: \label{ftfig4}
2051: \end{figure}
2052:
2053: Quantum beats predicted here for spontaneous emission from two
2054: nonidentical atoms are fully equivalent to the quantum beats
2055: predicted recently by Zhou and Swain~\cite{zs98} in a single
2056: three-level $V$ system with correlated spontaneous transitions. For
2057: the initial conditions used here that initially only one of the atoms
2058: was excited, the initial population distributes equally between the
2059: states $\ket {s^{\prime}}$ and $\ket {a^{\prime}}$. Since the
2060: transitions are correlated through the dissipative term
2061: $\Gamma_{a^{\prime}s^{\prime}}$, the system of two nonidentical atoms
2062: behaves as a three-level $V$ system with spontaneously correlated
2063: transitions.
2064:
2065:
2066: \subsubsection{The case of $\Delta =0, \Gamma_{1}\neq \Gamma_{2}$
2067: and $\Omega_{12}\gg \Gamma_{1},\Gamma_{2}$}\label{sec412}
2068:
2069:
2070: We now wish to show how quantum beats can be obtained in two
2071: nonidentical atoms that have equal frequencies but different damping
2072: rates. According to Eqs.~(\ref{t93}) and (\ref{t96}), the symmetric
2073: and antisymmetric transitions are correlated not only through the
2074: spontaneously induced coherences $\Gamma_{as}$, but also through the
2075: coherent coupling $\Delta_{c}$. One can see from Eq.~(\ref{t93}) that
2076: for small interatomic separations $\Gamma_{as}\approx 0$. However, the
2077: coherent coupling parameter $\Delta_{c}$, which is proportional to
2078: $\Omega_{12}$, is very large, and we will
2079: show that the coherent coupling $\Delta_{c}$ can also lead to quantum
2080: beats and the superradiant effect. In the case of $\Delta =0,
2081: \Gamma_{1}\neq \Gamma_{2}$ and $\Omega_{12}\gg \Gamma_{1},\Gamma_{2}$,
2082: the approximate solution of Eq.~(\ref{t97}) leads to the following
2083: expression for the total radiation intensity
2084: %
2085: \begin{eqnarray}
2086: I\left(t\right) &=&
2087: e^{-\frac{1}{2}\left(\Gamma_{1}+\Gamma_{2}\right)t}
2088: \left\{\frac{1}{2}\left(\Gamma_{1}-\Gamma_{2}\right)\cos
2089: 2\Omega_{12}t \right. \nonumber \\
2090: &&\left. +\frac{1}{2}\left(\Gamma_{1}+\Gamma_{2}\right) \cosh
2091: \Gamma_{12}t -\Gamma_{12}\sinh \Gamma_{12}t
2092: \right\} \ .\label{t101}
2093: \end{eqnarray}
2094: The intensity displays quantum-beat oscillations at frequency
2095: $2\Omega_{12}$ corresponding to the frequency splitting between the
2096: $\ket {s^{\prime}}$ and $\ket {a^{\prime}}$ states. The amplitude of
2097: the oscillations is equal to $\left(\Gamma_{1}-\Gamma_{2}\right)/2$
2098: that is proportional to the coherent coupling $\Delta_{c}$. For
2099: $\Gamma_{1}=\Gamma_{2}$ the coherent coupling parameter $\Delta_{c}=0$
2100: and no quantum beats occur. In this case the intensity exhibits pure
2101: exponential decay. This is shown in Fig.~\ref{ftfig5}, where we plot the
2102: time evolution of $I\left(t\right)$ for interatomic separation
2103: $r_{12}=\lambda/12$, and different ratios
2104: $\Gamma_{2}/\Gamma_{1}$. Similar to the case discussed in
2105: Sec.~\ref{sec411}, the intensity exhibits quantum beats and the
2106: superradiant effect. For $r_{12}=\lambda/12$ the collective damping
2107: $\Gamma_{12}\approx \sqrt{\Gamma_{1}\Gamma_{2}}$, and then the
2108: parameter $\Gamma_{as}\approx 0$, indicating that the
2109: quantum beats and the superradiant effect result from the coherent
2110: coupling between the $\ket {s^{\prime}}$ and $\ket {a^{\prime}}$
2111: states.
2112: %
2113: \begin{figure}[t]
2114: \begin{center}
2115: \includegraphics[width=10cm]{ftfig5.eps}
2116: \end{center}
2117: \caption{Time evolution of the total radiation intensity for
2118: $r_{12}=\lambda/12, \Delta =0,
2119: \bar{\mu}\perp \bar{r}_{12}$, and different $\Gamma_{2}/\Gamma_{1}$:
2120: $\Gamma_{2}/\Gamma_{1}=1$ (solid
2121: line), $\Gamma_{2}/\Gamma_{1} = 2.5$ (dashed line),
2122: $\Gamma_{2}/\Gamma_{1} =5$ (dashed-dotted line).}
2123: \label{ftfig5}
2124: \end{figure}
2125:
2126:
2127:
2128: \subsubsection{Two identical atoms in nonequivalent positions in a
2129: driving field}\label{sec413}
2130:
2131:
2132: Quantum beats and superradiant effect induced by interference between
2133: different transitions in the system of two nonidentical atoms also occur
2134: in other situations. For example, quantum beats can appear in a system of
2135: two identical atoms that experience different amplitude or phase of a
2136: coherent driving field~\cite{fs,rfd}.
2137:
2138: Consider the Hamiltonian~(\ref{t44}) of the interaction between
2139: coherent laser field and two identical atoms. In the interaction
2140: picture, the Hamiltonian can be written as
2141: \begin{eqnarray}
2142: \hat{H}_{L} &=&
2143: -\frac{1}{2}\hbar \left[\left(\Omega_{1}S_{1}^{+}
2144: +\Omega_{2}S_{2}^{+}\right)
2145: + \rm{H.c.}\right] \ ,\label{t102}
2146: \end{eqnarray}
2147: where $\Omega_{i}$ is the Rabi frequency of the driving field at the
2148: position of the $i$th atom.
2149:
2150: For the atoms in a running-wave laser field with $\vec{k}_{L}
2151: \cdot \vec{r}_{i}\neq 0$, the Rabi frequency is a
2152: complex parameter, which may be written as
2153: \begin{eqnarray}
2154: \Omega_{i} =\Omega e^{i\vec{k}_{L}
2155: \cdot \vec{r}_{i}} \ , \label{t103}
2156: \end{eqnarray}
2157: where $\Omega =|\vec{\mu}_{i}\cdot \vec{E}_{L}|/\hbar$ is the maximum
2158: Rabi frequency and $\vec{k}_{L}$ is the wave vector of the driving
2159: field. Thus, in the running-wave laser field the atoms experience
2160: different phases of the driving field.
2161:
2162: For the atoms in a standing-wave laser field and $\vec{k}_{L}
2163: \cdot \vec{r}_{i}\neq 0$, the Rabi frequency is a
2164: real parameter, which may be written as
2165: \begin{eqnarray}
2166: \Omega_{i} =\Omega \cos\left(\vec{k}_{L}
2167: \cdot \vec{r}_{i}\right) \ . \label{t104}
2168: \end{eqnarray}
2169: Hence, in the standing-wave laser field the atoms experience
2170: different amplitudes of the driving field.
2171:
2172: In the following, we choose the reference frame such that the atoms
2173: are at the positions $\vec{r}_{1}=(r_{1},0,0)$ and
2174: $\vec{r}_{2}=(r_{2},0,0)$ along the $x$-axis, with distance $r_{12}$
2175: apart. In this case,
2176: \begin{eqnarray}
2177: \Omega_{1} =\Omega e^{i\vec{k}_{L}
2178: \cdot \vec{r}_{1}} \ ,\qquad
2179: \Omega_{2} =\Omega e^{i\vec{k}_{L}
2180: \cdot \vec{r}_{2}} \ , \label{t105}
2181: \end{eqnarray}
2182: for the atoms in the running-wave field, and
2183: \begin{eqnarray}
2184: \Omega_{1} =\Omega \cos\left(\vec{k}_{L}
2185: \cdot \vec{r}_{1}\right) \ ,\qquad
2186: \Omega_{2} =\Omega \cos\left(\vec{k}_{L}
2187: \cdot \vec{r}_{2}\right) \ , \label{t106}
2188: \end{eqnarray}
2189: for the atoms in the standing-wave field.
2190:
2191: With the above choice of the Rabi frequencies, the
2192: Hamiltonian~(\ref{t102}) takes the form
2193: \begin{eqnarray}
2194: \hat{H}_{L} &=&
2195: -\frac{1}{2}\hbar \left(\Omega S_{s}^{+}
2196: + \rm{H.c.}\right) \ ,\label{t107}
2197: \end{eqnarray}
2198: where $S_{s}^{+}= S_{1}^{+}\exp(i\vec{k}_{L}\cdot \vec{r}_{1}) +
2199: S_{2}^{+}\exp(i\vec{k}_{L}\cdot \vec{r}_{2})$ for the running-wave
2200: field, and
2201: $S_{s}^{+}= S_{1}^{+}\cos(\vec{k}_{L}\cdot \vec{r}_{1}) +
2202: S_{2}^{+}\cos(\vec{k}_{L}\cdot \vec{r}_{2})$ for the standing-wave
2203: field. The operator $S_{s}^{+}$
2204: corresponds to the symmetric superposition operator defined in
2205: Eq.(\ref{t88}). Following the procedure, we developed in
2206: Sec.~\ref{sec322}, we find that the transformation coefficients
2207: $u$ and $v$ are
2208: \begin{eqnarray}
2209: u &=& \frac{e^{i\vec{k}_{L}\cdot \vec{r}_{1}}}{\sqrt{2}} \ ,\qquad
2210: v =\frac{e^{i\vec{k}_{L}\cdot \vec{r}_{2}}}{\sqrt{2}} \
2211: ,\label{t108}
2212: \end{eqnarray}
2213: for the running-wave field, and
2214: \begin{eqnarray}
2215: u &=& \frac{\cos (\vec{k}_{L}\cdot \vec{r}_{1})}
2216: {\sqrt{\cos^{2}(\vec{k}_{L}\cdot \vec{r}_{1})
2217: +\cos^{2}(\vec{k}_{L}\cdot \vec{r}_{2})}}
2218: \ ,\nonumber \\
2219: v &=& \frac{\cos(\vec{k}_{L}\cdot \vec{r}_{2})}
2220: {\sqrt{\cos^{2}(\vec{k}_{L}\cdot \vec{r}_{1})
2221: +\cos^{2}(\vec{k}_{L}\cdot \vec{r}_{2})}} \ ,\label{t109}
2222: \end{eqnarray}
2223: for the standing-wave field.
2224:
2225: Using the transformation coefficients~(\ref{t108}) and~(\ref{t109}),
2226: we find that the spontaneously induced coherences $\Gamma_{as}$ and
2227: the coherent coupling $\Delta_{c}$ between the symmetric and
2228: antisymmetric transitions are
2229: \begin{eqnarray}
2230: \Gamma_{as} &=& -i\Gamma \sin(\vec{k}_{L}\cdot \vec{r}_{12})
2231: \ ,\quad
2232: \Delta_{c} = i\Delta_{L}\sin(\vec{k}_{L}\cdot \vec{r}_{12})
2233: \ ,\label{t110}
2234: \end{eqnarray}
2235: for the running-wave field, and
2236: \begin{eqnarray}
2237: \Gamma_{as} &=& -\Gamma_{12}\frac{\sin^{2}(\vec{k}_{L}\cdot
2238: \vec{r}_{12})}{1+\cos^{2}(\vec{k}_{L}\cdot \vec{r}_{12})}
2239: \ ,\quad
2240: \Delta_{c} = \Omega_{12}\frac{\sin^{2}(\vec{k}_{L}\cdot
2241: \vec{r}_{12})}{1+\cos^{2}(\vec{k}_{L}\cdot \vec{r}_{12})}
2242: \ ,\label{t111}
2243: \end{eqnarray}
2244: for the standing-wave field, where, for simplicity, we have chosen
2245: the reference frame such that $r_{1}=0$ and $r_{2}=r_{12}$.
2246:
2247: First, we note that no quantum beats can be obtained for the
2248: direction of propagation of the laser field perpendicular to the
2249: interatomic axis, because $\sin(\vec{k}_{L}
2250: \cdot \vec{r}_{12})= 0$; however, quantum beats occur for directions
2251: of propagation different from the perpendicular to $\vec{r}_{12}$.
2252: One can see from Eqs.~(\ref{t110}) and (\ref{t111}) that in the case
2253: of the running-wave field and $\Delta_{L}=0$, the symmetric and antisymmetric
2254: transitions are correlated only through the spontaneously induced
2255: coherences $\Gamma_{as}$. In the case of the standing-wave field, both
2256: coupling parameters $\Gamma_{as}$ and $\Delta_{c}$ are different from
2257: zero. However, for interatomic separations $r_{12}<\lambda$, the
2258: parameter $\Gamma_{as}$ is much smaller than $\Delta_{c}$, indicating
2259: that in this case the coherent coupling dominates over the
2260: spontaneously induced coherences. These simple analysis of the
2261: parameters $\Gamma_{as}$ and $\Delta_{c}$ show that one should obtain
2262: quantum beats in the total radiation intensity of the fluorescence
2263: field emitted from two identical atoms.
2264: %
2265: \begin{figure}[t]
2266: \begin{center}
2267: \includegraphics[width=10cm]{ftfig6.eps}
2268: \end{center}
2269: \caption{Time evolution of the total radiation intensity for
2270: the running-wave driving field with $\Omega =0.2\Gamma,
2271: \vec{k}_{L}\parallel \vec{r}_{12}$
2272: and different interatomic separations; $r_{12}=0.2\lambda$ (solid
2273: line), $r_{12}=0.16\lambda$ (dashed line), $r_{12}=0.14\lambda$
2274: (dashed-dotted line).}
2275: \label{ftfig6}
2276: \end{figure}
2277: %
2278: Figures~\ref{ftfig6} and~\ref{ftfig7} show the time evolution of the
2279: total radiation
2280: intensity, obtained by numerical solutions of the equations of motion
2281: for the atomic correlation functions. The equations are found from the
2282: master equation~(\ref{t42}), which in the case of the running- or
2283: standing-wave driving field leads to a closed set of fifteen equations
2284: of motion for the atomic correlation functions~\cite{fs,rfd}. In
2285: Fig.~\ref{ftfig6}, we present the time-dependent total radiation
2286: intensity for the running-wave driving field with $\Omega
2287: =0.2\Gamma, \vec{k}_{L}\parallel \vec{r}_{12}$
2288: and different interatomic separations. Fig.~\ref{ftfig7} shows the total
2289: radiation intensity for the same parameters as in Fig.~\ref{ftfig6}, but
2290: the standing-wave driving field. As predicted by Eqs.~(\ref{t110}) and
2291: (\ref{t111}), the intensity exhibits quantum beats. The amplitude and
2292: frequency of the oscillations is dependent on the interatomic
2293: interactions and vanishes for large interatomic separations as well as
2294: for separations very small compared with the resonant wavelength.
2295: This is easily explained in the framework of collective states of a
2296: two-atom system. For a weak driving field, the population oscillates
2297: between the intermediate states $\ket s$, $\ket a$ and the ground
2298: state $\ket g$. When interatomic separations are large, $\Omega_{12}$ is
2299: approximately zero, and then the transitions $\ket s \rightarrow \ket
2300: g$ and $\ket a \rightarrow \ket g$ have the same frequency. Therefore,
2301: there are no quantum beats in the emitted field. On the other hand,
2302: for very small interatomic separations, $\vec{k}_{L}\cdot
2303: \vec{r}_{12}\approx 0$, and then the coupling parameters
2304: $\Gamma_{as}$ and $\Delta_{c}$ vanish, resulting in the disappearance
2305: of the quantum beats.
2306: %
2307: \begin{figure}[t]
2308: \begin{center}
2309: \includegraphics[width=10cm]{ftfig7.eps}
2310: \end{center}
2311: \caption{Time evolution of the total radiation intensity for
2312: the same parameters as in Fig.~\ref{ftfig6}, but the standing-wave
2313: driving field.}
2314: \label{ftfig7}
2315: \end{figure}
2316:
2317:
2318:
2319: \section{Nonclassical states of light}\label{ftsec5}
2320:
2321:
2322: The interaction of light with atomic systems can lead to unique
2323: phenomena such as photon antibunching and squeezing. These effects
2324: are examples of a nonclassical light field, that is a field for which
2325: quantum mechanics is essential for its description. Photon antibunching
2326: is characteristic of a radiation in which the variance of the number of
2327: photons is less than the mean number of photons, i.e. the photons
2328: exhibit sub-Poissonian statistics. Squeezing is characteristic of a
2329: field with phase-sensitive quantum fluctuations, which in one of the
2330: two phase components are reduced below the vacuum (shot-noise) level.
2331: Since photon antibunching and squeezing are distinguishing features of
2332: light, it is clearly of interest to identify situations in which such
2333: fields can be generated. Photon antibunching has been predicted
2334: theoretically for the first time in resonance fluorescence of a
2335: two-level atom~\cite{cw76,km76}. Since then, a number of papers have
2336: appeared analyzing various schemes for generating photon antibunching
2337: offered by nonlinear optics~\cite{wal79,lo80,per80,paul82}. Squeezing
2338: has been extensively studied since the theoretical work by Walls and
2339: Zoller~\cite{wz81} and Mandel~\cite{man82} on reduction of noise and
2340: photon statistics in resonance fluorescence of a two-level atom.
2341: Several experimental groups have been successful in producing
2342: nonclassical light. Photon antibunching has been observed in resonance
2343: fluorescence from a dilute atomic beam of sodium atoms driven by a
2344: coherent laser field~\cite{kdm77,kdm78,cre}. More recently, beautiful
2345: measurements of photon antibunching have been made
2346: on trapped atoms~\cite{dw87}, and a cavity QED system~\cite{fost00}.
2347: On the other hand, squeezed light was first observed by Slusher {\it et
2348: al.}~\cite{slush} in four-wave mixing experiments. After that observation
2349: squeezed light has been observed in many other nonlinear processes,
2350: with a recent development being the availability of a tunable source
2351: of squeezed light exhibiting a noise reduction of $\sim 70\%$ below
2352: the shot-noise level. The experimental observation of photon antibunching
2353: and squeezing have provided direct evidence of the quantum nature of light,
2354: and these two phenomena were precursors of much of the present work on
2355: nonclassical light fields. An extensive literature on various aspects of
2356: photon antibunching and squeezing now exists and is reviewed in
2357: several articles~\cite{jmo,josab,dodo}.
2358:
2359: The objective of this section is to concentrate on collective two-atom
2360: systems as a potential source for photon antibunching and squeezing.
2361: We understand collective effects in a broad sense, that for
2362: two or more atoms all effects that cannot be explained by the properties
2363: of individual atoms are considered as collective. This definition of
2364: collective effects thus includes, for example, both the resonance
2365: fluorescence from a system of two atoms in free space and also collective
2366: behaviour of two atoms strongly coupled to the same cavity mode in the
2367: good cavity limit. Moreover, we emphasize the role of the
2368: interatomic interactions in the generation of nonclassical light.
2369: We also relate the nonclassical effects to the degree of entanglement
2370: in the system.
2371:
2372:
2373: \subsection{Photon antibunching}\label{ftsec51}
2374:
2375:
2376: Photon antibunching is described through the normalized second-order
2377: correlation function, defined as~\cite{mw}
2378: \begin{eqnarray}
2379: g^{(2)}\left(\vec{R}_{1},t_{1};\vec{R}_{2},t_{2}\right) =
2380: \frac{G^{(2)}\left(\vec{R}_{1},t_{1};\vec{R}_{2},t_{2}\right)}
2381: {G^{(1)}\left(\vec{R}_{1},t_{1}\right)G^{(1)}\left(\vec{R}_{2},t_{2}\right)}
2382: \ ,\label{t112}
2383: \end{eqnarray}
2384: where
2385: \begin{eqnarray}
2386: G^{(2)}&&\left(\vec{R}_{1},t_{1}; \vec{R}_{2},t_{2}\right) \nonumber
2387: \\
2388: &&= \langle \vec{E}^{(-)}\left(\vec{R}_{1},t_{1}\right)\vec{E}^{(-)}
2389: \left(\vec{R}_{2},t_{2}\right)\vec{E}^{(+)}\left(\vec{R}_{2}, t_{2}\right)
2390: \vec{E}^{(+)}\left(\vec{R}_{1}, t_{1}\right)\rangle \ ,\label{t113}
2391: \end{eqnarray}
2392: is the two-time second-order correlation function of the EM field
2393: detected at a point $\vec{R}_{1}$ at time $t_{1}$ and at a point
2394: $\vec{R}_{2}$ at time $t_{2}$, and
2395: \begin{eqnarray}
2396: G^{(1)}\left(\vec{R}_{i},t_{i}\right) &=& \langle
2397: \vec{E}^{(-)}\left(\vec{R}_{i},t_{i}\right)
2398: \vec{E}^{(+)}\left(\vec{R}_{i},t_{i}\right)\rangle \ ,\label{t114}
2399: \end{eqnarray}
2400: is the first-order correlation function of the field (intensity) detected
2401: at a point $R_{i}$ at time $t_{i} (i=1,2)$.
2402:
2403: The correlation function $G^{(2)}(\vec{R}_{1},t_{1};
2404: \vec{R}_{2},t_{2})$ is proportional to a joint probability of
2405: finding one photon around the direction $\vec{R}_{1}$ at time $t_{1}$
2406: and another photon around the direction $\vec{R}_{2}$ at the moment of
2407: time $t_{2}$. For a coherent light, the probability of finding a
2408: photon around $\vec{R}_{1}$ at time $t_{1}$ is independent of the
2409: probability of finding another photon around $\vec{R}_{2}$ at time
2410: $t_{2}$, and then $G^{(2)}(\vec{R}_{1},t_{1};
2411: \vec{R}_{2},t_{2})$ simply factorizes into
2412: $G^{(1)}(\vec{R}_{1},t_{1})
2413: G^{(1)}(\vec{R}_{2},t_{2})$ giving
2414: $g^{(2)}(\vec{R}_{1},t_{1};\vec{R}_{2},t_{2})=1$. For a
2415: chaotic (thermal) field the second-order correlation function for
2416: $t_{1}=t_{2}$ is greater than for $t_{2}-t_{1}=\tau >0$ giving
2417: $g^{(2)}(\vec{R}_{1},t_{1};\vec{R}_{2},t_{1})
2418: >g^{(2)}(\vec{R}_{1},t_{1};\vec{R}_{2},t_{1}+\tau )$. This
2419: is a manifestation of the tendency of photons to be emitted in
2420: correlated pairs, and is called photon bunching. Photon antibunching,
2421: as the name implies, is the opposite of bunching, and describes a
2422: situation in which fewer photons appear close together than further
2423: apart. The condition for photon antibunching is
2424: $g^{(2)}(\vec{R}_{1},t_{1};\vec{R}_{2},t_{1})
2425: <g^{(2)}(\vec{R}_{1},t_{1};\vec{R}_{2},t_{1}+\tau )$ and
2426: implies that the probability of detecting two photons at the same
2427: time $t$ is smaller than the probability of detecting two photons at
2428: different times $t$ and $t+\tau$. Moreover, the fact that there is a
2429: small probability of detecting photon pairs with zero time separation
2430: indicated that the one-time correlation function
2431: $g^{(2)}(\vec{R}_{1},t;\vec{R}_{2},t)$ is smaller than
2432: one. This effect is called photon anticorrelation. The normalized
2433: one-time second-order correlation function carries also information
2434: about photon statistics, which is given by the Mandel's $Q$ parameter
2435: defined as~\cite{man82}
2436: \begin{eqnarray}
2437: Q = qT\left[g^{(2)}\left(\vec{R}_{1},t; \vec{R}_{2},t\right)
2438: -1\right] \ ,\label{t115}
2439: \end{eqnarray}
2440: where $q$ is the quantum efficiency of the detector and $T$ is the
2441: photon counting time.
2442:
2443: We can relate the field correlation functions~(\ref{t113})
2444: and~(\ref{t114}) to the correlation functions of the atomic operators,
2445: which will allow us to apply directly the master equation~(\ref{t42})
2446: to calculate photon antibunching in a collective atomic system.
2447: The relation between the positive frequency part of the electric field
2448: operator at a point $\vec{R}=R\bar{R}$, in the far-field zone,
2449: and the atomic dipole operators $S_{i}^{-}$, is given by the well-known
2450: expression~\cite{leh,ag74}
2451: \begin{eqnarray}
2452: \vec{E}^{(+)}\left(\vec{R},t\right) &=&
2453: \vec{E}^{(+)}_{0}\left(\vec{R},t\right) \nonumber \\
2454: &-& \sum_{i=1}^{2}\frac{\omega_{i}}{c^{2}}
2455: \frac{\bar{R}\times \left(\bar{R}\times
2456: \vec{\mu}_{i}\right)}{R}
2457: S_{i}^{-}\left(t -\frac{R}{c}\right) \exp
2458: \left(-ik\bar{R}\cdot \vec{r}_{i}\right) \ ,\label{t116}
2459: \end{eqnarray}
2460: where $\omega_{i}$ is the angular frequency of the $i$th atom located
2461: at a point $\vec{r}_{i}$, and $\vec{E}^{(+)}_{0}\left(\vec{R},t\right)$
2462: denotes the positive frequency part of the field in the absence of the
2463: atoms.
2464:
2465: If we assume that initially the field is in the vacuum state,
2466: then the free-field part $\vec{E}^{(+)}_{0}(\vec{R},t)$ does
2467: not contribute to the expectation values of the normally ordered
2468: operators. Hence, substituting Eq.~(\ref{t116}) into Eqs.~(\ref{t113})
2469: and~(\ref{t114}), we obtain
2470: \begin{eqnarray}
2471: G^{\left( 2\right)}\left( \vec{R},t;\vec{R},t+\tau \right)
2472: &=& u(\vec{R}_{1})u(\vec{R}_{2})
2473: \sum_{i,j,k,l=1}^{N}\left(\Gamma _{i}\Gamma_{j}\Gamma
2474: _{k}\Gamma_{l}\right)^{\frac{1}{2}} \nonumber \\
2475: &\times& \left\langle S^{+}_{i}\left( t\right) S^{+}_{k}\left(t+\tau \right)
2476: S^{-}_{l}\left( t+\tau\right) S^{-}_{j}\left(t\right) \right\rangle
2477: \nonumber \\
2478: &\times& \exp \left[ik\left(\bar{R}_{1}\cdot \vec{r}_{ij}+\bar{R}_{2}
2479: \cdot \vec{r}_{kl}\right)\right] \ ,\label{t117}\\
2480: G^{\left( 1\right) }\left(\vec{R},t\right)
2481: &=& u(\vec{R})
2482: \sum_{i,j=1}^{N}\left(\Gamma_{i}\Gamma_{j}\right)^{\frac{1}{2}}\left\langle
2483: S_{i}^{+}\left( t\right) S^{-}_{j}\left( t\right) \right\rangle
2484: \nonumber \\
2485: &\times& \exp \left(ik\bar{R}\cdot \vec{r}_{ij}\right)
2486: \ ,\label{t118}
2487: \end{eqnarray}
2488: where $\tau =t_{2}-t_{1}$, $\Gamma_{i}$ is the damping rate of the
2489: $i$th atom, and $u(\vec{R})$ is a constant given in
2490: Eq.~(\ref{t70}). The second-order correlation function~(\ref{t117})
2491: involves two-time atomic correlation function that can be calculated
2492: from the master equation~(\ref{t36}) or~(\ref{t42}) and applying the
2493: quantum regression theorem~\cite{lax}. From the quantum regression
2494: theorem, it is well known that for $\tau >0$ the two-time correlation
2495: function $\langle S^{+}_{i}\left( t\right) S^{+}_{k}\left(t+\tau
2496: \right)S^{-}_{l}\left( t+\tau\right) S^{-}_{j}\left(t\right)
2497: \rangle$ satisfies the same equation of motion as the one-time
2498: correlation function $\langle S_{k}^{+}\left( t\right)
2499: S^{-}_{l}\left( t\right) \rangle$.
2500:
2501: We shall first of all consider the simplest collective system for
2502: photon antibunching; two identical atoms in the Dicke model. Whilst
2503: this model is not well satisfied with the present sources of two-atom
2504: systems, it does enable analytic treatments that allow to understand
2505: the role of the collective damping in the generation of nonclassical
2506: light.
2507:
2508: For the two-atom Dicke model the master equation~(\ref{t42}) reduces
2509: to
2510: \begin{eqnarray}
2511: \frac{\partial \hat{\rho}}{\partial t} &=& \frac{1}{2}i\Omega
2512: \left[S^{+}+S^{-}, \hat{\rho}\right] -\frac{1}{2}\Gamma
2513: \left(S^{+}S^{-}\hat{\rho} +\hat{\rho}S^{+}S^{-}
2514: -2S^{-}\hat{\rho}S^{+}\right) \ ,\label{t119}
2515: \end{eqnarray}
2516: where $S^{\pm}=S^{\pm}_{1}+S^{\pm}_{2}$ and $S^{z}=S^{z}_{1}+S^{z}_{2}$
2517: are the collective atomic operators and $\Omega$ is the Rabi frequency
2518: of the driving field, which in the Dicke model is the same for both
2519: atoms. For simplicity, the laser frequency $\omega_{L}$ is taken to
2520: be exactly equal to the atomic resonant frequency $\omega_{0}$.
2521:
2522: The secular approximation technique has been suggested by
2523: Agarwal {\it et al.}~\cite{aga79} and Kilin~\cite{kil80},
2524: which greatly simplifies the master equation~(\ref{t119}). Hassan {\it
2525: et al.}~\cite{has82} and Cordes~\cite{cor82,cor87} have generalised the
2526: method to include non-zero detuning of the laser field and the
2527: quasistatic dipole-dipole potential. The technique is a modification of
2528: a collective dressed-atom approach developed by Freedhoff~\cite{fr79} and
2529: is valid if the Rabi frequency of the driving field
2530: is much greater than the damping rates of the atoms, $\Omega \gg
2531: \Gamma$. To implement the technique, we transform the collective
2532: operators into new (dressed) operators
2533: \begin{eqnarray}
2534: S^{\pm} &=& \pm \frac{1}{2}i\left(R^{+} +R^{-}\right) +R^{z} \
2535: ,\nonumber \\
2536: S^{z} &=& -\frac{1}{2}i\left(R^{+} -R^{-}\right) \ .\label{t120}
2537: \end{eqnarray}
2538: The operators $R$ are a rotation of the operators $S$. For a strong
2539: driving field, the operators $R^{\pm}$ vary rapidly with time,
2540: approximately as $\exp (\pm i\Omega t)$, while $R^{z}$ varies slowly
2541: in time. By expressing the operators $S^{\pm}$ and $S^{z}$ in terms of
2542: the operators $R^{\pm}$ and $R^{z}$, and substituting into the master
2543: equation~(\ref{t119}), we find that certain terms are slowly varying
2544: in time while others oscillate rapidly. The secular approximation then
2545: involves dropping the rapidly oscillating terms that results in an
2546: approximate master equation of the form
2547: \begin{eqnarray}
2548: \frac{\partial \hat{\rho}}{\partial t} = i\Omega
2549: \left[R^{z}, \hat{\rho}\right] &-& \frac{1}{2}\Gamma \left\{
2550: \left(R^{z}R^{z}\hat{\rho} +\hat{\rho}R^{z}R^{z}
2551: -2R^{z}\hat{\rho}R^{z}\right)\right. \nonumber \\
2552: &&\left. +\frac{1}{4}\left(R^{+}R^{-}\hat{\rho}
2553: +\hat{\rho}R^{+}R^{-} -2R^{-}\hat{\rho}R^{+}\right)\right.
2554: \nonumber \\
2555: &&\left. +\frac{1}{4}\left(R^{-}R^{+}\hat{\rho}
2556: +\hat{\rho}R^{-}R^{+} -2R^{+}\hat{\rho}R^{-}\right)\right\}
2557: \ .\label{t121}
2558: \end{eqnarray}
2559: The master equation~(\ref{t121}) enables to obtain equations of motion
2560: for the expectation value of an arbitrary combination of the
2561: transformed operators $R$. In particular, the master equation leads to
2562: simple equations of motion for the expectation values required to calculate
2563: the normalized second-order correlation function. The required equations of
2564: motion are given by
2565: \begin{eqnarray}
2566: \frac{d}{dt}\left\langle R^{z}\right\rangle &=& -\frac{1}{2}\Gamma
2567: \left\langle R^{z}\right\rangle \ ,\nonumber \\
2568: \frac{d}{dt}\langle R^{\pm}\rangle &=&
2569: -\left(\frac{3}{4}\Gamma \pm i\Omega \right)
2570: \langle R^{\pm}\rangle \ ,\nonumber \\
2571: \frac{d}{dt}\langle R^{+}R^{+}\rangle &=&
2572: -\left(\frac{5}{2}\Gamma +2i\Omega \right)
2573: \langle R^{+}R^{+}\rangle \ .\label{t122}
2574: \end{eqnarray}
2575: The solution of these decoupled differential equations is
2576: straightforward. Performing the integration and applying the quantum
2577: regression theorem~\cite{lax}, we obtain from
2578: Eqs.~(\ref{t122}) and (\ref{t112}) the following solution for the
2579: normalized second-order correlation function~\cite{ftk83,hds}
2580: \begin{eqnarray}
2581: g^{(2)}\left(\tau \right) &\equiv&
2582: \lim_{t\rightarrow \infty} g^{(2)}\left(\vec{R}_{1},t;\vec{R}_{2},
2583: t+\tau \right) = 1
2584: +\frac{1}{32}\exp \left(-\frac{3}{2}\Gamma \tau \right) \nonumber
2585: \\
2586: && +\frac{3}{32}\exp \left(-\frac{5}{2}\Gamma \tau \right)\cos
2587: \left(2\Omega \tau \right)
2588: -\frac{3}{8}\exp \left(-\frac{3}{4}\Gamma \tau \right) \cos
2589: \left(\Omega \tau \right) \ .\label{t123}
2590: \end{eqnarray}
2591: The correlation function $g^{(2)}\left(\tau \right)$ is shown in
2592: Fig.~\ref{ftfig8}
2593: as a function of $\tau$ for different $\Omega$. For $\tau =0$, the
2594: correlation function $g^{(2)}\left(0 \right)=0.75$, showing the photon
2595: anticorrelation in the emitted fluorescence field. As $\tau$
2596: increases, the correlation function increases $(g^{(2)}\left(\tau
2597: \right) > g^{(2)}\left(0 \right))$, which reflects photon
2598: antibunching in the emitted field. However, the photon anticorrelation
2599: in the two-atom fluorescence field is reduced compared to that for a
2600: single atom, for which $g^{(2)}\left(0 \right)=0$. This result
2601: indicates that the collective damping reduces the photon
2602: anticorrelations in the emitted fluorescence field.
2603: %
2604: \begin{figure}[t]
2605: \begin{center}
2606: \includegraphics[width=10cm]{ftfig8.eps}
2607: \end{center}
2608: \caption{The normalised second-order correlation function
2609: $g^{(2)}\left(\tau \right)$ as a function of $\tau$ and different
2610: $\Omega$; $\Omega =2.5\Gamma$ (solid line), $\Omega =10\Gamma$ (dashed
2611: line).}
2612: \label{ftfig8}
2613: \end{figure}
2614:
2615:
2616: As we have mentioned above, in the Dicke model the dipole-dipole
2617: interaction between the atoms is
2618: ignored. This approximation has no justification, since for small
2619: interatomic separations the dipole-dipole parameter $\Omega_{12}$,
2620: which varies as $(k_{0}r_{12})^{-3}$, is very large and goes to
2621: infinity as $r_{12}$ goes to
2622: zero (see Fig.~\ref{ftfig1}). Moreover, the Dicke model does not
2623: correspond to the experimentally realistic systems in which atoms are
2624: separated by distances comparable to the resonant wavelength.
2625: Ficek {\it et al.}~\cite{ftk83} and Lawande {\it et al.}~\cite{law85}
2626: have shown that the dipole-dipole
2627: interaction does not considerably affect the anticorrelation effect
2628: predicted in the Dicke model. Richter~\cite{rich1} has shown that the
2629: value $g^{(2)}\left(0 \right)=0.75$ can in fact be reduced such that
2630: even the complete photon anticorrelation $g^{(2)}\left(0 \right)=0$ can
2631: be obtained, if the dipole-dipole is included and the laser frequency
2632: is detuned from the atomic transition frequency. To show this, we
2633: calculate the normalized second-order correlation function~(\ref{t112})
2634: for the steady-state fluorescence field from two identical atoms $(N=2)$,
2635: and $\tau =0$. In this case, the correlation function~(\ref{t112}) with
2636: Eqs.~(\ref{t117}) and (\ref{t118}) can be written as
2637: \begin{eqnarray}
2638: g^{(2)}\left(0 \right) =\frac{2U
2639: \left\{1+\cos\left[k\vec{r}_{12}\cdot
2640: \left(\bar{R}_{1}-\bar{R}_{2}\right)\right]\right\}}{\left[
2641: 1+W\cos\left(k\vec{r}_{12}\cdot
2642: \bar{R}_{1}\right)\right] \left[
2643: 1+W\cos\left(k\vec{r}_{12}\cdot
2644: \bar{R}_{2}\right)\right]} \ ,\label{t124}
2645: \end{eqnarray}
2646: where $U$ and $W$ are the steady-state atomic correlation functions
2647: \begin{eqnarray}
2648: U = \frac{\langle
2649: S^{+}_{1}S^{+}_{2}S^{-}_{1}S^{-}_{2}\rangle}{\langle
2650: S^{+}_{1}S^{-}_{1} +S^{+}_{2}S^{-}_{2}\rangle^{2}} \ ,\qquad
2651: W = \frac{\langle
2652: S^{+}_{1}S^{-}_{2} +S^{+}_{2}S^{-}_{1}\rangle}{\langle
2653: S^{+}_{1}S^{-}_{1} +S^{+}_{2}S^{-}_{2}\rangle} \ .\label{t125}
2654: \end{eqnarray}
2655: The steady-state correlation functions are easily obtained from the
2656: master equation~(\ref{t42}). We can simplify the solutions assuming
2657: that the atoms are in equivalent positions in the driving field,
2658: which can be achieved by propagating the laser field in the direction
2659: perpendicular to the
2660: interatomic axis. In this case we get analytical solutions, otherwise
2661: for $\vec{k}_{L}\cdot \vec{r}_{12}\neq 0$ numerical methods are
2662: more appropriate~\cite{fs,rfd,rich2}. With $\vec{k}_{L}\cdot
2663: \vec{r}_{12}=0$ the master equation~(\ref{t42}) leads to a closed set
2664: of nine equations of motion for the atomic correlation functions.
2665: This set of equations can be solved exactly in the
2666: steady-state~\cite{ftk84}, and the solutions for $U$ and $W$ are
2667: \begin{eqnarray}
2668: U &=& \frac{\Omega^{4}
2669: +\left(\Gamma^{2}+4\Delta_{L}^{2}\right)\Omega^{2}
2670: +\left(\Gamma^{2}+4\Delta_{L}^{2}\right)\left[ \frac{1}{4}
2671: \left(\Gamma +\Gamma_{12}\right)^{2}
2672: +\left(\Delta_{L}-\Omega_{12}\right)^{2}\right]}
2673: {\left(\Gamma^{2} +4\Delta_{L}^{2} +2\Omega^{2}\right)^{2}}
2674: \ ,\nonumber \\
2675: W &=& \frac{\left(\Gamma^{2}+4\Delta_{L}^{2}\right)}
2676: {\left(\Gamma^{2}+4\Delta_{L}^{2}+2\Omega^{2}\right)} \
2677: .\label{t126}
2678: \end{eqnarray}
2679: One can see from Eqs.~(\ref{t124}) and (\ref{t126}) that there are two
2680: different processes which can lead to the total anticorrelation,
2681: $g^{(2)}\left(0 \right)=0$. The
2682: first one involves an observation of the fluorescence field with two
2683: detectors located at different points. If the correlation function is
2684: measured using two detectors, $\bar{R}_{1}\neq \bar{R}_{2}$, and then
2685: we obtain $g^{(2)}\left(0 \right)=0$ whenever the positions of the
2686: detectors are such that
2687: \begin{eqnarray}
2688: \left\{1+\cos\left[k\vec{r}_{12}\cdot
2689: \left(\bar{R}_{1}-\bar{R}_{2}\right)\right]\right\} =0 \
2690: ,\label{t127}
2691: \end{eqnarray}
2692: which happens when
2693: \begin{eqnarray}
2694: k\vec{r}_{12}\cdot \left(\bar{R}_{1}-\bar{R}_{2}\right) =
2695: \left(2n+1\right)\pi \ ,\quad n=0,\pm 1,\pm 2 \ldots \label{t128}
2696: \end{eqnarray}
2697: In other words, two photons can never be simultaneously detected at two
2698: points separated by an odd number of $\lambda/2r_{12}$, despite the fact
2699: that one photon can be detected anywhere. This complete anticorrelation
2700: effect is due to spatial interference between different photons and
2701: reflects the fact that one photon must have come from one source and
2702: one from the other, but we cannot tell which came from which.
2703:
2704: It should be emphasised that this effect is independent of the interatomic
2705: interactions and the Rabi frequency of the
2706: driving field. The vanishing of $g^{(2)}\left(0 \right)$ for two photons at
2707: widely separated points $\vec{R}_{1}$ and $\vec{R}_{2}$ is an example of
2708: quantum-mechanical nonlocality, that the outcome of a detection measurement
2709: at $\vec{R}_{1}$ appears to be influenced by where we have chosen to locate
2710: the $\vec{R}_{2}$ detector. At certain positions $\vec{R}_{2}$ we can never
2711: detect a photon at $\vec{R}_{1}$ when there is a photon detected at
2712: $\vec{R}_{2}$, whereas at other position $\vec{R}_{2}$ it is possible.
2713: The photon correlation argument shows clearly that quantum theory does not
2714: in general describe an objective physical reality independent of
2715: observation.
2716:
2717: The second process involves the shift of the collective atomic states
2718: due to the dipole-dipole interaction that can lead
2719: to $g^{(2)}\left(0 \right)=0$ even if the correlation function is
2720: measured with a single detector $(\bar{R}_{1}=\bar{R}_{2})$ or two
2721: detectors in configurations different from that given by
2722: Eq.~(\ref{t128}). For a weak driving field $(\Omega \ll \Gamma)$ and
2723: large detunings such that $\Delta_{L} =\Omega_{12}\gg \Gamma$, the
2724: correlation function~(\ref{t112}) with $\bar{R}_{1}=\bar{R}_{2}$
2725: simplifies to
2726: \begin{eqnarray}
2727: g^{(2)}\left(0 \right) \approx \frac{\left(\Gamma
2728: +\Gamma_{12}\right)^{2}}{4\Delta_{L}^{2}} \ .\label{t129}
2729: \end{eqnarray}
2730: %
2731: \begin{figure}[t]
2732: \begin{center}
2733: \includegraphics[width=10cm]{ftfig9.eps}
2734: \end{center}
2735: \caption{The normalised second-order correlation function
2736: $g^{(2)}\left(0 \right)$ as a function of $\Delta_{L}$ for
2737: $\bar{R}_{1}=\bar{R}_{2}=\bar{R}, \vec{r}_{12}\perp \bar{R},
2738: \bar{\mu}\perp \bar{r}_{12}$,
2739: $\Omega =0.5\Gamma$ and different $r_{12}$; $r_{12} =10\lambda$
2740: (solid line), $r_{12} =0.15\lambda$ (dashed line), $r_{12}=0.08\lambda$
2741: (dashed-dotted line).}
2742: \label{ftfig9}
2743: \end{figure}
2744: %
2745: Thus, a pronounced photon anticorrelation, $g^{(2)}\left(0
2746: \right)\approx 0$, can be obtained for large detunings such that
2747: $\Delta_{L} =\Omega_{12}$, i.e., when the dipole-dipole interaction
2748: shift of the collective states and the
2749: detuning cancel out mutually. The correlation function
2750: $g^{(2)}\left(0 \right)$ of the steady-state fluorescence field is
2751: illustrated graphically in Fig.~\ref{ftfig9} as a function of $\Delta_{L}$
2752: for the single detector configuration with $\bar{R}_{1}=\bar{R}_{2}=\bar{R}$,
2753: and different $r_{12}$. The graphs show that $g^{(2)}\left(0 \right)$ strongly
2754: depends on $\Delta_{L}$, and the total photon anticorrelation can be
2755: obtained for $\Delta_{L} =\Omega_{12}$. Referring to Fig.~\ref{ftfig2},
2756: the condition $\Delta_{L} =\Omega_{12}$ corresponds to the laser frequency
2757: tuned to the resonance with the $\ket g \rightarrow \ket s$
2758: transition. Since the other levels are far from the resonance, the
2759: two-atom system behaves like a single two-level system with the
2760: ground state $\ket g$ and the excited state $\ket s$.
2761:
2762:
2763: \subsection{Squeezing}\label{ftsec52}
2764:
2765:
2766: To understand squeezed light, recall that the electric
2767: field amplitude $\vec{E}\left( \vec{r}\right) $ may
2768: be expressed by positive- and negative-frequency parts
2769: \begin{equation}
2770: \vec{E}\left( \vec{r}\right) =\vec{E}^{\left( +\right) }\left( \vec{r}
2771: \right) +\vec{E}^{\left( -\right) }\left( \vec{r}\right) \ ,
2772: \label{t130}
2773: \end{equation}
2774: where
2775: \begin{equation}
2776: \vec{E}^{\left( +\right) }\left( \vec{r}\right) =\left( \vec{E}^{\left(
2777: -\right) }\left( \vec{r}\right) \right) ^{\dagger }=-i\sum_{\vec{k}s}%
2778: \left( \hbar \omega _{k}/2\varepsilon_{o}
2779: V\right) ^{1/2}\bar{e}_{\vec{k}s}\hat{a}_{\vec{k}s}e^{i\vec{k} \cdot
2780: \vec{r}} \ ,\label{t131}
2781: \end{equation}
2782: and $\omega_{k}=c\left| \vec{k}\right| $ is the angular
2783: frequency of the mode $\vec{k}$.
2784:
2785: We introduce two Hermitian combinations (quadrature components) of the field
2786: components that are $\pi /2$ out of phase as
2787: \begin{eqnarray}
2788: \vec{E}_{\theta } &=&\vec{E}^{\left( +\right) }\left(
2789: \vec{R}\right) e^{i\theta }+\vec{E}^{\left( -\right) }\left( \vec{R}\right)
2790: e^{-i\theta } \ , \nonumber \\
2791: \vec{E}_{\theta -\pi /2} &=&-i\left( \vec{E}^{\left(
2792: +\right) }\left( \vec{R}\right) e^{i\theta }-\vec{E}^{\left( -\right)
2793: }\left( \vec{R}\right) e^{-i\theta }\right) \ , \label{t132}
2794: \end{eqnarray}
2795: where
2796: \begin{equation}
2797: \theta =\omega t-\vec{k}\cdot \vec{R} \ , \label{t133}
2798: \end{equation}
2799: and $\omega $ is the angular frequency of the quadrature components.
2800:
2801: The quadrature components do not commute, satisfying the commutation
2802: relation
2803: \begin{equation}
2804: \left[ \vec{E}_{\theta } ,\vec{E}_{\theta -\pi/2} \right] =2iC \
2805: ,\label{t134}
2806: \end{equation}
2807: where $C$ is a positive number
2808: \begin{equation}
2809: C=\sum_{\vec{k}s}\left| \hbar \omega _{k}/2\varepsilon_{o}
2810: V \right| \ .\label{t135}
2811: \end{equation}
2812: Hence the two quadrature components cannot be simultaneously precisely
2813: measured, and from the Heisenberg uncertainty principle, we find that the
2814: variances $\left\langle \Delta \vec{E}_{\theta }^{2} \right\rangle$ and
2815: $\left\langle \Delta \vec{E}_{\theta -\pi /2}^{2}
2816: \right\rangle $ satisfy the inequality
2817: \begin{equation}
2818: \left\langle \Delta \vec{E}_{\theta }^{2} \right\rangle
2819: \left\langle \Delta \vec{E}_{\theta -\pi /2}^{2}
2820: \right\rangle \geq C^{2} \ , \label{t136}
2821: \end{equation}
2822: where the equality holds for a minimum uncertainty state of the field.
2823:
2824: The variances $\left\langle \Delta \vec{E}_{\theta }^{2}\right\rangle $
2825: and $\left\langle \Delta \vec{E}_{\theta -\pi/2}^{2} \right\rangle $ depend on
2826: the state of the field and can be larger or smaller than $C$. A
2827: chaotic state of
2828: the field leads to the variances in both components larger than $C$:
2829: \begin{eqnarray}
2830: \left\langle \Delta \vec{E}_{\theta }^{2}\right\rangle \geq C
2831: \quad &\mbox{and}& \quad \left\langle \Delta \vec{E}_{\theta -\pi /2}^{2}
2832: \right\rangle \geq C \ .\label{t137}
2833: \end{eqnarray}
2834: If the field is in a coherent or vacuum state
2835: \begin{equation}
2836: \left\langle \Delta \vec{E}_{\theta }^{2}\right\rangle
2837: = \left\langle \Delta \vec{E}_{\theta -\pi /2}^{2}\right\rangle =C
2838: \ ,\label{t138}
2839: \end{equation}
2840: which is an example of a minimum uncertainty state.
2841:
2842: A squeezed state of the field is defined to be one in which the
2843: variance in one of the two quadrature components is less than that
2844: for the vacuum field
2845: \begin{eqnarray}
2846: \left\langle \Delta \vec{E}_{\theta }^{2}\right\rangle
2847: <C \quad &\mbox{or}& \quad \left\langle \Delta \vec{E}_{\theta -\pi
2848: /2}^{2}\right\rangle < C \ . \label{t139}
2849: \end{eqnarray}
2850: The variances can be expressed as
2851: \begin{eqnarray}
2852: \left\langle \Delta \vec{E}_{\theta }^{2} \right\rangle
2853: &=& C + \left\langle :\Delta \vec{E}_{\theta}^{2} :\right\rangle \
2854: ,\nonumber \\
2855: \left\langle \Delta \vec{E}_{\theta -\pi /2}^{2}
2856: \right\rangle &=&C+\left\langle :\Delta \vec{E}_{\theta
2857: -\pi /2}^{2}:\right\rangle \ ,\label{t140}
2858: \end{eqnarray}
2859: where the colon stands for normal ordering of the operators.
2860:
2861: As the squeezed state has been defined by the requirement that either $%
2862: \left\langle \Delta \vec{E}_{\theta }^{2}\right\rangle
2863: $ or $\left\langle \Delta \vec{E}_{\theta -\pi /2}^{2}
2864: \right\rangle $ be below the vacuum level $C$, it follows immediately from
2865: Eq.~(\ref{t140}) that either
2866: \begin{eqnarray}
2867: \left\langle :\Delta \vec{E}_{\theta }^{2}
2868: :\right\rangle < 0 \quad &\mbox{or}& \quad \left\langle :\Delta
2869: \vec{E}_{\theta -\pi/2}^{2} :\right\rangle < 0 \label{t141}
2870: \end{eqnarray}
2871: for the field in a squeezed state.
2872:
2873: We now determine the relation between variances in the field and the
2874: atomic dipole operators. Using Eq.~(\ref{t116}), which relates the field
2875: operators to the atomic dipole operators, we obtain
2876: \begin{eqnarray}
2877: \left\langle :\Delta \vec{E}_{\alpha }^{2}:\right\rangle =
2878: u\left(\vec{R}\right)\left[\langle \left(\Delta
2879: S_{\alpha}\right)^{2}\rangle +\frac{1}{2}\langle
2880: S_{3}\rangle\right] \ ,\label{t142}
2881: \end{eqnarray}
2882: where $\alpha =\theta ,\theta -\pi/2$, $S_{\alpha}$ and $S_{3}$ are
2883: real (phase) operators defined as
2884: \begin{eqnarray}
2885: S_{\theta} &=& \frac{1}{2}\left(S_{\theta}^{+}
2886: +S_{\theta}^{-}\right) \ ,\qquad
2887: S_{\theta -\pi/2} = \frac{1}{2i}\left(S_{\theta}^{+}
2888: -S_{\theta}^{-}\right) \ ,\label{t143}
2889: \end{eqnarray}
2890: and
2891: \begin{eqnarray}
2892: S_{3} = \frac{1}{2}\left[S_{\theta}^{+},
2893: S_{\theta}^{-}\right] \ ,\label{t144}
2894: \end{eqnarray}
2895: with
2896: \begin{eqnarray}
2897: S_{\theta}^{\pm} =\sum_{i=1}^{N}S_{i}^{\pm}\exp \left[\pm
2898: i\left(k\bar{R}\cdot \vec{r}_{i} -\theta\right) \right] \
2899: .\label{t145}
2900: \end{eqnarray}
2901:
2902: We first consider quantum fluctuations in the fluorescence field emitted
2903: by two identical atoms in the Dicke model. To simplify the
2904: calculations we will treat only the case of zero detuning, $\Delta_{L}=0$.
2905: Assuming that initially $(t=0)$ the atoms were in their ground states, we
2906: find from Eqs.~(\ref{t122}) and (\ref{t142}) the following
2907: expressions for the time-dependent variances~\cite{ftk87s}
2908: \begin{eqnarray}
2909: F_{\theta =0}\left( t\right) &\equiv&
2910: \left\langle :\Delta \vec{E}_{\theta =0 }^{2}:\right\rangle /\left(
2911: 2u(\vec{R})\right) = \frac{1}{3} -\frac{1}{8}\exp
2912: \left(-\frac{5}{2}\Gamma t\right)\cos\left(2\Omega t\right)
2913: \nonumber \\
2914: &+& \frac{1}{24}\exp \left(-\frac{3}{2}\Gamma t\right)
2915: -\frac{1}{2}\exp\left(-\frac{3}{2}\Gamma t\right)
2916: \sin^{2}\left(\Omega t\right) \nonumber \\
2917: &-& \frac{1}{4}\exp\left(-\frac{3}{4}\Gamma
2918: t\right)\cos\left(\Omega t\right) \ ,\label{t146}
2919: \end{eqnarray}
2920: and
2921: \begin{eqnarray}
2922: F_{\theta =\pi/2}\left( t\right) &\equiv&
2923: \left\langle :\Delta \vec{E}_{\theta =\pi/2 }^{2}:\right\rangle
2924: /\left(2u(\vec{R})\right) = \frac{1}{3}
2925: -\frac{1}{12}\exp \left(-\frac{3}{2}\Gamma t\right) \nonumber \\
2926: &-& \frac{1}{4}\exp\left(-\frac{3}{4}\Gamma
2927: t\right)\cos\left(\Omega t\right) \ .\label{t147}
2928: \end{eqnarray}
2929: In writing Eqs.~(\ref{t146}) and (\ref{t147}), we have assumed that the
2930: angular frequency of the quadrature components is equal to the laser
2931: frequency, $\omega =\omega_{L}$, and we have normalised the variances
2932: such that $F(t)$ determines fluctuations per atom.
2933: It is easily to show that the variance $F_{\theta =\pi/2}\left(
2934: t\right)$ is positive for all times $t$, and squeezing
2935: $(F_{\theta}<0)$ can be observed in the variance $F_{\theta =0}\left(
2936: t\right)$. The time dependence of the variance $F_{\theta =0}\left(
2937: t\right)$ is shown in Fig.~\ref{ftfig10} for two different values of
2938: the Rabi frequency. It is seen that squeezing appears in the transient
2939: regime of resonance fluorescence and its maximum value (minimum of $F$)
2940: moves towards shorter times as $\Omega$ increases. The optimum squeezing
2941: reaches a value of $-1/16$ at a very short time. This value is equal
2942: to the maximum possible squeezing in a single two-level
2943: atom~\cite{wz81,ftk84a,bk88,dfk94}. Thus,
2944: the collective damping does not affect squeezing in the two-atom
2945: Dicke model. This is in contrast to the photon anticorrelation
2946: effect which is greatly reduced by the collective damping.
2947: %
2948: \begin{figure}[t]
2949: \begin{center}
2950: \includegraphics[width=10cm]{ftfig10.eps}
2951: \end{center}
2952: \caption{The variance $F_{\theta =0}\left( t\right)$ as a function
2953: of time for different $\Omega$; $\Omega =100\Gamma$
2954: (solid line), $\Omega = 200\Gamma$ (dashed line).}
2955: \label{ftfig10}
2956: \end{figure}
2957:
2958: Figure~\ref{ftfig10} shows that in the two-atom Dicke model
2959: there is no squeezing in the steady-state resonance fluorescence when
2960: the atoms are excited by a strong laser field. Ficek {\it et
2961: al.}~\cite{ftk84} and Richter~\cite{rich3} have shown that similarly
2962: as in the case of photon anticorrelations, a large squeezing can
2963: be obtained in the steady-state resonance fluorescence from a
2964: strongly driven two-atom system, if the dipole-dipole interaction is included
2965: and the laser frequency is detuned from the atomic transition
2966: frequency. This is shown in Fig.~\ref{ftfig11}, where we plot $F_{\theta =0}$,
2967: calculated from Eq.~(\ref{t142}) and the master equation~(\ref{t42}),
2968: for the steady-state resonance fluorescence from two identical atoms,
2969: with $\vec{r}_{12}\perp \bar{R}$, $\Omega =0.5\Gamma$,
2970: $\vec{k}_{L}\cdot \vec{r}_{12}= 0$ and different~$r_{12}$. It is
2971: evident from Fig.~\ref{ftfig11} that a large squeezing can be obtained
2972: for a finite $\Delta_{L}$ and its maximum shifts towards larger
2973: $\Delta_{L}$ as the interatomic separation decreases. Similar as in
2974: the case of photon anticorrelations, the maximum squeezing appears
2975: at $\Delta_{L}=\Omega_{12}$, and can again be attributed to the
2976: shift of the collective energy states due to the dipole-dipole
2977: interaction.
2978:
2979: The variance $F_{\theta =0}$, shown in Fig.~\ref{ftfig11}, exhibits not
2980: only the large squeezing at finite detuning $\Delta_{L}$, but also a small
2981: squeezing near $\Delta_{L}=0$. In contrast to the
2982: squeezing at finite $\Delta_{L}$, which has a clear physical
2983: interpretation, the source of squeezing at $\Delta_{L}$ is not easy
2984: to understand. To find the source of squeezing at $\Delta_{L}=0$,
2985: we simplify the calculations assuming that the angular frequency of the
2986: quadrature components $\omega =\omega_{L}$ and the fluorescence field
2987: is observed in the direction perpendicular to the interatomic axis,
2988: $\bar{R}\perp \vec{r}_{12}$. In this case, the variance
2989: $\left\langle :\Delta \vec{E}_{\alpha }^{2}:\right\rangle$, written
2990: in terms of the density matrix elements of the collective system, is
2991: given by~\cite{ft94}
2992: \begin{eqnarray}
2993: F_{\alpha} \equiv
2994: \left\langle :\Delta \vec{E}_{\alpha}^{2}:\right\rangle /\left(
2995: 2u(\vec{R})\right) &=& \frac{1}{4}\left\{2\rho_{ee} +2\rho_{ss}
2996: +\rho_{eg}e^{2i\alpha}
2997: +\rho_{ge}e^{-2i\alpha}\right. \nonumber \\
2998: &-&\left. \left[\left(\rho_{es}+\rho_{sg}\right)e^{i\alpha}
2999: +\left(\rho_{se}+\rho_{gs}\right)e^{-i\alpha}\right]^{2}\right\}
3000: \ .\label{t148}
3001: \end{eqnarray}
3002: This equation shows that the variance depends on phase $\alpha$
3003: not only through the one-photon coherences $\rho_{es}$ and
3004: $\rho_{sg}$, but also through the two-photon coherences $\rho_{eg}$
3005: and $\rho_{ge}$. This dependence suggests that there are two
3006: different processes that can lead to squeezing in the two-atom
3007: system. The one-photon coherences cause squeezing near one-photon
3008: resonances $\ket e \rightarrow \ket s$ and $\ket s \rightarrow \ket
3009: g$, whereas the two-photon coherences cause squeezing near the
3010: two-photon resonance $\ket g \rightarrow \ket e$.
3011: %
3012: \begin{figure}[t]
3013: \begin{center}
3014: \includegraphics[width=10cm]{ftfig11.eps}
3015: \end{center}
3016: \caption{The steady-state variance $F_{\theta =0}$ as a function
3017: of $\Delta_{L}$ for
3018: $\vec{r}_{12}\perp \bar{R}$, $\bar{\mu}\perp \bar{r}_{12}$,
3019: $\Omega =0.5\Gamma$,
3020: $\vec{k}_{L}\cdot \vec{r}_{12}= 0$ and different $r_{12}$;
3021: $r_{12} =10\lambda$ (solid line), $r_{12} =0.15\lambda$ (dashed line),
3022: $r_{12}=0.08\lambda$ (dashed-dotted line).}
3023: \label{ftfig11}
3024: \end{figure}
3025:
3026: To show this, we calculate the steady-state populations and coherences
3027: from the master equation~(\ref{t42}). We use the set of the collective
3028: states~(\ref{t63}) as an appropriate representation for the density
3029: operator
3030: \begin{equation}
3031: \hat{\rho} = \sum_{ij} \rho_{ij}\ket i\bra j ,\quad i,j=g,s,a,e
3032: \ ,\label{t149}
3033: \end{equation}
3034: where $\rho_{ij}$ are the density matrix elements in the basis of the
3035: collective states.
3036:
3037: After transforming to the collective state basis, the master
3038: equation~(\ref{t42}) leads to a closed system of fifteen equations
3039: of motion for the density matrix elements~\cite{ftk83}. However, for
3040: a specifically chosen geometry for the driving field, namely that the field
3041: is propagated perpendicularly to the atomic axis
3042: $(\vec{k}_{L}\cdot \vec{r}_{12}=0)$, the system of equations decouples into
3043: nine equations for symmetric and six equations for antisymmetric combinations
3044: of the density matrix elements~\cite{ftk83,hsf82,fs,rfd,rich1,rich2}. In this
3045: case, we can solve the system analytically, and find that the steady-state
3046: values of the populations and coherences are~\cite{ftk83,rich1}
3047: \begin{eqnarray}
3048: \rho_{ee} &=& \rho_{aa} =\frac{\tilde{\Omega}^{4}}{Z} \ ,\qquad
3049: \rho_{ss} = \frac{\tilde{\Omega}^{2}\left(\Gamma^{2}
3050: +4\Delta^{2}_{L}\right) +\tilde{\Omega}^{4}}{Z} \ ,\nonumber \\
3051: \rho_{es} &=& i\tilde{\Omega}^{3}\left(\Gamma+2i\Delta_{L}\right)/Z \
3052: ,\nonumber \\
3053: \rho_{sg} &=& -i\tilde{\Omega}\left\{\Gamma \tilde{\Omega}
3054: \left(\tilde{\Omega}
3055: +2i\Delta_{L}\right)\right. \nonumber \\
3056: &&\left. +\left(\Gamma^{2}+4\Delta_{L}^{2}\right)
3057: \left[\frac{1}{2}\left(\Gamma +\Gamma_{12}\right)
3058: +i\left(\Delta_{L}-\Omega_{12}\right)\right]\right\}/Z \
3059: ,\nonumber \\
3060: \rho_{eg} &=& \tilde{\Omega}^{2}\left(\Gamma +2i\Delta_{L}\right)
3061: \left[\frac{1}{2}\left(\Gamma +\Gamma_{12}\right)
3062: +i\left(\Delta_{L}-\Omega_{12}\right)\right]/Z \ ,\label{t150}
3063: \end{eqnarray}
3064: where
3065: \begin{equation}
3066: Z= 4\tilde{\Omega}^{4}+\left(\Gamma^{2}+4\Delta^{2}_{L}\right)
3067: \left\{2\tilde{\Omega}^{2}
3068: +\left[\frac{1}{4}\left(\Gamma +\Gamma_{12}\right)^{2}
3069: +\left(\Delta_{L} -\Omega_{12}\right)^{2}\right]\right\} \ ,\label{t151}
3070: \end{equation}
3071: and $\tilde{\Omega}=\Omega/\sqrt{2}$.
3072:
3073: Near the one-photon resonance $\ket s \rightarrow \ket g$ the
3074: detuning $\Delta_{L}=\Omega_{12}$, and assuming that $\Omega_{12}\gg
3075: \Omega, \Gamma$, the coherences reduce to
3076: \begin{eqnarray}
3077: \rho_{es} &=& \frac{\tilde{\Omega}}{8\Omega_{12}} \ ,\qquad
3078: \rho_{sg} = \frac{-i\left(\Gamma
3079: +\Gamma_{12}\right)}{4\tilde{\Omega}} \ ,\qquad
3080: \rho_{eg} = \frac{i\left(\Gamma +\Gamma_{12}\right)}
3081: {8\Omega_{12}} \ .\label{t152}
3082: \end{eqnarray}
3083: It is clear from Eq.~(\ref{t152}) that near the $\ket g \rightarrow
3084: \ket s$ resonance the coherence $\rho_{sg}$ is large, whereas the
3085: two-photon coherence is of order of $\Omega^{-1}_{12}$ and thus is
3086: negligible for large $\Omega_{12}$.
3087:
3088: Near two-photon resonance, $\Delta_{L}\approx 0$, and it follows
3089: from Eq.~(\ref{t150}) that in the limit of $\Omega_{12}\gg
3090: \Omega, \Gamma$ the coherences reduce to
3091: \begin{eqnarray}
3092: \rho_{es} = \frac{i\tilde{\Omega}^{3}}{\Gamma \Omega_{12}^{2}}
3093: \ ,\qquad
3094: \rho_{sg} = -\frac{\tilde{\Omega}}{\Omega_{12}} \ ,\qquad
3095: \rho_{eg} = -\frac{i\tilde{\Omega}^{2}}{\Gamma \Omega_{12}}
3096: \ .\label{t153}
3097: \end{eqnarray}
3098: In this regime, the coherences $\rho_{sg}$ and $\rho_{eg}$ are of order
3099: of magnitude $\Omega^{-1}_{12}$, but $\rho_{eg}$ dominates over the
3100: one-photon coherence $\rho_{sg}$ when the driving field is
3101: strong~\cite{va92}.
3102: %
3103: \begin{figure}[t]
3104: \begin{center}
3105: \includegraphics[width=10cm]{ftfig12.eps}
3106: \end{center}
3107: \caption{The steady-state variance $F_{\alpha}$ as a
3108: function of $\Delta_{L}$ for $r_{12}=0.05\lambda$, $\Omega
3109: =3\Gamma$, $\bar{R}\perp \vec{r}_{12}$,
3110: $\bar{\mu}\perp \bar{r}_{12}$ and different phases $\alpha$:
3111: $\alpha =\pi/2$ (dashed line), $\alpha =3\pi/4$ (solid line).}
3112: \label{ftfig12}
3113: \end{figure}
3114:
3115: The steady-state variance $F_{\alpha}$, calculated from Eqs.~(\ref{t148})
3116: and~(\ref{t150}), is plotted in Fig.~\ref{ftfig12} as a function of
3117: $\Delta_{L}$ for $r_{12}=0.05\lambda$, $\Omega =3\Gamma$,
3118: $\bar{R}\perp \vec{r}_{12}$ and different phases $\alpha$.
3119: The variance shows a strong dependence on~$\alpha$ near the one- and
3120: two-photon resonances. Moreover, a large squeezing is found at these
3121: resonances. It is also seen that near the two-photon resonance a
3122: change by $\pi/4$ of the phase $\alpha$ changes a dispersion-like
3123: structure of $F_{\alpha}$ into an absorption-like type. According to
3124: Eqs.~(\ref{t148}) and (\ref{t150}), the variance $F_{\alpha}$ for
3125: $\Omega_{12}\gg \Omega \gg \Gamma$, can be written as
3126: \begin{eqnarray}
3127: F_{\alpha} = \frac{\Omega^{2}}{\Omega_{12}}\left[
3128: \frac{\Delta_{L}}{\left(\Gamma^{2} +4\Delta_{L}^{2}\right)} \cos
3129: 2\alpha + \frac{\Gamma}{\left(\Gamma^{2} +4\Delta_{L}^{2}\right)}
3130: \sin 2\alpha \right] \ ,\label{t154}
3131: \end{eqnarray}
3132: where we retained only those terms which contribute near the
3133: two-photon resonance. Equation~(\ref{t154}) predicts a dispersion-like
3134: structure for $\alpha =0$ or $\pi/2$, and an absorption-like
3135: structure for $\alpha =\pi/4$. Moreover, we see that the presence of
3136: the dipole-dipole interaction is essential to obtain squeezing near
3137: the two-photon resonance. The emergence of an additional dipole-dipole
3138: interaction induced squeezing is a clear indication of a totally
3139: different process which can appear in a two-atom system. The
3140: dipole-dipole interaction shifts the collective states that induces
3141: two-photon transitions responsible for the origin of the two-photon
3142: coherence.
3143:
3144:
3145: \section{Quantum interference of optical fields}\label{ftsec6}
3146:
3147:
3148: In the classical theory of optical interference the EM field is
3149: represented by complex vectorial amplitudes $\vec{E}(\vec{R},t)$
3150: and $\vec{E}^{\ast}(\vec{R},t)$, and the first- and second-order
3151: correlation functions are defined in a similar way as the correlation
3152: functions~(\ref{t113}) and~(\ref{t114}) of the field operators
3153: $\vec{E}^{(+)}(\vec{R},t)$ and $\vec{E}^{(-)}(\vec{R},t)$.
3154: This could suggest that the only difference between the classical
3155: and quantum correlation functions is the classical amplitudes
3156: $\vec{E}^{\ast}(\vec{R},t)$ and $\vec{E}(\vec{R},t)$ are
3157: replaced by the field operators $\vec{E}^{(-)}(\vec{R},t)$ and
3158: $\vec{E}^{(+)}(\vec{R},t)$. This is true as long as the first-order
3159: correlation functions (coherences) are considered, where the interference
3160: effects do not distinguish between the quantum and classical theories of
3161: the EM field~\cite{bsp}. However, there are significant differences
3162: between the classical and quantum descriptions of the field in the
3163: properties of the second-order correlation function~\cite{mw,rich4}.
3164:
3165:
3166: \subsection{First-order interference}\label{ftsec61}
3167:
3168:
3169: The simplest system in which the first-order interference can be
3170: demonstrated is the Young's double slit experiment in which two
3171: light beams of amplitudes $\vec{E}_{1}(\vec{r}_{1}, t_{1})$ and
3172: $\vec{E}_{2}(\vec{r}_{2}, t_{2})$, produced at two slits located at
3173: $\vec{r}_{1}$ and $\vec{r}_{2}$, respectively, incident on a detector
3174: located at a point $\vec{R}$ far away from the slits. The resulting
3175: average intensity of the two fields measured by the detector can be
3176: written as~\cite{mw}
3177: \begin{eqnarray}
3178: \langle I(\vec{R},t)\rangle &=& \sigma \left\{\langle
3179: I_{1}(\vec{R}_{1},t-t_{1})\rangle +\langle
3180: I_{2}(\vec{R}_{2},t-t_{2})\rangle\right. \nonumber \\
3181: &&\left. + 2{\rm Re}\left\langle \vec{E}_{1}^{\ast}(\vec{R}_{1},t-t_{1})
3182: \vec{E}_{2}(\vec{R}_{2},t-t_{2})\right\rangle
3183: \right\} \ ,\label{t155}
3184: \end{eqnarray}
3185: where $\sigma$ is a constant that depends on the geometry and the size
3186: of the slits, and $I_{i}(\vec{R}_{i},t-t_{i})=
3187: \left\langle \vec{E}_{i}^{\ast}(\vec{R}_{i},t-t_{i})
3188: \vec{E}_{i}(\vec{R}_{i},t-t_{i})\right\rangle$.
3189:
3190: If the observation point $\vec{R}$ lies in the far field zone of the
3191: radiation emitted by the slits, the fields at the observation point can be
3192: approximated by plane waves for which we can write
3193: \begin{eqnarray}
3194: \vec{E}_{i}(\vec{R}_{i}, t-t_{i}) &\approx&
3195: \vec{E}_{i}(\vec{R},t)\exp \left[-i\left(\omega_{i}t -k_{i}\bar{R}\cdot
3196: \vec{r}_{i} +\phi_{i}\right)\right] \ ,\label{t156}
3197: \end{eqnarray}
3198: where $k_{i}=\omega_{i}/c$, $\omega_{i}$ is the angular frequency of the
3199: $i$th field and $\phi_{i}$ is its initial phase.
3200:
3201: For perfectly correlated fields with equal amplitudes and frequencies,
3202: and fixed the phase difference $\phi_{1}-\phi_{2}$, the average intensity
3203: detected at the point~$\vec{R}$ is given by
3204: \begin{eqnarray}
3205: \langle I(\vec{R},t)\rangle =
3206: 2\sigma \langle I_{0}\rangle\left(1+\cos k\bar{R}\cdot \vec{r}_{12}\right)
3207: \ ,\label{t157}
3208: \end{eqnarray}
3209: where $I_{0}=I_{1}=I_{2}$.
3210:
3211: Equation~(\ref{t157}) shows that the average intensity depends on the
3212: position $\bar{R}$ of the detector, and small changes in the position
3213: $\vec{R}$ of the detector lead to minima and maxima in the detected
3214: intensity. The usual measure of the minima and maxima of the intensity,
3215: called the interference fringes or interference pattern, is a visibility
3216: defined as
3217: \begin{equation}
3218: {\cal {V}} = \frac{I_{max} -I_{min}}{I_{max} +I_{min}} \ ,\label{t158}
3219: \end{equation}
3220: where $I_{max}$ corresponds to $\cos(k\bar{R}\cdot \vec{r}_{12}) =1$,
3221: whereas $I_{min}$ corresponds to $\cos (k\bar{R}\cdot \vec{r}_{12})
3222: =-1$ of the field intensity~(\ref{t157}). The visibility of the
3223: interference fringes corresponds to the degree of coherence between
3224: two fields. Hence, two classical fields
3225: of equal amplitudes and frequencies, and fixed the phase difference
3226: produce maximum possible interference pattern with the maximum
3227: visibility of 100\%.
3228:
3229: For a quantum field, the electric field components can be expressed
3230: in terms of plane waves as
3231: \begin{eqnarray}
3232: \vec{E}^{(+)}\left(\vec{r},t\right) =
3233: \left(\vec{E}^{(-)}\left(\vec{r},t\right)\right)^{\dagger} =
3234: -i\sum_{\vec{k}s}\left(\frac{\hbar\omega_{k}}{2\epsilon_{0}V}\right)
3235: ^{\frac{1}{2}}\bar{e}_{\vec{k}s}\hat{a}_{\vec{k}s}e^{i\left(\vec{k}\cdot
3236: \vec{r}-\omega_{k}t\right)} \ ,\label{t159}
3237: \end{eqnarray}
3238: where $V$ is the volume occupied by the field, $\hat{a}_{\vec{k}s}$ is
3239: the annihilation operator for the $\vec{k}$th mode of the field of the
3240: polarization $\bar{e}_{\vec{k}s}$ and $\omega_{k}$ is the angular
3241: frequency of the mode.
3242:
3243: It is easily to show, that in the case of interference of quantum
3244: fields, the average intensity detected at the point $\vec{R}$ has the
3245: same form as for the classical fields, Eq.~(\ref{t157}), with
3246: $\langle I_{0}\rangle$ given by
3247: \begin{eqnarray}
3248: \langle I_{0}\rangle &=& \sum_{\vec{k}s}\frac{\hbar\omega_{k}}
3249: {2\epsilon_{0}V}\langle\hat{n}_{\vec{k}s}\rangle \
3250: ,\label{t160}
3251: \end{eqnarray}
3252: where $\langle\hat{n}_{\vec{k}s}\rangle
3253: =\langle\hat{a}_{\vec{k}s}^{\dagger}\hat{a}_{\vec{k}s}\rangle$ is
3254: the average number of photons in the mode $\vec{k}$.
3255:
3256: Thus, interference effects involving the first-order coherences cannot
3257: distinguish between the quantum and classical theories of the EM field.
3258:
3259:
3260: \subsection{Second-order interference}\label{ftsec62}
3261:
3262:
3263: The second-order correlation function has completely different coherence
3264: properties than the first-order correlation function. An interference
3265: pattern can be observed in the second-order correlation function even if
3266: the fields are produced by two independent sources for which the phase
3267: difference $\phi_{1}-\phi_{2}$ is completely random~\cite{man65,man65a}.
3268: In this case the second-order correlation function, observed at two
3269: points $\vec{R}_{1}$ and $\vec{R}_{2}$, is given by~\cite{rich4}
3270: \begin{eqnarray}
3271: G^{(2)}(\vec{R}_{1},t_{1};\vec{R}_{2},t_{2}) &=&
3272: \langle I_{1}^{2}\left(t_{1}\right)\rangle
3273: +\langle I_{2}^{2}\left(t_{2}\right)\rangle
3274: +2\langle I_{1}\left(t_{1}\right)I_{2}\left(t_{2}\right)\rangle \nonumber \\
3275: &+& 2\langle I_{1}\left(t_{1}\right)I_{2}\left(t_{2}\right)\rangle
3276: \cos\left[ k\vec{r}_{12}\cdot
3277: \left(\bar{R}_{1}-\bar{R}_{2}\right)\right] \ .\label{t161}
3278: \end{eqnarray}
3279: Clearly, the second-order correlation function of two independent
3280: fields exhibits a cosine modulation
3281: with the separation $\vec{R}_{1}-\vec{R}_{2}$ of the two detectors. This is
3282: an interference although it involves a correlation function that is of the
3283: second order in the intensity. Similar to the first-order
3284: correlation function, the sharpness of the fringes depends on the relative
3285: intensities of the fields. For classical fields of equal intensities,
3286: $I_{1}=I_{2}=I_{0}$, the correlation function~(\ref{t161}) reduces to
3287: \begin{equation}
3288: G^{(2)}(\vec{R}_{1},t;\vec{R}_{2},t) =
3289: 4\langle I_{0}^{2}\rangle
3290: \left\{1+\frac{1}{2}\cos \left[k\vec{r}_{12}\cdot \left(\bar{R}_{1}
3291: -\bar{R}_{2}\right)\right]\right\} \ .\label{t162}
3292: \end{equation}
3293:
3294: In analogy to the visibility in the first-order correlation function, we can
3295: define the visibility of the interference pattern of the intensity
3296: correlations as
3297: \begin{equation}
3298: {\cal{V}}^{(2)} =
3299: \frac{G^{(2)}_{max}-G^{(2)}_{min}}{G^{(2)}_{max}+G^{(2)}_{min}} \ ,
3300: \label{t163}
3301: \end{equation}
3302: and find from Eq.~(\ref{t162}) that in the case of classical fields an
3303: interference pattern can be observed with the maximum possible visibility
3304: of ${\cal{V}}^{(2)}=1/2$. Thus, two independent fields of random and
3305: uncorrelated phases can exhibit an interference pattern in the intensity
3306: correlation with a maximum visibility of $50\%$.
3307:
3308: As an example of second-order interference with quantum fields, consider
3309: the simple case of two single-mode fields of equal
3310: frequencies and polarizations. Suppose that there are initially $n$ photons
3311: in the field~$E_{1}$ and $m$ photons in the field~$E_{2}$, and the state
3312: vectors of the fields are the Fock states $|\psi_{1}\rangle =|n\rangle$ and
3313: $|\psi_{2}\rangle =|m\rangle$. The initial state of the two fields is the
3314: direct product of the single-field states, $|\psi\rangle =|n\rangle|m\rangle$.
3315: Inserting Eq.~(\ref{t159}) into Eq.~(\ref{t113}) and taking the expectation
3316: value with respect to the initial state of the fields, we find
3317: \begin{eqnarray}
3318: G^{(2)}\left(\vec{R}_{1},t_{1};\vec{R}_{2},t_{2}\right) &=&
3319: \left(\frac{\hbar\omega}{2\epsilon_{0}V}\right)^{2}\left\{ n\left(n-1\right)
3320: +m\left(m-1\right)\right. \nonumber \\
3321: &&\left. +2nm\left[1+\cos k\vec{r}_{12}\cdot \left(\bar{R}_{1}
3322: -\bar{R}_{2}\right)\right]\right\} \ .\label{t164}
3323: \end{eqnarray}
3324: We note that the first two terms on the right-hand side of
3325: Eq.~(\ref{t164}) vanish when the number of photons in each field is smaller
3326: than 2, i.e. $n<2$ and $m<2$. In this limit the correlation
3327: function~(\ref{t164}) reduces to
3328: \begin{equation}
3329: G^{(2)}\left(\vec{R}_{1},t_{1};\vec{R}_{2},t_{2}\right) =
3330: 2\left(\frac{\hbar\omega}{2\epsilon_{0}V}\right)^{2}
3331: \left[1+\cos k\vec{r}_{12}\cdot \left(\bar{R}_{1}-\bar{R}_{2}
3332: \right)\right] \ .\label{t165}
3333: \end{equation}
3334: Thus, perfect interference pattern with the visibility ${\cal{V}}^{(2)}=1$
3335: can be observed in the second-order correlation function of two quantum fields
3336: each containing only one photon. According to Eq.~(\ref{t162}), the classical
3337: theory predicts only a visibility of ${\cal{V}}^{(2)}=0.5$. For $n,m\gg 1$,
3338: the first two terms on the right-hand side of Eq.~(\ref{t164}) are different
3339: from zero $(m(m-1)\approx n(n-1)\approx n^{2})$, and then the quantum
3340: correlation function (\ref{t164}) reduces to that of the classical field.
3341:
3342: The visibility of the interference pattern of the intensity correlations
3343: provides a means of testing for quantum correlations between two light fields.
3344: Mandel~{\it et al.}~\cite{mand1,mand2,mand3} have measured the visibility in
3345: the interference of signal and idler modes simultaneously generated in the
3346: process of degenerate parametric down conversion, and observed a visibility of
3347: about $75\%$, that is a clear violation of the upper bound of $50\%$ allowed by
3348: classical correlations. Richter~\cite{rich90} have extended the analysis of the
3349: visibility into the third-order correlation function, and have also found
3350: significant differences in the visibility of the interference pattern of the
3351: classical and quantum fields.
3352:
3353:
3354:
3355: \subsection{Quantum interference in two-atom systems}\label{ftsec7}
3356:
3357:
3358: In the Young's interference experiment the slits can be replaced by
3359: two atoms and interference effects can be observed between
3360: coherent or incoherent fields emitted from the atoms. The advantage
3361: of using atoms instead of slits is that a given time each atom cannot
3362: emit more than one photon. Therefore, the atoms can be regarded as
3363: sources of single photon fields.
3364:
3365: Using Eq.~(\ref{t116}), we can write the visibility as
3366: \begin{eqnarray}
3367: V &=& \frac{\left\langle S_{1}^{+}S_{2}^{-}+ S_{2}^{+}S_{1}^{-}
3368: \right\rangle}{\left\langle S_{1}^{+}S_{1}^{-}+
3369: S_{2}^{+}S_{2}^{-}\right\rangle} \ ,\label{t166}
3370: \end{eqnarray}
3371: which shows that the interference effects can be studied in terms of
3372: the atomic correlation functions.
3373:
3374: There have been several theoretical studies of the fringe visibility in
3375: the fluorescence field emitted by two coupled atoms~\cite{sckd}, and the
3376: Young's interference-type pattern has recently been observed experimentally
3377: in the resonance fluorescence of two trapped ions~\cite{eich}. The
3378: experimental results have been explained theoretically by Wong {\it et al.}
3379: \cite{wong}, and can be understood by treating the ions as independent
3380: radiators which are synchronized by the constant phase of the driving field.
3381: It has been shown that for a weak driving field, the fluorescence field is
3382: predominantly composed of an elastic component and therefore the ions behave
3383: as point sources of coherent light producing an interference pattern. Under
3384: strong excitation the fluorescence field is mostly composed of the incoherent
3385: part and consequently there is no interference pattern. To show this,
3386: we consider a two-atom system driven by a coherent laser field
3387: propagating in the direction perpendicular to the interatomic axis. In
3388: this case, we can use the master equation~(\ref{t42}) and obtain the analytical
3389: formula for the fringe visibility of the steady-state fluorescence field
3390: as~\cite{fr}
3391: \begin{eqnarray}
3392: V &=& \frac{\left(\Gamma^{2}+4\Delta_{L}^{2}\right)}
3393: {\left(\Gamma^{2}+4\Delta_{L}^{2}\right)
3394: +2\Omega^{2}} \ .\label{t167}
3395: \end{eqnarray}
3396: It is seen that in this specific case, the visibility is positive for
3397: all parameter values and is independent of the interatomic
3398: interactions. For a weak driving field, $\Omega \ll \Gamma,
3399: \Delta_{L}$, the fringe visibility $|V|\approx 1$, whereas $|V|\approx 0$
3400: for $\Omega \gg \Gamma, \Delta_{L}$, showing that and there is no
3401: interference pattern when the atoms are driven by a strong field.
3402: For moderate Rabi frequencies, $\Omega \approx \Gamma$, the
3403: visibility may be improved by detuning the laser field from the atomic
3404: resonance. Kochan {\it et al.} \cite{kochan} have shown that the
3405: interference pattern of the strongly driven atoms can also be improved by
3406: placing the atoms inside an optical cavity. The coupling of the atoms to
3407: the cavity mode induces atomic correlations which improves the fringe
3408: visibility.
3409:
3410: Here, we derive general criteria for the first- and second-order
3411: interference in the fluorescence field emitted from two two-level
3412: atoms. Using these criteria, we can easily predict conditions for
3413: quantum interference in the two atom system. In this approach, we
3414: apply the collective states of a two-atom system, and write the atomic
3415: correlation functions in terms of the density matrix elements of the
3416: collective system as
3417: \begin{eqnarray}
3418: \left\langle S_{1}^{+}S_{1}^{-}\right\rangle +
3419: \left\langle S_{2}^{+}S_{2}^{-}\right\rangle &=&
3420: \rho_{ss}+\rho_{aa}+2\rho_{ee} \ ,\nonumber \\
3421: \left\langle S_{1}^{+}S_{2}^{-}\right\rangle &=&
3422: \frac{1}{2}\left(\rho_{ss}-\rho_{aa}+\rho_{as}-\rho_{sa}\right)
3423: \ ,\nonumber \\
3424: \left\langle S_{1}^{+}S_{2}^{+}S_{1}^{-}S_{2}^{-}\right\rangle &=&
3425: \rho_{ee} \ ,\label{t168}
3426: \end{eqnarray}
3427: where $\rho_{ii} (i=a,s,e)$ are the populations of the collective
3428: states and $\rho_{sa}, \rho_{as}$ are coherences.
3429:
3430: From the relations~(\ref{t168}), we find that in terms of the density
3431: matrix elements the first-order correlation function can be written as
3432: \begin{eqnarray}
3433: G^{\left( 1\right) }\left(\vec{R},t\right)
3434: &=& \Gamma u(\vec{R})\left\{ 2\rho_{ee}\left(t\right)
3435: +\rho_{ss}\left(t\right)\left(1
3436: +\cos k\bar{R}\cdot \vec{r}_{12}\right)\right. \nonumber \\
3437: &+&\left. \rho_{aa}\left(t\right)\left(1-\cos
3438: k\bar{R}\cdot \vec{r}_{12}\right)\right. \nonumber \\
3439: &+&\left. i\left(\rho_{sa}\left(t\right)-\rho_{as}\left(t\right)\right)
3440: \sin k\bar{R}\cdot \vec{r}_{12}\right\} \ ,\label{t169}
3441: \end{eqnarray}
3442: and the second-order correlation function takes the form
3443: \begin{eqnarray}
3444: G^{\left( 2\right)}\left( \vec{R}_{1},t;\vec{R}_{2},t\right)
3445: &=& 4\Gamma^{2}u(\vec{R}_{1}) u(\vec{R}_{2})
3446: \nonumber \\
3447: &\times& \rho_{ee}\left(t\right)\left[1+\cos k\left(\bar{R}_{1}
3448: -\bar{R}_{2}\right)\cdot \vec{r}_{12}\right] \ .\label{t170}
3449: \end{eqnarray}
3450: It is evident from Eq.~(\ref{t169}) that first-order correlation
3451: function can exhibit an interference pattern only if $\rho_{ss}\neq
3452: \rho_{aa}$ and/or Im$(\rho_{sa})\neq 0$. This happens when $\langle
3453: e_{1}|\langle g_{2}|\hat{\rho} |e_{2}\rangle |g_{1}\rangle$ and
3454: $\langle g_{1}|\langle e_{2}|\hat{\rho} |g_{2}\rangle |e_{1}\rangle$
3455: are different from zero, i.e. when there are nonzero coherences between
3456: the atoms. Sch\"{o}n and Beige~\cite{sb} have arrived to the same
3457: conclusion using the quantum jump method. On the other hand, the
3458: second-order correlation function is independent of the populations of
3459: the entangled states $\rho_{ss}, \rho_{aa}$ and the coherences, and
3460: exhibit an interference pattern when $\rho_{ee}(t)\neq 0$.
3461:
3462: We now examine some specific processes in which one can create unequal
3463: populations of the $|s\rangle$ and $|a\rangle$ states. Dung and
3464: Ujihara~\cite{du} have shown that spontaneous emission from two
3465: identical atoms, with initially only one atom excited, can exhibit an
3466: interference pattern. Their results can be easily interpreted in terms
3467: of the populations $\rho_{ss}(t)$ and $\rho_{aa}(t)$. If initially
3468: only one atom was excited; $\rho_{ee}(0)=0$ and
3469: $\rho_{ss}(0)=\rho_{aa}(0)=\rho_{sa}(0)=\rho_{as}(0)=\frac{1}{2}$.
3470: Using the master
3471: equation~(\ref{t42}) with $\Omega_{1}=\Omega_{2}=0$, we find that the
3472: time evolution of the populations $\rho_{ss}(t)$ and $\rho_{aa}(t)$ is
3473: given by
3474: \begin{eqnarray}
3475: \rho_{ss}\left(t\right) &=& \rho_{ss}\left(0\right) \exp \left[-\left(\Gamma
3476: +\Gamma_{12}\right)t\right] \ ,\nonumber \\
3477: \rho_{aa}\left(t\right) &=& \rho_{aa}\left(0\right) \exp \left[-\left(\Gamma
3478: -\Gamma_{12}\right)t\right] \ .\label{t171}
3479: \end{eqnarray}
3480: Since the populations decay with different rates, the symmetric state
3481: decays with an enhanced rate $\Gamma +\Gamma_{12}$, while the
3482: antisymmetric state decays with a reduced rate $\Gamma -\Gamma_{12}$,
3483: the populations $\rho_{aa}(t)$ is larger than $\rho_{ss}(t)$ for all
3484: $t>0$. Hence, an interference pattern can be observed for $t>0$.
3485: This effect arises from the presence of the interatomic
3486: interactions $(\Gamma_{12}\neq 0)$. Thus, for two independent atoms
3487: the populations decay with the same rate resulting in the
3488: disappearance of the interference pattern.
3489:
3490: When the atoms are driven by a coherent laser field, an interference
3491: pattern can be observed even in the absence of the interatomic
3492: interactions. To show this, we consider the steady-state
3493: solutions~(\ref{t150}) for the populations of
3494: the collective atomic states.
3495: It is evident from Eq.~(\ref{t150}) that $\rho_{ss}>\rho_{aa}$ even in
3496: the absence of the interatomic interactions
3497: $(\Gamma_{12}=\Omega_{12}=0)$. Hence, an interference pattern can be
3498: observed even for two independent atoms. In this case the
3499: interference pattern results from the coherent synchronization of the
3500: oscillations of the atoms by the constant coherent phase of the
3501: driving laser field.
3502:
3503: We have shown that the first-order coherence is sensitive to the
3504: interatomic interactions and the excitation field. In contrast, the
3505: second-order correlation function can exhibit an interference pattern
3506: independent of the interatomic interactions and the excitation
3507: process~\cite{ftk88,sko,sko1}. According to Eq.(\ref{t170}), to observe
3508: an interference pattern in the second-order correlation function, it is
3509: enough to produce a non-zero population in the state $|e\rangle$. The
3510: interference results from the detection process that a detector does not
3511: distinguish between two simultaneously detected photons. As an
3512: example, consider spontaneous emission from two identical and also
3513: nonidentical atoms with initially both atoms excited.
3514:
3515: For two identical atoms, we can easily find from Eqs.~(\ref{t42}) and
3516: (\ref{t170}), and the quantum regression theorem~\cite{lax}, that the
3517: two-time second-order correlation function is given by
3518: \begin{eqnarray}
3519: G^{\left( 2\right)}\left( \vec{R}_{1},t;\vec{R}_{2},t+\tau \right)
3520: &=& \frac{1}{2}\Gamma^{2}u(\vec{R}_{1}) u(\vec{R}_{2})
3521: \exp\left[-\Gamma \left(2t +\tau \right)\right] \nonumber \\
3522: &\times& \left\{\left[1+\cos \left(k\bar{R}_{1}\cdot
3523: \vec{r}_{12}\right) \cos \left(k\bar{R}_{2}\cdot
3524: \vec{r}_{12}\right)\right]\cosh \left(\Gamma_{12}\tau \right)\right.
3525: \nonumber \\
3526: &-&\left. \left[\cos \left(k\bar{R}_{1}\cdot \vec{r}_{12}\right)
3527: +\cos \left(k\bar{R}_{2}\cdot \vec{r}_{12}\right)\right] \sinh
3528: \left(\Gamma_{12}\tau \right)\right. \nonumber \\
3529: &+&\left. \sin \left(k\bar{R}_{1}\cdot
3530: \vec{r}_{12}\right) \sin \left(k\bar{R}_{2}\cdot
3531: \vec{r}_{12}\right) \cos \left(2\Omega_{12}\tau \right)\right\}
3532: \ .\label{t172}
3533: \end{eqnarray}
3534: The above equation shows that the two-time second-order correlation
3535: function exhibits a sinusoidal modulation in space and time.
3536: This modulation can be interpreted both in terms of interference
3537: fringes and quantum beats. The frequency of quantum beats is
3538: $2\Omega_{12}$ and the amplitude of these beats depends on the
3539: direction of observation in respect to the interatomic axis.
3540: The quantum beats vanish for directions $\theta_{1}=90^{o}$ or
3541: $\theta_{2}=90^{o}$, where $\theta_{1}(\theta_{2})$ is the angle
3542: between $\vec{r_{12}}$ and $\bar{R}_{1}(\bar{R}_{2})$, and the
3543: amplitude of the beats has its maximum for two photons detected in
3544: the direction $\theta_{1}=\theta_{2}=0^{o}$. This directional effect
3545: is connected with the fact that the antisymmetric state $\ket a$ does
3546: not radiate in the direction perpendicular to the interatomic axis. We
3547: will discuss this directional effect in more details in
3548: Sec.~\ref{ftsec91}. For independent atoms, $\Gamma_{12}=0,
3549: \Omega_{12}=0$, and then the correlation function~(\ref{t172})
3550: reduces to
3551: \begin{eqnarray}
3552: G^{\left( 2\right)}\left( \vec{R}_{1},t;\vec{R}_{2},t+\tau \right)
3553: &=& \frac{1}{2}\Gamma^{2}u(\vec{R}_{1}) u(\vec{R}_{2})
3554: \exp\left[-\Gamma \left(2t +\tau \right)\right] \nonumber \\
3555: &\times& \left[1+\cos k\left(\bar{R}_{1}
3556: -\bar{R}_{2}\right)\cdot \vec{r}_{12}\right] \ ,\label{t173}
3557: \end{eqnarray}
3558: which shows that the time modulation vanishes. This implies that quantum
3559: beats are absent in spontaneous emission from two independent atoms, but
3560: the spatial modulation is still present.
3561:
3562: The situation is different for two nonidentical atoms. In this case,
3563: the two-time second-order correlation function exhibits quantum beats
3564: even if the atoms are independent. For $\Gamma_{12}=0$ and
3565: $\Omega_{12}=0$, the master equation~(\ref{t42}) leads to the
3566: following correlation function
3567: \begin{eqnarray}
3568: G^{\left( 2\right)}\left( \vec{R}_{1},t;\vec{R}_{2},t+\tau \right)
3569: &=& \frac{1}{2}\Gamma^{2}u(\vec{R}_{1}) u(\vec{R}_{2})
3570: \exp\left[-\Gamma \left(2t +\tau \right)\right] \nonumber \\
3571: &\times& \left\{\cosh
3572: \frac{1}{2}\left(\Gamma_{2}-\Gamma_{1}\right)\tau \right. \nonumber \\
3573: &+&\left. \cos \left[ k\left(\bar{R}_{1}
3574: -\bar{R}_{2}\right)\cdot \vec{r}_{12} -2\Delta \tau \right]\right\}
3575: \ .\label{t174}
3576: \end{eqnarray}
3577: Thus, for independent nonidentical atoms, the correlation function
3578: shows a sinusoidal modulation both in space and time. We note that
3579: the modulation term in Eq.~(\ref{t174}) is the same as that obtained
3580: by Mandel~\cite{man64}, who considered the second-order
3581: correlation function for two beams emitted by independent lasers.
3582:
3583:
3584: \section{Selective excitation of the collective atomic
3585: states}\label{ftsec8}
3586:
3587:
3588: In the previous section, we have shown that nonclassical effects in
3589: coherently driven two-atom systems reflect the preparation of the system
3590: in a superposition of two collective states. In particular, for the
3591: total photon anticorrelation and maximum squeezing, the two-atom
3592: system is in a superposition of the ground and the entangled
3593: symmetric states. The other states are not populated. We now consider
3594: excitation processes which can lead to a preparation of the two-atom
3595: system in only one of the collective states. In particular, we will focus on
3596: processes which can prepare the two-atom system in the entangled symmetric
3597: state $\ket s$. Our main interest, however, is in the preparation of the
3598: system in the maximally entangled antisymmetric state $\ket a$ which, under
3599: the condition $\Gamma_{12}=\sqrt{\Gamma_{1}\Gamma_{2}}$, is a
3600: decoherence-free state. The central idea is to choose the distance
3601: between the atoms such that the resulting level shift is large enough
3602: to consider the possible transitions between the collective states
3603: separately. This will allow to make a selective excitation of the
3604: symmetric and antisymmetric states and therefore to create controlled
3605: entanglement between the atoms.
3606:
3607:
3608: \subsection{Preparation of the symmetric state by a pulse
3609: laser}\label{ftsec81}
3610:
3611:
3612: Beige {\it et al.}~\cite{be} have shown that a system of two identical
3613: two-level atoms may be prepared in the symmetric state $\ket s$ by a short
3614: laser pulse. The conditions for a selective excitation of
3615: the collective atomic states can be analyzed from the interaction
3616: Hamiltonian of the laser field with the two-atom system. We make the
3617: unitary transformation
3618: \begin{equation}
3619: \tilde{H}_{L} = e^{i\hat{H}_{a}t/\hbar}\hat{H}_{L}e^{-i\hat{H}_{a}t/\hbar}
3620: \ ,\label{t175}
3621: \end{equation}
3622: where
3623: \begin{eqnarray}
3624: \hat{H}_{a} &=& \hbar \left\{ \Delta_{L}\left( \ket e\bra e -\ket g
3625: \bra g \right)
3626: +\left(\Delta_{L}+\Omega_{12}\right)\ket s\bra s\right. \nonumber \\
3627: &&+\left.\left(\Delta_{L}-\Omega_{12}\right)\ket a\bra a\right\}
3628: \ ,\label{t176}
3629: \end{eqnarray}
3630: and find that in the case of identical atoms, $\Gamma_{1}=\Gamma_{2}$ and
3631: $\Delta =0$, the transformed interaction Hamiltonian $\tilde{H}_{L}$
3632: is given by
3633: \begin{eqnarray}
3634: \tilde{H}_{L} &=& -\frac{\hbar}{2\sqrt{2}}\left\{
3635: \left(\Omega_{1}+\Omega_{2}\right)\left(S^{+}_{es}
3636: e^{i\left(\Delta_{L}+\Omega_{12}\right)t}
3637: +S^{+}_{sg}e^{i\left(\Delta_{L}-\Omega_{12}\right)t}\right)\right. \nonumber \\
3638: &+&\left. \left(\Omega_{2}-\Omega_{1}\right)\left(S^{+}_{ag}
3639: e^{i\left(\Delta_{L}+\Omega_{12}\right)t}
3640: +S^{+}_{ea}e^{i\left(\Delta_{L}-\Omega_{12}\right)t}\right)
3641: +{\rm H.c.}\right\} \ .\label{t177}
3642: \end{eqnarray}
3643: The Hamiltonian~(\ref{t177}) represents the interaction of the laser field with
3644: the collective two-atom system, and in the transformed form contains terms
3645: oscillating at frequencies $(\Delta_{L}\pm \Omega_{12})$, which correspond to
3646: the two separate groups of transitions between the collective atomic
3647: states at frequencies $\omega_{L}=\omega_{0}+\Omega_{12}$ and
3648: $\omega_{L}=\omega_{0}-\Omega_{12}$.
3649: The $\Delta_{L}+\Omega_{12}$ frequencies are separated from
3650: $\Delta_{L}-\Omega_{12}$ frequencies by $2\Omega_{12}$, and hence the
3651: two groups of the transitions evolve separately when $\Omega_{12}\gg \Gamma$.
3652: Depending on the frequency, the laser can be selectively tuned to one
3653: of the two groups of the transitions.
3654: When $\omega_{L}=\omega_{0}+\Omega_{12}$ $(\Delta_{L}-\Omega_{12}=0)$ the laser
3655: is tuned to exact resonance with the $\ket e -\ket a$ and $\ket g -\ket s$
3656: transitions, and then the terms, appearing in the
3657: Hamiltonian~(\ref{t177}), and corresponding to these transitions have no
3658: explicit time dependence. In contrast,
3659: the $\ket g -\ket a$ and $\ket e -\ket s$ transitions are
3660: off-resonant and the terms corresponding to these transitions have an
3661: explicit time dependence exp$\left(\pm 2i\Omega_{12}t\right)$.
3662: If $\Omega_{12}\gg \Gamma$, the off-resonant terms rapidly oscillate with
3663: the frequency $2\Omega_{12}$, and then we can make a secular
3664: approximation in which we neglect all those rapidly
3665: oscillating terms. The interaction Hamiltonian can then be written in
3666: the simplified form
3667: \begin{equation}
3668: \tilde{H}_{L} = -\frac{\hbar}{2\sqrt{2}}\left[
3669: \left(\Omega_{1}+\Omega_{2}\right)S^{+}_{sg}
3670: + \left(\Omega_{2}-\Omega_{1}\right)S^{+}_{ea} + {\rm H.c.}\right]
3671: \ .\label{t178}
3672: \end{equation}
3673: It is seen that the laser field couples to the transitions with significantly
3674: different Rabi frequencies. The coupling strength of the laser to the
3675: $\ket g -\ket s$ transition is proportional to the sum of the Rabi frequencies
3676: $\Omega_{1}+\Omega_{2}$, whereas the coupling strength of the laser to the
3677: $\ket a -\ket e$ transition is proportional to the difference of the Rabi
3678: frequencies $\Omega_{1}-\Omega_{2}$. According to Eq.~(\ref{t46}) the Rabi
3679: frequencies $\Omega_{1}$ and $\Omega_{2}$ of two identical atoms differ only
3680: by the phase factor exp$(i\vec{k}_{L}\cdot \vec{r}_{12})$. Thus, in order to
3681: selectively excite the $\ket g -\ket s$ transition, the driving
3682: laser field should be in phase with both atoms, i.e. $\Omega_{1}=\Omega_{2}$.
3683: This can be achieved by choosing the propagation
3684: vector $\vec{k}_{L}$ of the laser orthogonal to the line joining the atoms.
3685: Under this condition we can make a further simplification and truncate the
3686: state vector of the system into two states $\ket g$ and $\ket s$.
3687: In this two-state approximation we find from the Schr\"{o}dinger
3688: equation the time evolution of the population $P_{s}(t)$ of the state
3689: $\ket s$ as
3690: \begin{equation}
3691: P_{s}\left(t\right) = \sin^{2}\left(\frac{1}{\sqrt{2}}\Omega t\right)
3692: \ ,\label{t179}
3693: \end{equation}
3694: where $\Omega =\Omega_{1}=\Omega_{2}$.
3695:
3696: The population oscillates with the Rabi frequency of the $\ket s
3697: -\ket g$ transition and at certain times
3698: $P_{s}(t)=1$ indicating that all the population is in the symmetric state.
3699: This happens at times
3700: \begin{equation}
3701: T_{n} = \left(2n+1\right)\frac{\pi}{\sqrt{2}\Omega} ,\quad n=0,1,\ldots .
3702: \label{t180}
3703: \end{equation}
3704: Hence, the system can be prepared in the state $\ket s$ by simply applying
3705: a laser pulse, for example, with the duration $T_{0}$, that is a standard
3706: $\pi$ pulse.
3707:
3708: The two-state approximation is of course an idealization, and a
3709: possibility that all the transitions can be driven by the laser
3710: imposes significant limits on the Rabi frequency and the duration
3711: of the pulse. Namely, the Rabi frequency cannot be too strong in order to
3712: avoid the
3713: coupling of the laser to the $\ket s -\ket e$ transition, which could lead to
3714: a slight pumping of the population to the state $\ket e$. On the other hand,
3715: the Rabi frequency cannot be too small as for a small $\Omega$ the duration
3716: of the pulse, required for the complete
3717: transfer of the population into the state $\ket s$, becomes longer and
3718: then spontaneous emission can occur during the excitation process. Therefore,
3719: the transfer of the population to the state $\ket s$ cannot be made
3720: arbitrarily fast and, in addition, requires a careful estimation
3721: of the optimal Rabi frequency, which could be difficult to achieve in a real
3722: experimental situation.
3723:
3724:
3725: \subsection{Preparation of the antisymmetric state}\label{ftsec82}
3726:
3727: \subsubsection{Pulse laser}\label{ftsec821}
3728:
3729: If we choose the laser frequency such that $\Delta_{L}+\Omega_{12}=0$, the
3730: laser field is then resonant to the $\ket a -\ket g$ and $\ket e -\ket s$
3731: transitions and, after the secular approximation, the
3732: Hamiltonian~(\ref{t177}) reduces to
3733: \begin{equation}
3734: \tilde{H}_{L} = -\frac{\hbar}{2\sqrt{2}}\left[
3735: \left(\Omega_{2}-\Omega_{1}\right)S^{+}_{ag}
3736: + \left(\Omega_{1}+\Omega_{2}\right)S^{+}_{es} + {\rm H.c.}\right]
3737: \ .\label{t181}
3738: \end{equation}
3739: Clearly, for $\Omega_{1}=-\Omega_{2}$ the laser couples only to the
3740: $\ket a -\ket g$ transition. Thus, in order to selectively excite the
3741: $\ket g -\ket a$ transition, the atoms should experience opposite phases
3742: of the laser field. This can be achieved by choosing the propagation
3743: vector $\vec{k}_{L}$ of the laser along the interatomic axis, and the
3744: atomic separations such that
3745: \begin{equation}
3746: \vec{k}_{L}\cdot \vec{r}_{12} = \left(2n+1\right)\pi ,\quad n=0,1,2,\ldots
3747: \ ,\label{t182}
3748: \end{equation}
3749: which corresponds to a situation that the atoms are separated by a distance
3750: $r_{12}=(2n+1)\lambda/2$.
3751:
3752: The smallest distance at which the atoms could experience opposite phases
3753: corresponds to $r_{12}=\lambda/2$. However, at this particular separation
3754: the dipole-dipole interaction parameter $\Omega_{12}$ is small, see
3755: Fig.~\ref{ftfig1}, and then all of the transitions between the collective
3756: states occur at approximately the same frequency. In this case the secular
3757: approximation is not valid and we cannot separate the transitions at
3758: $\Delta_{L}+\Omega_{12}$ from the transitions at $\Delta_{L}-\Omega_{12}$.
3759:
3760: One possible solution to the problem of the selective excitation
3761: with opposite phases is to use a standing laser field instead of the
3762: running-wave field.
3763: If the laser amplitudes differ by the sign, i.e. $\vec{E}_{L_{1}}=
3764: -\vec{E}_{L_{2}}=\vec{E}_{0}$, and $\vec{k}_{L_{1}}\cdot \vec{r}_{1} =
3765: -\vec{k}_{L_{2}}\cdot \vec{r}_{2}$, the Rabi frequencies experienced by the
3766: atoms are
3767: \begin{eqnarray}
3768: \Omega_{1} &=& \frac{2i}{\hbar} \vec{\mu}_{1} \cdot \vec{E}_{0}
3769: \sin \left(\frac{1}{2}\vec{k}_{L}\cdot \vec{r}_{12}\right) , \nonumber \\
3770: \Omega_{2} &=& -\frac{2i}{\hbar} \vec{\mu}_{2}\cdot \vec{E}_{0}
3771: \sin \left(\frac{1}{2}\vec{k}_{L}\cdot \vec{r}_{12}\right) \
3772: ,\label{t183}
3773: \end{eqnarray}
3774: where $\vec{k}_{L}=\vec{k}_{L_{1}}=\vec{k}_{L_{2}}$ and
3775: we have chosen the reference frame such that $\vec{r}_{1}=\frac{1}{2}
3776: \vec{r}_{12}$ and $\vec{r}_{2}=-\frac{1}{2}\vec{r}_{12}$. It follows
3777: from Eq.~(\ref{t183}) that the Rabi frequencies oscillate with opposite
3778: phases independent of the separation between the atoms. However, the
3779: magnitude of the Rabi frequencies decreases with decreasing $r_{12}$.
3780:
3781:
3782: \subsubsection{Indirect driving through the symmetric
3783: state}\label{ftsec822}
3784:
3785:
3786: We now turn to the situation of non-identical atoms and consider
3787: different possible processes of the population transfer to the
3788: antisymmetric state which could be present even if the antisymmetric
3789: state does not decay to the ground level. This can happen when
3790: $\Gamma_{12}=\sqrt{\Gamma_{1}\Gamma_{2}}$, i.e. when the separation
3791: between the atoms is negligible small.
3792: Under this condition the antisymmetric state is also decoupled from
3793: the driving field. According to Eq.~(\ref{t95}), the antisymmetric state can
3794: still be coupled, through the coherent interaction $\Delta_{c}$, to the
3795: symmetric state $\ket s$. However, this coupling appears only for nonidentical
3796: atoms.
3797:
3798: From the master equation~(\ref{t42}), we find that under the condition
3799: $\Gamma_{12}=\sqrt{\Gamma_{1}\Gamma_{2}}$ the equation of motion for the
3800: population of the antisymmetric state $\ket a$ is given by~\cite{afs}
3801: \begin{eqnarray}
3802: \dot{\rho}_{aa} &=& \frac{\left(\Gamma_{1}-\Gamma_{2}\right)^{2}}
3803: {\Gamma_{1}+\Gamma_{2}}\rho_{ee}
3804: +i\Delta _{c}\left( \rho _{sa}-\rho _{as}\right)
3805: \nonumber \\
3806: &&-\frac{1}{2}i\Omega \frac{\left(\Gamma_{1}-\Gamma_{2}\right)}
3807: {\sqrt{\Gamma_{1}^{2}+\Gamma_{2}^{2}}}\left(
3808: \rho _{ea}-\rho_{ae}\right) \ . \label{t184}
3809: \end{eqnarray}
3810: The equation (\ref{t184}) shows that the non-decaying antisymmetric state
3811: $\ket a$ can be populated by spontaneous emission from the
3812: upper state $\ket e$ and also by the coherent interaction
3813: with the state $\ket s$. The first condition is satisfied
3814: only when $\Gamma_{1}\neq \Gamma_{2}$, while the other condition is satisfied
3815: only when $\Delta _{c}\neq 0$. Thus, the transfer of population to the state
3816: $\ket a$ from the upper state $\ket e$ and the symmetric
3817: state $\ket s$ does not appear when the atoms are identical,
3818: but is possible for nonidentical atoms.
3819: %
3820: \begin{figure}[t]
3821: \begin{center}
3822: \includegraphics[width=10cm]{ftfig13.eps}
3823: \end{center}
3824: \caption{The steady-state population of the maximally entangled
3825: antisymmetric state
3826: $\ket a$ for $\Omega =10\Gamma_{1}, \Omega_{12}=10\Gamma_{1}$ and
3827: $\Gamma_{2}=\Gamma_{1}, \Delta =\Gamma_{1}$ (solid line),
3828: $\Gamma_{2}= 2\Gamma_{1}, \Delta =0$ (dashed line).}
3829: \label{ftfig13}
3830: \end{figure}
3831:
3832: We illustrate this effect in Fig.~\ref{ftfig13}, where we plot the
3833: steady-state population of the maximally entangled state
3834: $\ket a$ as a function of $\Delta _{L}$ for two different types
3835: of nonidentical atoms. In the first case the atoms have the same damping
3836: rates $(\Gamma_{1}=\Gamma_{2})$ but different transition frequencies
3837: $(\Delta \neq 0)$, while in the
3838: second case the atoms have the same frequencies $(\Delta =0)$ but different
3839: damping rates $(\Gamma_{1}\neq \Gamma_{2})$. It is seen from
3840: Fig.~\ref{ftfig13} that in both
3841: cases the antisymmetric state can be populated even if is not directly driven
3842: from the ground state. The population is transferred to $\ket a$
3843: through the coherent interaction $\Delta_{c}$ which leaves the other excited
3844: states completely unpopulated.
3845: %
3846: \begin{figure}[t]
3847: \begin{center}
3848: \includegraphics[width=10cm]{ftfig14.eps}
3849: \end{center}
3850: \caption{The steady-state populations of the upper state $\ket e$ (solid line)
3851: and the symmetric state $\ket s$ (dashed line) for
3852: $\Gamma_{2}=\Gamma_{1}, \Omega =10\Gamma_{1}, \Omega_{12}=10\Gamma_{1}$
3853: and $\Delta =\Gamma_{1}$.}
3854: \label{ftfig14}
3855: \end{figure}
3856: %
3857: This is shown in Fig.~\ref{ftfig14}, where we plot the steady-state populations
3858: $\rho _{ss}$ and $\rho_{ee}$ of the states $\ket s$ and $\ket e$. It is
3859: apparent from Fig.~\ref{ftfig14} that at $\Delta_{L}= -\Omega_{12}$ the states
3860: $\ket s$ and $\ket e$ are not populated. However, the population is not
3861: entirely trapped in the antisymmetric state $\ket a$, but rather in a linear
3862: superposition of the antisymmetric and ground states. This is
3863: illustrated in Fig.~\ref{ftfig15}, where we plot the steady-state population
3864: $\rho_{aa}$ for the same parameters as in Fig.~\ref{ftfig14}, but different
3865: $\Omega $. Clearly, for a small $\Omega$ the steady-state population
3866: $\rho_{aa}\approx \frac{1}{2}$, and the amount of the population increases with
3867: increasing $\Omega$. The population $\rho_{aa}$ attains the maximum value
3868: $\rho_{aa}\approx 1$ for a very strong driving field.
3869: %
3870: \begin{figure}[ht]
3871: \begin{center}
3872: \includegraphics[width=10cm]{ftfig15.eps}
3873: \end{center}
3874: \caption{The steady-state population of the antisymmetric state $\ket a$
3875: for $\Gamma_{2}=\Gamma_{1}, \Omega_{12}=10\Gamma_{1}, \Delta
3876: =\Gamma_{1}$ and different $\Omega$: $\Omega =\Gamma_{1}$ (solid
3877: line), $\Omega =5\Gamma_{1}$ (dashed line), $\Omega =20\Gamma_{1}$
3878: (dashed-dotted line). }
3879: \label{ftfig15}
3880: \end{figure}
3881:
3882: This result shows that we can relatively easily prepare two nonidentical atoms
3883: in the maximally entangled antisymmetric state. The
3884: closeness of the prepared state to the ideal one is measured by the fidelity
3885: $F$. Here $F$ is equal to the obtained maximum population in the state $%
3886: \left| a\right\rangle $. For $\Omega \gg \Gamma$ the fidelity of the prepared
3887: state is maximal, equal to~1. As we have already mentioned, the system has the
3888: advantage that the maximally entangled state $\left| a\right\rangle $ does not
3889: decay, i.e. is a decoherence-free state.
3890:
3891:
3892: \subsubsection{Atom-cavity-field interaction}\label{ftsec823}
3893:
3894:
3895: There have been several proposals to generate the antisymmetric
3896: state $\ket a$ in a system of two identical atoms interacting with a
3897: single-mode cavity field. In this case, collective effects arise
3898: from the interaction between the atoms induced by a strong coupling
3899: of the atoms to the cavity mode. An excited atom emits a photon
3900: into the cavity mode that is almost immediately absorbed by the
3901: second atom.
3902: Plenio {\it et al.}~\cite{phbk}
3903: have considered a system of two atoms trapped inside an optical
3904: cavity and separated by a distance much larger than the optical
3905: wavelength. This allows for the selective excitation of only one of
3906: the atoms. In this scheme the generation of the antisymmetric state
3907: relies on the concept of conditional dynamics due to continuous
3908: observation of the cavity field. If only one atom is excited and no
3909: photon is detected outside the cavity, the atoms are prepared in a dark
3910: state~\cite{yzm}, which is equivalent to the antisymmetric state
3911: $\ket a$.
3912:
3913: In earlier studies, Phoenix and Barnett~\cite{pb}, Kudrayvtsev and
3914: Knight~\cite{kk} and
3915: Cirac and Zoller~\cite{cz} have analyzed two-atom Jaynes-Cummings
3916: models for a violation of Bell's inequality, and have shown that the
3917: atoms moving across a single-mode cavity can be prepared in the
3918: antisymmetric state via the interaction with the cavity field. In this
3919: scheme, the preparation of the antisymmetric state takes place in two
3920: steps. In the first step, one atom initially prepared in its excited
3921: state $\ket {e_{1}}$ is sent through a single-mode cavity being in the
3922: vacuum state $\ket 0_{c}$. During the interaction with the cavity mode,
3923: the atomic population undergoes the vacuum Rabi oscillations, and the
3924: interaction time was varied by selecting different atomic velocities.
3925: If the velocity of the atom is such that the interaction time of the
3926: atom with the cavity mode is equal to quarter of the vacuum Rabi
3927: oscillations, the state of the combined system, the atom plus the
3928: cavity mode, is a superposition state
3929: \begin{equation}
3930: \ket {a_{1}c} =\frac{1}{\sqrt{2}}\left(\ket {e_{1}}\ket 0_{c} -\ket
3931: {g_{1}}\ket 1_{c}\right) \ .\label{t185}
3932: \end{equation}
3933: Hence, the state of the total system, two atoms plus the cavity mode,
3934: after the first atom has crossed the cavity is
3935: \begin{equation}
3936: \ket {\Psi_{1}} =\frac{1}{\sqrt{2}}\left(\ket {e_{1}}\ket 0_{c}
3937: -\ket {g_{1}}\ket 1_{c}\right)\ket {g_{2}} \ .\label{t186}
3938: \end{equation}
3939: If we now send the second atom, being in its ground state, with the
3940: selected velocity such that during the interaction with the cavity
3941: mode the atom undergoes half of the vacuum Rabi oscillation, the final
3942: state of the system becomes
3943: \begin{eqnarray}
3944: \ket {\Psi_{12}c} &=& \frac{1}{\sqrt{2}}\left(\ket {e_{1}}\ket
3945: 0_{c}\ket {g_{2}} -\ket {g_{1}}\ket 0_{c}\ket {e_{2}}\right)
3946: \nonumber \\
3947: &=& \frac{1}{\sqrt{2}}\left(\ket {e_{1}}\ket {g_{2}}
3948: -\ket {g_{1}}\ket {e_{2}}\right)\ket 0_{c} = \ket a \ket 0_{c}
3949: \ .\label{t187}
3950: \end{eqnarray}
3951: Thus, the final state of the system is a product state of the atomic
3952: antisymmetric state $\ket a$ and the vacuum state of the cavity mode.
3953: In this scheme the cavity mode is left in the vacuum state which
3954: prevents the antisymmetric state from any noise of the cavity. The
3955: scheme to entangle two atoms in a cavity, proposed by Cirac and
3956: Zoller~\cite{cz}, has recently been realized experimentally by Hagly
3957: {\it et al.}~\cite{hag}.
3958:
3959: Gerry~\cite{ge} has proposed a similar method based on a dispersive
3960: interaction of the atoms with a cavity mode prepared in a coherent
3961: state $\ket \alpha$. The atoms enter the cavity in superposition states
3962: \begin{eqnarray}
3963: \ket {a_{1}} &=& \frac{1}{\sqrt{2}}\left(\ket {e_{1}} +i\ket
3964: {g_{1}}\right) \ , \nonumber \\
3965: \ket {a_{2}} &=& \frac{1}{\sqrt{2}}\left(\ket {e_{2}} -i\ket
3966: {g_{2}}\right) \ .\label{t188}
3967: \end{eqnarray}
3968: After passage of the second atom, the final state of the system is
3969: \begin{eqnarray}
3970: \ket {\Psi_{12}c} &=& \frac{1}{2}\left\{ \left(\ket {g_{1}}\ket {g_{2}}
3971: +\ket {e_{1}}\ket {e_{2}}\right)\ket {-\alpha}\right. \nonumber \\
3972: &&+\left. i\left(\ket {e_{1}}\ket {g_{2}} -\ket {g_{1}}\ket
3973: {e_{2}}\right)\ket {\alpha}\right\} \ .\label{t189}
3974: \end{eqnarray}
3975: Thus, if the cavity field is measured and found in the state
3976: $\ket {\alpha}$,
3977: the atoms are in the antisymmetric state. If the cavity field is
3978: found in the state $\ket {-\alpha}$, the atoms are in the entangled
3979: state
3980: \begin{equation}
3981: \ket {\Psi_{12}(-\alpha)} = \frac{1}{2}\left(\ket {g_{1}}\ket {g_{2}}
3982: +\ket {e_{1}}\ket {e_{2}}\right) \ .\label{t190}
3983: \end{equation}
3984: The state~(\ref{t190}) is called as a two photon entangled state. In
3985: section~\ref{ftsec10}, we will discuss another method of preparing
3986: the system in the two-photon entangled state based on the interaction
3987: of two atoms with a squeezed vacuum field.
3988:
3989:
3990:
3991: \subsection{Entanglement of two distant atoms}\label{ftsec83}
3992:
3993:
3994: In the previous subsection, we have discussed different excitation
3995: processes which can prepare two atoms in the antisymmetric state.
3996: The analysis involved single mode cavities, but ignored spontaneous
3997: emission from the atoms and the cavity damping. Here, we will extend
3998: this analysis to include spontaneous emission from the atoms and the
3999: cavity damping~\cite{wan01}. We will show that two atoms separated by
4000: an arbitrary distance $r_{12}$ and interacting with a strongly damped
4001: cavity mode can behave as the Dicke model even if there is no assumed
4002: interaction between the atoms.
4003:
4004: Consider two identical atoms separated by a large distance such that
4005: $\Gamma_{12}\approx 0$ and $\Omega_{12}\approx 0$. The interatomic
4006: axis is oriented perpendicular to the direction of the cavity mode
4007: (cavity axis) which is driven by an external coherent laser field of the
4008: Rabi frequency $\Omega$. The atoms are coupled to the cavity mode with
4009: coupling constant $g$, and damped at the rate $\Gamma$ by spontaneous
4010: emission to modes other than the privileged cavity mode. For
4011: simplicity, we assume that the frequencies of the laser field
4012: $\omega_{L}$ and the cavity mode $\omega_{c}$ are both equal to the
4013: atomic transition frequency $\omega_{0}$. The master equation for the
4014: density operator $\hat{\rho}_{ac}$ of the system of two atoms plus
4015: cavity field has the form
4016: \begin{eqnarray}
4017: \frac{\partial \hat{\rho}_{ac} }{\partial t} &=&
4018: -\frac{1}{2}i\Omega \left[\hat{a}+\hat{a}^{\dagger},
4019: \hat{\rho}_{ac}\right]
4020: -\frac{1}{2}ig\left[S^{-}\hat{a}^{\dagger}+\hat{a}S^{+},
4021: \hat{\rho}_{ac}\right] \nonumber \\
4022: && -\frac{1}{2}\Gamma
4023: \hat{L}_{a}\hat{\rho}_{ac}
4024: -\frac{1}{2}\Gamma_{c}\hat{L}_{c}\hat{\rho}_{ac} \ ,\label{t191}
4025: \end{eqnarray}
4026: where
4027: \begin{eqnarray}
4028: \hat{L}_{a}\hat{\rho}_{ac} &=& \sum_{i=1}^{2}\left( \hat{\rho}_{ac}
4029: S_{i}^{+}S_{i}^{-}+S_{i}^{+}S_{i}^{-}\hat{\rho}_{ac}
4030: -2S_{i}^{-}\hat{\rho}_{ac} S_{i}^{+}\right) \ ,\nonumber \\
4031: \hat{L}_{c}\hat{\rho}_{ac} &=&
4032: \hat{a}^{\dagger}\hat{a}\hat{\rho}_{ac}
4033: +\hat{\rho}_{ac}\hat{a}^{\dagger}\hat{a}
4034: -2\hat{a}\hat{\rho}_{ac}\hat{a}^{\dagger} \ ,\label{t192}
4035: \end{eqnarray}
4036: are operators representing the damping of the atoms by spontaneous
4037: emission and of the field by cavity decay, respectively; $\hat{a}$
4038: and $\hat{a}^{\dagger}$ are the cavity-mode annihilation and creation
4039: operators, $S^{\pm}= S^{\pm}_{1}+S^{\pm}_{2}$ are collective atomic
4040: operators, and $\Gamma_{c}$ is the cavity damping rate.
4041:
4042: To explore the dynamics of the atoms, we assume the "bad-cavity''
4043: limit of $\Gamma_{c}\gg g \gg \Gamma$. This enables us to
4044: adiabatically eliminate the cavity mode and obtain a master equation
4045: for the reduced density operator of the atoms. We make the unitary
4046: transformation
4047: \begin{eqnarray}
4048: \hat{\rho}_{T} &=& \hat{D}(-\eta )\hat{\rho}_{ac}\hat{D}(\eta )
4049: \ ,\label{t193}
4050: \end{eqnarray}
4051: where
4052: \begin{eqnarray}
4053: \hat{D}(\eta ) = e^{\eta \left(\hat{a}+\hat{a}^{\dagger}\right)}
4054: \label{t194}
4055: \end{eqnarray}
4056: is the displacement operator, and $\eta =\Omega/\Gamma_{c}$.
4057:
4058: The master equation for the transformed operator reduces to
4059: \begin{eqnarray}
4060: \frac{\partial \hat{\rho}_{T} }{\partial t} &=&
4061: \frac{1}{2}ig\eta \left[S^{+}+S^{-},
4062: \hat{\rho}_{T}\right]
4063: -\frac{1}{2}ig\left[S^{-}\hat{a}^{\dagger}+\hat{a}S^{+},
4064: \hat{\rho}_{T}\right] \nonumber \\
4065: && -\frac{1}{2}\Gamma
4066: \hat{L}_{a}\hat{\rho}_{T}
4067: -\frac{1}{2}\Gamma_{c}\hat{L}_{c}\hat{\rho}_{T} \ .\label{t195}
4068: \end{eqnarray}
4069: We now introduce the photon number representation for the density
4070: operator $\hat{\rho}_{T}$ with respect to the cavity mode
4071: \begin{eqnarray}
4072: \hat{\rho}_{T} = \sum_{m,n=0}^{\infty} \rho_{mn}\ket m\bra n
4073: \ ,\label{t196}
4074: \end{eqnarray}
4075: where $\rho_{mn}$ are the density matrix elements in the basis of the
4076: photon number states of the cavity mode. Since the cavity mode is
4077: strongly damped, we can neglect populations of the highly excited
4078: cavity modes and limit the expansion to $m,n=1$. Under this
4079: approximation, the master equation~(\ref{t195}) leads to the
4080: following set of coupled equations of motion for the density matrix
4081: elements
4082: \begin{eqnarray}
4083: \dot{\rho}_{00} &=& \hat{L}\rho_{00}
4084: -\frac{1}{2}ig\left(S^{+}\rho_{10}-\rho_{01}S^{-}\right)
4085: +\Gamma_{c}\rho_{11} \ ,\nonumber \\
4086: \dot{\rho}_{10} &=& \hat{L}\rho_{10}
4087: -\frac{1}{2}ig\left(S^{-}\rho_{00}-\rho_{11}S^{-}\right)
4088: -\frac{1}{2}\Gamma_{c}\rho_{10} \ ,\nonumber \\
4089: \dot{\rho}_{11} &=& \hat{L}\rho_{11}
4090: -\frac{1}{2}ig\left(S^{-}\rho_{01}-\rho_{10}S^{+}\right)
4091: -\Gamma_{c}\rho_{11} \ ,\label{t197}
4092: \end{eqnarray}
4093: where $\hat{L}\rho_{ij}=\frac{1}{2}ig\eta \left[S^{+}+S^{-},
4094: \rho_{ij}\right] -\frac{1}{2}\Gamma \hat{L}_{a}\rho_{ij}$.
4095:
4096: We note that the field-matrix elements $\rho_{mn}$ are still operators
4097: with respect to the atoms. Moreover
4098: \begin{eqnarray}
4099: \rho_{00}+\rho_{11} ={\rm Tr}_{F}\left(\hat{\rho}_{T}\right)
4100: = \hat{\rho} \label{t198}
4101: \end{eqnarray}
4102: is the reduced density operator of the atoms.
4103:
4104: For the case of a strong cavity damping the most populated state of
4105: the cavity field is the ground state $\ket 0$, and then we can assume
4106: that the coherence $\rho_{10}$ changes slowly in time, so that we can
4107: take $\dot{\rho}_{10}=0$. Hence, we find that
4108: \begin{eqnarray}
4109: \rho_{10}\approx -\frac{ig}{\Gamma_{c}}\left(S^{-}\rho_{00}
4110: -\rho_{11}S^{-}\right) \ .\label{t199}
4111: \end{eqnarray}
4112: Substituting this solution to $\dot{\rho}_{00}$ and
4113: $\dot{\rho}_{11}$, we get
4114: \begin{eqnarray}
4115: \dot{\rho}_{00} &=& \hat{L}\rho_{00}
4116: +\Gamma_{c}\rho_{11}
4117: -\frac{g^{2}}{2\Gamma_{c}}\left(S^{+}S^{-}\rho_{00}
4118: +\rho_{00}S^{+}S^{-} -2S^{+}\rho_{11}S^{-}\right)
4119: \ ,\nonumber \\
4120: \dot{\rho}_{11} &=& \hat{L}\rho_{11}
4121: -\Gamma_{c}\rho_{11}
4122: +\frac{g^{2}}{2\Gamma_{c}}\left(2S^{-}\rho_{00}S^{+}
4123: -S^{-}S^{+}\rho_{11} -\rho_{11}S^{-}S^{+}\right)
4124: \ .\label{t200}
4125: \end{eqnarray}
4126: Adding these two equations together and neglecting population of the
4127: state $\ket 1$, we obtain the master equation for the reduced density
4128: operator of the atoms as
4129: \begin{eqnarray}
4130: \frac{\partial \hat{\rho} }{\partial t} &=&
4131: \frac{1}{2}ig\eta \left[S^{+}+S^{-},
4132: \hat{\rho}\right]
4133: -\frac{1}{2}\Gamma
4134: \hat{L}_{a}\hat{\rho} \nonumber \\
4135: &&-\frac{g^{2}}{2\Gamma_{c}}\left(S^{+}S^{-}\hat{\rho}
4136: +\hat{\rho}S^{+}S^{-} -2S^{-}\hat{\rho}S^{+}\right)
4137: \ .\label{t201}
4138: \end{eqnarray}
4139: The first term in Eq.~(\ref{t201}) describes the interaction of the
4140: atoms with the driving field of an effective Rabi frequency
4141: $g\eta$. The second term represents the usual damping of
4142: the atoms by spontaneous emission, whereas the last term describes the
4143: damping of the collective system with an effective damping rate
4144: $g^{2}/\Gamma_{c}$. If we choose the parameters such that the collective
4145: damping is much larger than the spontaneous rates of the single atoms,
4146: the second term in Eq.~(\ref{t201}) can be ignored, and then the
4147: master equation~(\ref{t201}) describes the time evolution of the
4148: collective two-atom system. Thus, two independent atoms located
4149: inside a strongly damped one-mode cavity behave as a single collective
4150: small sample model (Dicke model) with the damping rate
4151: $g^{2}/\Gamma_{c}$. This model, however, requires that the atoms are
4152: strongly coupled to the cavity mode $(g\gg \Gamma)$ and are
4153: located inside the cavity such that the interatomic axis is
4154: perpendicular to the direction of the cavity mode and the driving
4155: field.
4156:
4157:
4158:
4159: \subsection{Preparation of a superposition of the antisymmetric and the
4160: ground states}\label{ftsec84}
4161:
4162:
4163:
4164: In the section~\ref{ftsec822}, we have shown that two nonidentical
4165: two-level atoms can be
4166: prepared in an arbitrary superposition of the maximally entangled
4167: antisymmetric state $\ket a$ and the ground state $\ket g$
4168: \begin{equation}
4169: \ket \Phi = \gamma \ket a +\sqrt{1-|\gamma|^{2}}\ket g \ .\label{t202}
4170: \end{equation}
4171: However, the preparation of the superposition state requires that the
4172: atoms have different transition frequencies. Recently, Beige {\it et
4173: al.}~\cite{bmk} have proposed a scheme in which the superposition
4174: state $\ket \Phi$ can be prepared in a system of two identical atoms
4175: placed at fixed positions inside an optical cavity.
4176:
4177: Here, we discuss an alternative scheme where the superposition state
4178: $\ket \Phi$ can be generated in two identical atoms driven in free space
4179: by a coherent laser field. This can happen when the atoms are in
4180: nonequivalent positions in the driving field, i.e. the atoms
4181: experience different intensities and phases of the driving field.
4182: For a comparison, we first consider a specific geometry for the driving
4183: field, namely that
4184: the field is propagated perpendicularly to the atomic axis
4185: $(\vec{k}_{L}\cdot \vec{r}_{12}=0)$. We find from Eq.~(\ref{t150}), that
4186: in this case the collective states are populated with the
4187: population distribution $\rho_{ee}=\rho_{aa}<\rho_{ss}$.
4188: The population distribution
4189: changes dramatically when the driving field propagates in directions
4190: different from perpendicular to the interatomic axis~\cite{hsf82,fs,rfd}.
4191: In this situation the populations strongly depend on the interatomic
4192: separation and the detuning $\Delta_{L}$. This can produce the interesting
4193: modification that the collective states can be selectively populated.
4194: We show this by solving numerically the system of~15 equations for
4195: the density matrix elements. The populations are plotted against the
4196: detuning $\Delta_{L}$ in Fig.~\ref{ftfig16} for the laser field
4197: propagating in the direction of the interatomic axis. We see from
4198: Fig.~\ref{ftfig16} that the collective states $\ket s$ and $\ket e$ are
4199: populated at $\Delta_{L}=0$ and $\Delta_{L} = \Omega_{12}$.
4200: The antisymmetric state is significantly populated at
4201: $\Delta_{L} = -\Omega_{12}$, and at this detuning the populations of the
4202: states $\ket s$ and $\ket e$ are close to zero. Since
4203: $\rho_{aa}<1$, the population is distributed between the antisymmetric
4204: and the ground states, and therefore at $\Delta_{L} =-\Omega_{12}$ the
4205: system is in a superposition of the maximally entangled state $\ket a$ and
4206: the ground state $\ket g$.
4207: %
4208: \begin{figure}[t]
4209: \begin{center}
4210: \includegraphics[width=10cm]{ftfig16.eps}
4211: \end{center}
4212: \caption{The steady-state populations of the collective atomic states
4213: of two identical atoms as a
4214: function of $\Delta_{L}$ for the driving field propagating in the
4215: direction of the interatomic axis, $\Omega =2.5\Gamma$,
4216: $r_{12}/\lambda =0.08$ and $\bar{\mu} \perp
4217: \bar{r}_{12}$: $\rho_{ee}$ (solid line), $\rho_{aa}$ (dashed line),
4218: $\rho_{ss}$ (dashed-dotted line).}
4219: \label{ftfig16}
4220: \end{figure}
4221:
4222: Turchette {\it et al.}~\cite{tur} have recently realized experimentally
4223: a superposition state of the ground state and a non-maximally
4224: entangled antisymmetric state in two trapped ions. In the experiment
4225: two trapped barium ions were sideband cooled to their motional ground
4226: states. Transitions between the states of the ions were induced by
4227: Raman pulses using co-propagating lasers. The ions were at positions
4228: that experience different Rabi frequencies $\Omega_{1}$ and
4229: $\Omega_{2}$ of the laser fields. By preparing the initial motional
4230: ground state with one ion excited $\ket {e_{1}}\ket {g_{2}}\ket 0$,
4231: and applying the laser fields for a time $t$, the following entangled
4232: state $\ket {\Psi \left(t\right)}$ was created
4233: \begin{eqnarray}
4234: \ket {\Psi \left(t\right)} &=& -\frac{i\Omega_{2}}{\Omega}\sin
4235: \left(\Omega t\right)\ket g \ket 1
4236: + \left\{\left[\frac{\Omega_{2}^{2}}{\Omega^{2}}\left(\cos \Omega t
4237: -1\right) +1\right]\ket {e_{1}}\ket {g_{2}}\right. \nonumber \\
4238: && +\left. \left[\frac{\Omega_{1}\Omega_{2}}{\Omega^{2}}\left(\cos \Omega t
4239: -1\right)\right]\ket {g_{1}}\ket {e_{2}}\right\}\ket 0 \ ,\label{t203}
4240: \end{eqnarray}
4241: where $\Omega^{2}=\Omega_{1}^{2}+\Omega_{2}^{2}$.
4242:
4243: For $\Omega t =\pi$ the entangled state~(\ref{t203}) reduces to a
4244: non-maximally entangled antisymmetric state
4245: \begin{equation}
4246: \ket {\Psi_{a}} = \left[\frac{\Omega_{1}^{2}-\Omega_{2}^{2}}{\Omega^{2}}
4247: \ket {e_{1}}\ket {g_{2}} -\frac{2\Omega_{1}\Omega_{2}}{\Omega^{2}}
4248: \ket {g_{1}}\ket {e_{2}}\right]\ket 0 \ .\label{t204}
4249: \end{equation}
4250:
4251: Franke {\it et al.}~\cite{fhb} have proposed to use the non-maximally
4252: entangled state~(\ref{t204}) to demonstrate the intrinsic difference
4253: between quantum and classical information transfers. The difference
4254: arises from the different ways in which the probabilities occur and is
4255: particularly clear in terms of entangled states.
4256:
4257:
4258:
4259: \section{Detection of the entangled states}\label{ftsec9}
4260:
4261:
4262:
4263: In this section we describe two possible methods for detection of
4264: entangled states of two interacting atoms. One is the observation of
4265: angular intensity distribution of the fluorescence field emitted by
4266: the system of two interacting atoms. The other is based on quantum
4267: interference in which one observes interference pattern of the
4268: emitted field. Beige {\it et al.}~\cite{bbtk} have proposed a scheme,
4269: based on the quantum Zeno effect, to observe a decoherence-free state
4270: in a system of two three-level atoms located inside an optical cavity.
4271: The two schemes discussed here involve two two-level atoms in free
4272: space.
4273:
4274:
4275: \subsection{Angular fluorescence distribution}\label{ftsec91}
4276:
4277:
4278: It is well known that the fluorescence field emitted from a two-atom
4279: system exhibits strong directional properties~\cite{ftk87,leh,rfd,du}.
4280: This property can be used to detect an internal state of two interacting
4281: atoms. To show this, we consider the fluorescence intensity, defined
4282: in Eq.~(\ref{t69}), that in terms of the density matrix elements of
4283: the collective atomic system can be written as
4284: \begin{eqnarray}
4285: I\left(\vec{R},t\right) &=& u(\vec{R}) \left\{
4286: \left(\rho_{ee}+\rho_{ss}\right)\left[1 +\cos\left(kr_{12}\cos\theta
4287: \right)\right]\right. \nonumber \\
4288: &&+\left. \left(\rho_{ee}+\rho_{aa}\right)\left[1 -\cos\left(kr_{12}
4289: \cos\theta \right)\right]\right. \nonumber \\
4290: &&+\left. i\left(\rho_{sa}-\rho_{as}\right)\sin\left(kr_{12}\cos\theta
4291: \right)\right\} \ ,\label{t205}
4292: \end{eqnarray}
4293: where $\theta$ is the angle between the observation direction $\vec{R}$
4294: and the vector $\vec{r}_{12}$.
4295:
4296: The first term in Eq.~(\ref{t205}) arises from the fluorescence emitted
4297: on the $\ket e \rightarrow \ket s \rightarrow \ket g$ transitions,
4298: which involve the symmetric state.
4299: The second term arises from the $\ket e \rightarrow \ket a \rightarrow
4300: \ket g$ transitions through the antisymmetric state. These two terms
4301: describe two different channels of transitions for which the angular
4302: distribution is proportional to $\left[1 \pm \cos\left(kr_{12}\cos\theta
4303: \right)\right]$. The last term in Eq.~(\ref{t205}) originates from
4304: interference between these two radiation channels. It is evident from
4305: Eq.~(\ref{t205}) that the angular distribution of the fluorescence field
4306: depends on the population of the entangled states $\ket s$ and $\ket
4307: a$. Moreover, independent of the interatomic separation~$r_{12}$, the
4308: antisymmetric state does not radiate in the direction perpendicular to
4309: the atomic axis, as for $\theta =\pi/2$ the factor
4310: $\left[1 -\cos\left(kr_{12}\cos\theta \right)\right]$ vanishes. In
4311: contrast, the symmetric state radiates in all directions.
4312: Hence, the spatial distribution of the fluorescence field is
4313: not spherical unless $\rho_{ss}=\rho_{aa}$ and then the angular
4314: distribution is spherically symmetric independent of the interatomic
4315: separation. Therefore, an asymmetry in the angular distribution of the
4316: fluorescence field would be a compelling evidence that
4317: the entangled states $\ket s$ and $\ket a$ are not equally populated.
4318: If the fluorescence is detected in the direction perpendicular to the
4319: interatomic axis the observed intensity (if any) would correspond to the
4320: fluorescence field emitted from the symmetrical state $\ket s$ and/or
4321: the upper state $\ket e$. On the other hand, if there is no
4322: fluorescence detected in the direction perpendicular to the atomic
4323: axis, the population is entirely in a superposition of the
4324: antisymmetric state $\ket a$ and the ground state $\ket g$.
4325:
4326: Guo and Yang~\cite{gy1,gy2} have analyzed spontaneous decay from two atoms
4327: initially prepared in an entangled state. They have shown that the
4328: time evolution of the population inversion, which is proportional to
4329: the total radiation intensity~(\ref{t71}), depends on the degree of
4330: entanglement of the initial state of the system. In Sections~\ref{sec411}
4331: and~\ref{sec412}, we have shown that in the case of two non-identical
4332: atoms the time evolution of the total radiation intensity
4333: $I\left(t\right)$ can exhibit quantum beats
4334: which result from the presence of correlations between the symmetric
4335: and antisymmetric states. In fact, quantum beats are present only if
4336: initially the system is in a non-maximally entangled state, and no
4337: quantum beats are predicted for maximally entangled as well as
4338: unentangled states.
4339:
4340:
4341: \subsection{Interference pattern with a dark center}\label{ftsec92}
4342:
4343:
4344: An alternative way to detect entangled states of a two-atom system
4345: is to observe an interference pattern of the fluorescence field
4346: emitted in the direction $\vec{R}$, not necessary perpendicular to
4347: the interatomic axis.
4348:
4349: This scheme is particularly useful for detection of the symmetric or
4350: the antisymmetric state. To show this, we consider the visibility in
4351: terms of the density matrix elements of the collective atomic system
4352: as
4353: \begin{equation}
4354: {\cal {V}} =
4355: \frac{\rho_{ss}-\rho_{aa}}{\rho_{ss}+\rho_{aa}+2\rho_{ee}} \
4356: .\label{t206}
4357: \end{equation}
4358: This simple formula shows that the sign of ${\cal {V}}$ depends on the
4359: population difference between the symmetric and antisymmetric states.
4360: For $\rho_{ss}>\rho_{aa}$ the visibility ${\cal {V}}$ is positive, and
4361: then the interference pattern exhibits a maximum (bright center),
4362: whereas for $\rho_{ss}<\rho_{aa}$ the visibility ${\cal {V}}$ is
4363: negative and then there is a minimum (dark center). The optimum
4364: positive (negative) value is ${\cal {V}} = 1$ $({\cal {V}} = -1)$, and
4365: there is no interference pattern when ${\cal{V}}=0$. The later happens
4366: when $\rho_{ss}=\rho_{aa}$.
4367:
4368: Similar to the fluorescence intensity distribution, the visibility can
4369: provide an information about the entangled states of a two-atom
4370: system. When the system is prepared in the antisymmetric state or in a
4371: superposition of the antisymmetric and the ground states,
4372: $\rho_{ss}=\rho_{ee}=0$, and then the visibility has the optimum
4373: negative value ${\cal {V}} = -1$. On the other hand, when the system
4374: is prepared in the symmetric state or in a linear superposition of the
4375: symmetric and ground states, the visibility has the maximum positive
4376: value ${\cal {V}} = 1$.
4377: %
4378: \begin{figure}[t]
4379: \begin{center}
4380: \includegraphics[width=10cm]{ftfig17.eps}
4381: \end{center}
4382: \caption{The visibility $V$ as a function of $\Delta_{L}$ for
4383: $r_{12}=0.1\lambda, \Omega =0.5\Gamma$ and various angles
4384: $\theta_{L}$; $\theta_{L}=\pi/2$ (solid line), $\theta_{L}=\pi/4$
4385: (dashed line), $\theta_{L}=0$ (dashed-dotted line).}
4386: \label{ftfig17}
4387: \end{figure}
4388:
4389: The earliest theoretical studies of the fringe visibility involved
4390: a coherent driving field which produces an interference pattern with a
4391: bright center.
4392: Recently, Meyer and Yeoman \cite{meyer} have shown that in contrast to the
4393: coherent excitation, the incoherent field produces an interference pattern
4394: with a dark center. Dung and Ujihara~\cite{du} have shown that the
4395: fringe contrast factor can be negative for spontaneous
4396: emission from two undriven atoms, with initially one atom excited.
4397: Interference pattern with a dark center can also be obtained with a
4398: coherent driving field~\cite{fr}. This happens when the atoms experience
4399: different phases and/or intensities of the driving field. To show this,
4400: we solve numerically the master equation~(\ref{t42}) for the
4401: steady-state density matrix elements of the driven system of two atoms.
4402: The visibility $V$ is plotted in Fig.~\ref{ftfig17} as a function of
4403: the detuning $\Delta_{L}$ for $r_{12}=0.1\lambda,
4404: \Omega =0.25\Gamma$ and various angles $\theta_{L}$ between the
4405: interatomic axis and the direction of propagation of the laser field.
4406: The visibility $V$ is positive for most values of $\Delta_{L}$, except
4407: $\Delta_{L}\approx -\Omega_{12}$. At this detuning the parameter $V$
4408: is negative and reaches the optimum negative value $V=-1$ indicating
4409: that the system produces interference pattern with a dark center. In
4410: Fig.~\ref{ftfig18}, we plot the populations of the symmetric and
4411: antisymmetric states for the same parameters as in Fig~\ref{ftfig17}.
4412: It is evident from Fig.~\ref{ftfig18}
4413: that at $\Delta_{L}=-\Omega_{12}$ the antisymmetric state is
4414: significantly populated, whereas the population of the symmetric
4415: state is close to zero. This population difference leads to negative
4416: values of $V$, as predicted by Eq.~(\ref{t206}) and seen in
4417: Fig.~\ref{ftfig17}. Experimental observation of the interference pattern
4418: with a dark center would be an interesting demonstration of the
4419: controled excitation of a two-atom system to the entangled antisymmetric
4420: state.
4421: %
4422: \begin{figure}[t]
4423: \begin{center}
4424: \includegraphics[width=10cm]{ftfig18.eps}
4425: \end{center}
4426: \caption{Populations of the symmetric and antisymmetric states for
4427: the same parameters as in Fig.~\ref{ftfig17}, with $\theta_{L}=0$. }
4428: \label{ftfig18}
4429: \end{figure}
4430:
4431:
4432:
4433: \section{Two-photon entangled states}\label{ftsec10}
4434:
4435:
4436:
4437: In our discussions to date on entanglement creation in two-atom
4438: systems, we have focused on different methods of creating entangled
4439: states of the form
4440: \begin{eqnarray}
4441: \ket \Psi = c_{1}\ket {e_{1}}\ket {g_{2}} \pm c_{2}\ket {g_{1}}
4442: \ket {e_{2}} \ .\label{t207}
4443: \end{eqnarray}
4444: As we have shown in Sec.~\ref{ftsec31}, entangled states of the above
4445: form are generated by the dipole-dipole interaction
4446: between the atoms and the preparation of these states is sensitive to
4447: the difference $\Delta$ between the atomic transition frequencies
4448: and to the atomic decay rates. These states are better known as the
4449: symmetric and antisymmetric collective atomic states.
4450:
4451: Apart from the symmetric and antisymmetric states,
4452: there are two other collective states of the two-atom system: the
4453: ground state $\ket g =\ket {g_{1}}\ket {g_{2}}$ and the upper state
4454: $\ket e =\ket {e_{1}}\ket {e_{2}}$, which are also product states of
4455: the individual atomic states. These states are
4456: not affected by the dipole-dipole interaction $\Omega_{12}$.
4457:
4458: In this section, we discuss a method of creating entanglement between
4459: these two states of the general form
4460: \begin{equation}
4461: \ket {\Upsilon} = c_{g}\ket g \pm c_{e}\ket e \ ,\label{t208}
4462: \end{equation}
4463: where $c_{g}$ and $c_{e}$ are transformation parameters such that
4464: $|c_{g}|^{2}+|c_{e}|^{2}=1$. The entangled states of the form
4465: (\ref{t208}) are known in literature as pairwise atomic
4466: states~\cite{pas1,pas2,pas3,pas4} or multi-atom squeezed
4467: states~\cite{mas}. According to Eq.~(\ref{t63}), the collective
4468: ground and excited states are separated in energy by $2\hbar \omega_{0}$,
4469: and therefore we can call the states $\ket \Upsilon$ as two-photon
4470: entangled (TPE) states.
4471:
4472: The two-photon entangled states cannot be generated by a simple coherent
4473: excitation. A coherent field applied to the two-atom system couples
4474: to one-photon transitions. The problem is that coherent excitation
4475: populates not only the upper state $\ket e$ but also the
4476: intermediate states $\ket s$ and $\ket a$, see Eq.~(\ref{t150}).
4477: The two-photon entangled states (\ref{t208}) are superpositions of
4478: the collective ground and excited states with no contribution from the
4479: intermediate collective states $\ket s$ and $\ket a$.
4480:
4481: The two-photon behavior of the entangled states (\ref{t208}) suggests
4482: that the simplest technique for generating the TPE
4483: states would be by applying a two-photon excitation process. An
4484: obvious candidate is a squeezed vacuum field which is characterized
4485: by strong two-photon correlations which would enable the transition
4486: $\ket g \rightarrow \ket e$ to occur effectively in a single step
4487: without populating the intermediate states. We will illustrate this
4488: effect by analyzing the populations of the collective atomic states of
4489: a two-atom system interacting with a squeezed vacuum field.
4490:
4491:
4492: \subsection{Populations of the entangled states in a squeezed vacuum}
4493: \label{ftsec101}
4494:
4495:
4496: The general master equation~(\ref{t36}) allows us to calculate the
4497: populations of the collective atomic states and coherences, which
4498: gives information about the stationary state of a two-atom system.
4499: We first consider a system of two identical atoms, separated by an
4500: arbitrary distance $r_{12}$ and interacting with a squeezed vacuum field.
4501: For simplicity, we assume that the carrier frequency $\omega_{s}$ of the
4502: squeezed vacuum field is resonant to the atomic transition frequency
4503: $\omega_{0}$, and the squeezed field is perfectly matched to the
4504: atoms, $D(\omega_{s})=1$ and $\theta_{s}=\pi$.
4505:
4506: From the master equation~(\ref{t36}), we find the following equations
4507: of motion for the populations of the collective states and the
4508: two-photon coherences of the collective system of two identical atoms
4509: \begin{eqnarray}
4510: \dot{\rho}_{ee} &=&
4511: -2\Gamma\left(\tilde{N}+1\right)\rho_{ee}
4512: +\tilde{N}\left[\left(\Gamma
4513: +\Gamma_{12}\right)\rho_{ss} +\left(\Gamma
4514: -\Gamma_{12}\right)\rho_{aa}\right]
4515: +\Gamma_{12}|\tilde{M}|\rho_{u} \ ,\nonumber \\
4516: \dot{\rho}_{ss} &=&
4517: \left(\Gamma +\Gamma_{12}\right)\left\{\tilde{N}
4518: -\left(3\tilde{N}+1\right)\rho_{ss} -\tilde{N}\rho_{aa}
4519: +\rho_{ee} -|\tilde{M}|\rho_{u}\right\} \ ,\nonumber \\
4520: \dot{\rho}_{aa} &=&
4521: \left(\Gamma -\Gamma_{12}\right)\left\{\tilde{N}
4522: -\left(3\tilde{N}+1\right)\rho_{aa} -\tilde{N}\rho_{ss}
4523: +\rho_{ee}
4524: +|\tilde{M}|\rho_{u}\right\} \ ,\nonumber \\
4525: \dot{\rho}_{u} &=& 2\Gamma_{12}|\tilde{M}|
4526: -\left(2\tilde{N}+1\right)\Gamma \rho_{u} \nonumber \\
4527: && -2|\tilde{M}|\left[\left(\Gamma
4528: +2\Gamma_{12}\right)\rho_{ss} -\left(\Gamma
4529: -2\Gamma_{12}\right)\rho_{aa}\right] \ ,\label{t209}
4530: \end{eqnarray}
4531: where $\tilde{N}=\tilde{N}\left(\omega_{0}\right)$, $\tilde{M}
4532: =\tilde{M}\left(\omega_{0}\right)$ and
4533: $\rho_{u}=\rho_{eg}\exp (-i\phi_{s}) +\rho_{ge}\exp (i\phi_{s})$.
4534:
4535: It is seen from Eq.~(\ref{t209}) that the evolution of the populations
4536: depends on the two-photon coherencies $\rho_{eg}$ and $\rho_{ge}$,
4537: which can transfer the population from the ground state $\ket g$
4538: directly to the upper state $\ket e$ leaving the states $\ket s$
4539: and $\ket a$ unpopulated. The evolution of the populations depends
4540: on $\Gamma_{12}$, but is completely independent of the dipole-dipole
4541: interaction $\Omega_{12}$.
4542:
4543: Similar to the interaction with the ordinary vacuum, discussed in
4544: Sec.~\ref{ftsec31}, the steady-state solution of Eqs.~(\ref{t209})
4545: depends on whether $\Gamma_{12}=\Gamma$ or $\Gamma_{12}\neq \Gamma$.
4546: Assuming that $\Gamma_{12}=\Gamma$ and setting the left-hand side of
4547: equations~(\ref{t209}) equal to zero, we obtain the steady-state
4548: solutions for the populations and the two-photon coherence in the
4549: Dicke model. A straightforward algebraic manipulation of
4550: Eqs.~(\ref{t209}) leads to the following steady-state solutions
4551: \begin{eqnarray}
4552: \rho_{ee} &=& \frac{\tilde{N}^{2}\left(2\tilde{N}+1\right)
4553: -\left(2\tilde{N}-1\right)|\tilde{M}|^{2}}{\left(2\tilde{N}+1\right)
4554: \left(3\tilde{N}^{2} +3\tilde{N} +1
4555: -3|\tilde{M}|^{2}\right)} \ ,\nonumber \\
4556: \rho_{ss} &=& \frac{\tilde{N}\left(\tilde{N}+1\right)
4557: -|\tilde{M}|^{2}}{3\tilde{N}^{2} +3\tilde{N} +1
4558: -3|\tilde{M}|^{2}} \ ,\nonumber \\
4559: \rho_{u} &=& \frac{2|\tilde{M}|}{\left(2\tilde{N}+1\right)
4560: \left(3\tilde{N}^{2} +3\tilde{N} +1
4561: -3|\tilde{M}|^{2}\right)} \ .\label{t210}
4562: \end{eqnarray}
4563:
4564: The steady-state populations depend strongly on the squeezing correlations
4565: $\tilde{M}$. For a classical squeezed
4566: field with the maximal correlations $\tilde{M}=\tilde{N}$ the steady-state
4567: populations reduces to
4568: \begin{eqnarray}
4569: \rho_{ss} &=& \frac{\tilde{N}}{3 \tilde{N} +1} \ ,\nonumber \\
4570: \rho_{ee} &=& \frac{2\tilde{N}^{2}}{\left( 2\tilde{N}+1\right)
4571: \left(3 \tilde{N} +1\right)} \ .\label{t211}
4572: \end{eqnarray}
4573: It is easily to check that in this case the populations obey a Boltzmann
4574: distribution with $\rho_{gg}>\rho_{ss}>\rho_{ee}$. Moreover, in the
4575: limit of low intensities $(\tilde{N}\ll 1)$ of the field, the
4576: population $\rho_{ee}$ is proportional to $\tilde{N}^{2}$, showing
4577: that in classical fields the population exhibits a quadratic dependence
4578: on the intensity.
4579:
4580: The population distribution is qualitatively different for a quantum
4581: squeezed field with the maximal correlations $|\tilde{M}|^{2}
4582: =\tilde{N}(\tilde{N}+1)$. In this case, the stationary populations of
4583: the excited collective states are
4584: \begin{eqnarray}
4585: \rho_{ss} &=& 0 \ , \nonumber \\
4586: \rho_{ee} &=& \frac{ \tilde{N}}{\left(2 \tilde{N}+1\right)}
4587: \ .\label{t212}
4588: \end{eqnarray}
4589: Clearly, the symmetric state is not populated. In this case the
4590: populations no longer obey the Boltzmann distribution. The
4591: population is distributed only between the ground state $\ket g$ and
4592: the upper state~$\ket e$. Moreover, it can be seen from Eq.~(\ref{t212})
4593: that for a weak quantum squeezed field the population $\rho_{ee}$ depends
4594: linearly on the intensity. This distinctive feature reflects the
4595: direct modifications of the two-photon absorption that the
4596: nonclassical photon correlations enable the transition
4597: $\ket g \rightarrow \ket e$ to occurs in a "single step" proportional
4598: to $\tilde{N}$. In other words, the nonclassical two-photon correlations
4599: entangle the ground state $\ket g$ and the upper state $\ket e$ with
4600: no contribution from the symmetric state $\ket s$.
4601:
4602: The question we are interested in concerns the final state of the
4603: system and its purity. To answer this question, we apply
4604: Eq.~(\ref{t210}) and find that in the steady-state, the density
4605: matrix of the system is given by
4606: \begin{eqnarray}
4607: \hat{\rho} &=& \left(
4608: \begin{array}{ccc}
4609: \rho_{gg} & 0 & \rho_{ge} \\
4610: 0 & \rho_{ss} & 0 \\
4611: \rho_{eg} & 0 & \rho_{ee}
4612: \end{array}
4613: \right) \ , \label{t213}
4614: \end{eqnarray}
4615: where $\rho_{ij}$ are the non-zero steady-state density matrix
4616: elements.
4617:
4618: It is evident from Eq.~(\ref{t213}) that in the squeezed vacuum the
4619: density matrix of the system is not diagonal due to the presence of
4620: the two-photon coherencies $\rho_{ge}$ and $\rho_{eg}$. This
4621: indicates that
4622: the collective states $\ket g$, $\ket s$ and $\ket e$ are no longer
4623: eigenstates of the system. The density matrix can be rediagonalized
4624: by including $\rho_{eg}$ and $\rho_{ge}$ to give the new (entangled)
4625: states
4626: \begin{eqnarray}
4627: \ket {\Upsilon_{1}} &=& \left[\left(P_{1}-\rho_{ee}\right)\ket g +
4628: \rho_{eg}\ket e
4629: \right]/\left[\left(P_{1}-\rho_{ee}\right)^{2}
4630: +\left|\rho_{eg}\right|^{2}\right]^{\frac{1}{2}} \ ,\nonumber \\
4631: \ket {\Upsilon_{2}} &=& \left[\rho_{ge}\ket g +
4632: \left(P_{2}-\rho_{gg}\right)\ket e
4633: \right]/\left[\left(P_{2}-\rho_{gg}\right)^{2}
4634: +\left|\rho_{eg}\right|^{2}\right]^{\frac{1}{2}} \ ,\nonumber \\
4635: \ket {\Upsilon_{3}} &=& \ket s \ ,\label{t214}
4636: \end{eqnarray}
4637: where the diagonal probabilities are
4638: \begin{eqnarray}
4639: P_{1} &=& \frac{1}{2}\left(\rho_{gg}+\rho_{ee}\right)
4640: +\frac{1}{2}\left[\left(\rho_{gg}-\rho_{ee}\right)^{2}
4641: +4\left|\rho_{eg}\right|^{2}\right]^{\frac{1}{2}} \ ,\nonumber \\
4642: P_{2} &=& \frac{1}{2}\left(\rho_{gg}+\rho_{ee}\right)
4643: -\frac{1}{2}\left[\left(\rho_{gg}-\rho_{ee}\right)^{2}
4644: +4\left|\rho_{eg}\right|^{2}\right]^{\frac{1}{2}} \ ,\nonumber \\
4645: P_{3} &=& \rho_{ss} \ .\label{t215}
4646: \end{eqnarray}
4647: In view of Eqs.~(\ref{t212}) and~(\ref{t214}), it is easy to see
4648: that the squeezed vacuum causes the system to decay into entangled
4649: states which are linear superpositions of the
4650: collective ground state $\ket g$ and the upper state $\ket e$.
4651: The intermediate symmetric state remains unchanged under the squeezed
4652: vacuum excitation.
4653: In general, the states~(\ref{t214}) are mixed states. However, for
4654: perfect correlations $|\tilde{M}|^{2}=\tilde{N}(\tilde{N}+1)$ the
4655: populations $P_{2}$ and $P_{3}$ are zero
4656: leaving the population only in the state $\ket {\Upsilon_{1}}$. Hence,
4657: the state $\ket {\Upsilon_{1}}$
4658: is a pure state of the system of two atoms driven by a squeezed vacuum
4659: field. From Eqs.~(\ref{t214}), we find that the pure entangled state
4660: $\ket {\Upsilon_{1}}$ is given by~\cite{pk90}
4661: \begin{equation}
4662: \ket {\Upsilon_{1}} = \frac{1}{\sqrt{2\tilde{N}+1}}
4663: \left[\sqrt{\tilde{N}+1}\ket g
4664: +\sqrt{\tilde{N}}\ket e\right] \ .\label{t216}
4665: \end{equation}
4666: The pure state~(\ref{t216}) is non-maximally entangled state, it
4667: reduces to a maximally entangled state for $\tilde{N}\gg 1$. The entangled
4668: state is analogous to the pairwise atomic
4669: state~\cite{pas1,pas2,pas3,pas4} or the multi-atom squeezed
4670: state~\cite{mas}, (see also Ref.~\cite{park}),
4671: predicted in the small sample model of two coupled atoms.
4672:
4673: The pure entangled state $\ket {\Upsilon_{1}}$ is characteristic not
4674: only of the two-atom Dicke model, but in general of the Dicke model of
4675: an even number of atoms~\cite{ap90}. The $N$-atom Dicke system
4676: interacting with a squeezed vacuum can decay to a state which the
4677: density operator is given by
4678: \begin{eqnarray}
4679: \hat{\rho} &=& C_{n}\left(\mu S^{-}+\nu S^{+}\right)^{-1}
4680: \left(\mu S^{+}+\nu S^{-}\right)^{-1} \ ,\label{t217}
4681: \end{eqnarray}
4682: if $N$ is odd, or
4683: \begin{eqnarray}
4684: \hat{\rho} =\ket \Upsilon \bra \Upsilon \ ,\label{t218}
4685: \end{eqnarray}
4686: if $N$ is even, where $C_{n}$ is the normalization constant,
4687: $S^{\pm}$ are the collective atomic operators, $\mu^{2} =\nu^{2}+1
4688: =\tilde{N}+1$, and $\ket \Upsilon$ is defined by
4689: \begin{eqnarray}
4690: \left(\mu S^{-}+\nu S^{+}\right)\ket \Upsilon =0 \ .\label{t219}
4691: \end{eqnarray}
4692: Thus, for an even number of atoms the stationary state of the system
4693: is the pure pairwise atomic state.
4694:
4695:
4696: \subsection{Effect of the antisymmetric state on the purity of the
4697: system}\label{ftsec102}
4698:
4699:
4700: The pure entangled state $\ket {\Upsilon_{1}}$ can be obtained for
4701: perfect matching of the squeezed modes to the atoms and interatomic
4702: separations much smaller than the optical
4703: wavelength. To achieve perfect matching, it is necessary
4704: to squeeze of all the modes to which the atoms are coupled. That is,
4705: the squeezed modes must occupy the whole $4\pi$ solid angle of the
4706: space surrounding the atoms. This is not possible to achieve with the
4707: present experiments in free space, and in order to avoid the difficulty
4708: cavity environments have been suggested~\cite{kimb1,kimb2}. Inside
4709: a cavity the atoms interact strongly only with the privileged cavity
4710: modes. By the squeezing of these cavity modes, which occupy only a
4711: small solid angle about the cavity axis, it would be possible to achieve
4712: perfect matching of the squeezed field to the atoms.
4713:
4714: There is, however, the practical problem to fulfil the second requirement
4715: that interatomic separations should be much smaller than the
4716: resonant wavelength. This assumption may prove difficult in
4717: experimental realization as with the present techniques two atoms can
4718: be trapped within distances of the order of a resonant
4719: wavelength~\cite{eich,deb,ber,tos}. As we have shown in
4720: Sec.~\ref{ftsec31}, the dynamics of such a system involve the
4721: antisymmetric state and are significantly different from the
4722: dynamics of the Dicke model.
4723:
4724: For two atoms separated by an arbitrary distance $r_{12}$, the
4725: collective damping $\Gamma_{12}\neq \Gamma$, and then the steady-state
4726: solutions of Eqs.~(\ref{t209}) are
4727: \begin{eqnarray}
4728: \rho_{ee} &=&\frac{\tilde{N}^{2}}{\left(2\tilde{N}+1\right)^{2}}
4729: +\frac{a^{2}|\tilde{M}|^{2}\left(
4730: 4\tilde{N}+1\right) }{G} \ , \nonumber \\
4731: \rho_{ss} &=&\frac{\tilde{N}\left(\tilde{N}+1\right) }
4732: {\left(2\tilde{N}+1\right)^{2}}-\frac{a|\tilde{M}|^{2}
4733: \left[2\left(2\tilde{N}+1\right)^{2}-a\right] }{G} \ , \nonumber \\
4734: \rho_{aa} &=& \frac{\tilde{N}\left(\tilde{N}+1\right) }
4735: {\left(2\tilde{N}+1\right)^{2}}
4736: +\frac{a|\tilde{M}|^{2}
4737: \left[2\left(2\tilde{N}+1\right)^{2}+a\right] }{G}, \nonumber \\
4738: \rho_{u} &=& \frac{2a\left(2\tilde{N}+1\right)^{3}|\tilde{M}|}{G}
4739: \ ,\label{t220}
4740: \end{eqnarray}
4741: where $a=\Gamma_{12}/\Gamma$, and
4742: \begin{equation}
4743: G = \left(2\tilde{N}+1\right)^{2}\left\{ \left(2\tilde{N}+1\right)^{4}
4744: +4|\tilde{M}|^{2}\left[ a^{2}-\left(2\tilde{N}+1\right)^{2}\right]
4745: \right\} \ .\label{t221}
4746: \end{equation}
4747: %
4748: \begin{figure}[t]
4749: \begin{center}
4750: \includegraphics[width=13cm]{ftfig19.eps}
4751: \end{center}
4752: \caption{The steady-state populations of the collective atomic states
4753: as a function of $r_{12}$ for (a) quantum squeezed field with
4754: $|\tilde{M}|^{2}=\tilde{N}(\tilde{N}+1)$, (b) classical
4755: squeezed field with $|\tilde{M}|=\tilde{N}$, and $\tilde{N}=0.05$,
4756: $\bar{\mu} \perp \bar{r}_{12}$: $\rho_{ee}$ (solid line),
4757: $\rho_{aa}$ (dashed line), $\rho_{ss}$ (dashed-dotted line).}
4758: \label{ftfig19}
4759: \end{figure}
4760:
4761: This result shows that the antisymmetric state is populated in the
4762: steady-state even for small interatomic separations $(\Gamma_{12}\approx
4763: \Gamma)$. For large interatomic separations $\Gamma_{12}\approx 0$, and
4764: then the symmetric
4765: and antisymmetric states are equally populated. When the interatomic
4766: separation decreases, the population of the state $\ket a$ increases,
4767: whereas the population of the state $\ket s$ decreases and
4768: $\rho_{ss}=0$ for very small interatomic separations. These features
4769: are illustrated in Fig.~\ref{ftfig19}(a), where we plot the steady-state
4770: populations as a function of the interatomic separation for the
4771: maximally correlated quantum squeezed field. We see that the collective
4772: states are unequally populated and in the case of small $r_{12}$, the
4773: state $\ket a$ is the most populated state of the system, whereas the
4774: state $\ket s$ is not populated. In Fig.~\ref{ftfig19}(b), we show the
4775: populations for the equivalent maximally correlated classical
4776: squeezed field, and in this case all states are populated independent
4777: of $r_{12}$.
4778:
4779: This fact can lead to a destruction of the purity of the stationary
4780: state of the system. To show this, we calculate the quantity
4781: \begin{equation}
4782: {\rm Tr}\left(\hat{\rho}^{2}\right) =\rho_{gg}^{2}+\rho_{ss}^{2}
4783: +\rho_{aa}^{2}+\rho_{ee}^{2}+\left|\rho_{u}\right|^{2} \ ,
4784: \label{t222}
4785: \end{equation}
4786: which determines the purity of the system. Tr$(\hat{\rho}^{2})=1$
4787: corresponds to a pure state of the system, while Tr$(\hat{\rho}^{2})<1$
4788: corresponds to a mixed state. Tr$(\hat{\rho}^{2})=1/4$ describes a
4789: completely mixed state of the system. In Fig.~\ref{ftfig20}, we display
4790: Tr$(\hat{\rho}^{2})$ as a function of the interatomic separation $r_{12}$
4791: for perfect correlations $|\tilde{M}|^{2}=\tilde{N}(\tilde{N}+1)$
4792: and various $\tilde{N}$. Clearly, the system is in a mixed state
4793: independent of the interatomic separation. Moreover, the purity decreases
4794: as $\tilde{N}$ increases.
4795: %
4796: \begin{figure}[t]
4797: \begin{center}
4798: \includegraphics[width=10cm]{ftfig20.eps}
4799: \end{center}
4800: \caption{Tr$(\hat{\rho}^{2})$ as a function of the interatomic separation
4801: for $|\tilde{M}|^{2}=\tilde{N}(\tilde{N}+1)$,
4802: $\bar{\mu} \perp \bar{r}_{12}$ and different $\tilde{N}$: $\tilde{N}=0.05$
4803: (solid line), $\tilde{N}=0.5$ (dashed line), $\tilde{N}=5$
4804: (dashed-dotted line). }
4805: \label{ftfig20}
4806: \end{figure}
4807:
4808:
4809: For small interatomic separation, the mixed state of the system is
4810: composed of two states: the TPE state $\ket {\Upsilon_{1}}$ and the
4811: antisymmetric state $\ket a$. We illustrate this by diagonalizing
4812: the steady-state density matrix of the system
4813: \begin{eqnarray}
4814: \hat{\rho} &=& \left(
4815: \begin{array}{cccc}
4816: \rho_{gg} & 0 & 0 & \rho_{ge} \\
4817: 0 & \rho_{aa} & 0 & 0 \\
4818: 0 & 0 & \rho_{ss} & 0 \\
4819: \rho_{eg} & 0 & 0 & \rho_{ee}
4820: \end{array}
4821: \right) \ , \label{t223}
4822: \end{eqnarray}
4823: from which we find the new (entangled) states
4824: \begin{eqnarray}
4825: \ket {\Upsilon_{1}} &=& \left[\left(P_{1}-\rho_{ee}\right)\ket g +
4826: \rho_{eg}\ket e
4827: \right]/\left[\left(P_{1}-\rho_{ee}\right)^{2}
4828: +\left|\rho_{eg}\right|^{2}\right]^{\frac{1}{2}} \ ,\nonumber \\
4829: \ket {\Upsilon_{2}} &=& \left[\rho_{ge}\ket g +
4830: \left(P_{2}-\rho_{gg}\right)\ket e
4831: \right]/\left[\left(P_{2}-\rho_{gg}\right)^{2}
4832: +\left|\rho_{eg}\right|^{2}\right]^{\frac{1}{2}} \ ,\nonumber \\
4833: \ket {\Upsilon_{3}} &=& \ket s \ ,\nonumber \\
4834: \ket {\Upsilon_{4}} &=& \ket a \ ,\label{t224}
4835: \end{eqnarray}
4836: where the diagonal probabilities (populations of the entangled states) are
4837: \begin{eqnarray}
4838: P_{1} &=& \frac{1}{2}\left(\rho_{gg}+\rho_{ee}\right)
4839: +\frac{1}{2}\left[\left(\rho_{gg}-\rho_{ee}\right)^{2}
4840: +4\left|\rho_{eg}\right|^{2}\right]^{\frac{1}{2}} \ ,\nonumber \\
4841: P_{2} &=& \frac{1}{2}\left(\rho_{gg}+\rho_{ee}\right)
4842: -\frac{1}{2}\left[\left(\rho_{gg}-\rho_{ee}\right)^{2}
4843: +4\left|\rho_{eg}\right|^{2}\right]^{\frac{1}{2}} \ ,\nonumber \\
4844: P_{3} &=& \rho_{ss} \ ,\nonumber \\
4845: P_{4} &=& \rho_{aa} \ .\label{t225}
4846: \end{eqnarray}
4847: %
4848: \begin{figure}[t]
4849: \begin{center}
4850: \includegraphics[width=13cm]{ftfig21.eps}
4851: \end{center}
4852: \caption{Populations of the entangled states~(\ref{t224})
4853: as a function of the interatomic separation for (a) quantum squeezed
4854: field with $|\tilde{M}|^{2}=\tilde{N}(\tilde{N}+1)$, (b)
4855: classical squeezed field with $|\tilde{M}|=\tilde{N}$, and $\bar{\mu}
4856: \perp \bar{r}_{12}$, $\tilde{N}=0.5$. In both frames $P_{1}$
4857: (solid line), $P_{2}$ (dashed line), $P_{3}$ (dashed-dotted line),
4858: $P_{4}$ (dotted line). }
4859: \label{ftfig21}
4860: \end{figure}
4861:
4862: Note, that the states $\ket {\Upsilon_{1}}, \ket {\Upsilon_{2}}$ and
4863: $\ket {\Upsilon_{3}}$ are the same as for the small sample model,
4864: discussed in the preceding section. This means that the presence of
4865: the antisymmetric state does not affect the two-photon entangled
4866: states, but it can affect the population distribution between the
4867: states and the purity of the system. In Fig.~\ref{ftfig21}, we plot
4868: the populations $P_{i}$ of the states $\ket {\Upsilon_{i}}$ as a
4869: function of the interatomic separation. The figure demonstrates that
4870: in the case of a quantum squeezed field the atoms are driven into a
4871: mixed state composed of {\it only} two entangled states
4872: $\ket {\Upsilon_{1}}$ and $\ket a$, and there is a vanishing
4873: probability that the system is in the states $\ket {\Upsilon_{2}}$
4874: and $\ket s$. In contrast, for a classical squeezed field, shown in
4875: Fig.~\ref{ftfig21}(b), the atoms are driven to a mixed state composed
4876: of all the entangled states.
4877:
4878: Following the discussion presented in Sec.~\ref{ftsec31}, we can argue
4879: that the system can decay to the pure TPE state $\ket {\Upsilon_{1}}$
4880: with the interatomic separation included. This can happen when
4881: the observation time is shorter than $\Gamma ^{-1}$.
4882: The antisymmetric state $\left|a\right\rangle $ decays on a time scale
4883: $\sim \left( \Gamma -\Gamma_{12}\right) ^{-1}$, and for
4884: $\Gamma _{12}\approx \Gamma $ the decay rate of the antisymmetric state
4885: is much longer than $\Gamma ^{-1}$. By contrast, the
4886: state $\left| s\right\rangle $ decays on a time scale $\sim \left( \Gamma
4887: +\Gamma _{12}\right) ^{-1},$ which for $\Gamma _{12}\approx \Gamma $ is
4888: shorter than $\Gamma ^{-1}$. Clearly, for observation times shorter
4889: than $\Gamma^{-1},$ the antisymmetric state does not participate in the
4890: interaction and the system reaches the steady-state only between the
4891: triplet states. Thus, for perfect matching of the squeezed
4892: modes to the atoms the symmetric state is not populated and then the
4893: system is in the pure TPE state $\ket {\Upsilon_{1}}$.
4894:
4895:
4896: \subsection{Two-photon entangled states for two non-identical
4897: atoms}\label{ftsec103}
4898:
4899:
4900: We now extend the analysis of the population distribution in a
4901: squeezed vacuum to the case of two nonidentical atoms.
4902: For two nonidentical atoms with $\Delta \neq 0$ and
4903: $\Gamma_{1}=\Gamma_{2}=\Gamma$, the master equation~(\ref{t36}) leads
4904: to the following equations of motion for the density matrix elements
4905: \begin{eqnarray}
4906: \dot{\rho}_{ee} &=&
4907: -2\Gamma\left(\tilde{N}+1\right)\rho_{ee}
4908: +\tilde{N}\left[\Gamma \left(\rho_{ss}
4909: +\rho_{aa}\right) +
4910: \Gamma_{12}\left(\rho_{ss}-\rho_{aa}\right)e^{i\Delta t}\right] \nonumber \\
4911: &&+\Gamma_{12}|\tilde{M}|\left(\rho_{eg}
4912: e^{-i\left[2\left(\omega_{s}-\omega_{0}\right)t +\phi_{s} \right]}
4913: +\rho_{ge}
4914: e^{i\left[2\left(\omega_{s}-\omega_{0}\right)t +\phi_{s} \right]}
4915: \right) \ ,\nonumber \\
4916: \dot{\rho}_{ss} &=&
4917: \left(\Gamma +\Gamma_{12}e^{i\Delta t}\right)\left[\tilde{N}
4918: -\left(3\tilde{N}+1\right)\rho_{ss} -\tilde{N}\rho_{aa}
4919: +\rho_{ee}\right] \nonumber \\
4920: && -\Gamma |\tilde{M}|\left(\rho_{eg}
4921: e^{-i\left[2\left(\omega_{s}-\omega_{1}\right)t +\phi_{s} \right]}
4922: +\rho_{ge}
4923: e^{i\left[2\left(\omega_{s}-\omega_{1}\right)t +\phi_{s} \right]}
4924: \right) \nonumber \\
4925: && -\Gamma_{12} |\tilde{M}|\left(\rho_{eg}
4926: e^{-i\left[2\left(\omega_{s}-\omega_{0}\right)t +\phi_{s} \right]}
4927: +\rho_{ge}
4928: e^{i\left[2\left(\omega_{s}-\omega_{0}\right)t +\phi_{s} \right]}
4929: \right) \ ,\nonumber \\
4930: \dot{\rho}_{aa} &=&
4931: \left(\Gamma -\Gamma_{12}e^{i\Delta t}\right)\left[\tilde{N}
4932: -\left(3\tilde{N}+1\right)\rho_{aa} -\tilde{N}\rho_{ss}
4933: +\rho_{ee}\right] \nonumber \\
4934: && +\Gamma |\tilde{M}|\left(\rho_{eg}
4935: e^{-i\left[2\left(\omega_{s}-\omega_{1}\right)t +\phi_{s} \right]}
4936: +\rho_{ge}
4937: e^{i\left[2\left(\omega_{s}-\omega_{1}\right)t +\phi_{s} \right]}
4938: \right) \nonumber \\
4939: && -\Gamma_{12} |\tilde{M}|\left(\rho_{eg}
4940: e^{-i\left[2\left(\omega_{s}-\omega_{0}\right)t +\phi_{s} \right]}
4941: +\rho_{ge}
4942: e^{i\left[2\left(\omega_{s}-\omega_{0}\right)t +\phi_{s} \right]}
4943: \right) \ ,\nonumber \\
4944: \dot{\rho}_{eg} &=&
4945: \left(\dot{\rho}_{ge}\right)^{\ast} =
4946: \Gamma_{12}|\tilde{M}|
4947: e^{i\left[2\left(\omega_{s}-\omega_{1}\right)t +\phi_{s} \right]}
4948: -\left(2\tilde{N}+1\right)\Gamma \rho_{eg} \nonumber \\
4949: && -\Gamma |\tilde{M}|
4950: e^{i\left[2\left(\omega_{s}-\omega_{1}\right)t +\phi_{s} \right]}
4951: \left(\rho_{ss} -\rho_{aa}\right) \nonumber \\
4952: && -2\Gamma_{12}|\tilde{M}|
4953: e^{i\left[2\left(\omega_{s}-\omega_{0}\right)t +\phi_{s} \right]}
4954: \left(\rho_{ss}+\rho_{aa}\right) \ ,\label{t226}
4955: \end{eqnarray}
4956: where $\omega_{0}=\frac{1}{2}\left(\omega_{1}+\omega_{2}\right)$.
4957:
4958: Equations~(\ref{t226}) contain time-dependent terms which oscillate
4959: at frequencies exp$(\pm i\Delta t)$ and exp$[\pm
4960: 2i(\omega_{s}-\omega_{0})t +\phi_{s}]$. If we tune the squeezed
4961: vacuum field to the middle of the frequency difference between the
4962: atomic frequencies, i.e. $\omega_{s} =(\omega_{1}+\omega_{2})/2$, the
4963: terms proportional to exp$[\pm 2i(\omega_{s}-\omega_{0})t +\phi_{s}]$
4964: become stationary in time. None of the other time dependent components
4965: is resonant with the frequency of the squeezed vacuum field.
4966: Consequently, for $\Delta \gg \Gamma$, the
4967: time-dependent components oscillate rapidly in time
4968: and average to zero over long times. Therefore, we can make a
4969: secular approximation in which we ignore the rapidly oscillating
4970: terms, and find the following steady-state solutions~\cite{fw97}
4971: \begin{eqnarray}
4972: \rho_{ee}
4973: &=&\frac{1}{4}\left\{\frac{\left(2\tilde{N}-1\right)}{2\tilde{N}+1}+
4974: \frac{1}{\left[\left(2\tilde{N}+1\right)^{2}
4975: -4a^{2}|\tilde{M}|^{2}\right] }\right\}
4976: \ ,\nonumber \\
4977: \rho_{ss} &=& \rho_{aa} =\frac{1}{4}\left\{ 1-\frac{1}{\left[
4978: \left(2\tilde{N}+1\right)^{2}-4a^{2}|\tilde{M}|^{2}\right] }\right\}
4979: \ ,\nonumber \\
4980: \rho_{u} &=&
4981: \frac{2a|\tilde{M}|}{\left(2\tilde{N}+1\right)
4982: \left[\left(2\tilde{N}+1\right)^{2}
4983: -4a^{2}|\tilde{M}|^{2}\right]}
4984: \ .\label{t227}
4985: \end{eqnarray}
4986: Equations (\ref{t227}) are quite different from Eqs.~(\ref{t220})
4987: and show that in the case of non-identical atoms
4988: the symmetric and antisymmetric states are equally
4989: populated independent of the interatomic separation. These are,
4990: however, some similarities to the steady-state solutions of the Dicke
4991: model that for small interatomic separations
4992: $\rho_{ss}=\rho_{aa} \approx 0$, and then only the collective ground
4993: and the upper states are populated.
4994:
4995: To conclude this section, we point out that by employing two spatially
4996: separated non-identical atoms of significantly different
4997: transition frequencies $(\Delta \gg \Gamma)$, it is possible to achieve
4998: the pure TPE state with the antisymmetric state fully participating in
4999: the interaction.
5000:
5001:
5002: \section{Mapping of entangled states of light on
5003: atoms}\label{ftsec104}
5004:
5005:
5006: The generation of the pure TPE state is an example of mapping of a
5007: state of quantum correlated light onto an atomic system. The
5008: two-photon correlations contained in the squeezed vacuum field can be
5009: completely transferred to the atomic system and can be measured, for
5010: example, by detecting fluctuations of the fluorescence field emitted by
5011: the atomic system. Squeezing in the fluorescence field is proportional to
5012: the squeezing in the atomic dipole operators (spin squeezing) which, on
5013: the other hand, can be found from the steady-state solutions for the
5014: density matrix elements.
5015:
5016:
5017: \subsection{Mapping of photon correlations}\label{ftsec1041}
5018:
5019:
5020: Equation~(\ref{t227}) shows that the collective damping parameter
5021: $\Gamma_{12}$ plays the role of a degree of the correlation
5022: transfer from the squeezed vacuum to the atomic system. For large
5023: interatomic separations, $\Gamma_{12} \approx 0$, and there is no
5024: transfer of the correlations to the system. In contrast, for very
5025: small separations, $\Gamma_{12}\approx \Gamma$, and then the
5026: correlations are completely transferred to the atomic system.
5027:
5028: However, the complete transfer of the correlations does not necessary
5029: mean that the two-photon correlations are stored in the pure TPE state.
5030: This happens only for two nonidentical atoms in the Dicke model, where
5031: the steady-state is the pure
5032: TPE state. For identical atoms separated by a finite distance $r_{12}$
5033: only a part of the correlations can be stored in the antisymmetric state.
5034: This can be shown, for example, by calculating of the interference
5035: pattern of the fluorescence field emitted by the system. Using the
5036: steady-state solutions~(\ref{t220}), we find that the visibility in
5037: the interference pattern is given by~\cite{ft}
5038: \begin{equation}
5039: {\cal{V}} =
5040: -\frac{2a|\tilde{M}|^{2}}{\tilde{N}\left(2\tilde{N}+1\right)^{3}
5041: +2|\tilde{M}|^{2}\left[a^{2}+\left(2\tilde{N}+1\right)
5042: -\left(2\tilde{N}+1\right)^{2}\right]} \ .\label{t228}
5043: \end{equation}
5044:
5045: \begin{figure}[t]
5046: \begin{center}
5047: \includegraphics[width=13cm]{ftfig22.eps}
5048: \end{center}
5049: \caption{The visibility ${\cal {V}}$ as a function of the interatomic
5050: separation $r_{12}$ for (a) a quantum squeezed field with
5051: $|\tilde{M}|^{2}=\tilde{N}(\tilde{N}+1)$, (b) a classical
5052: squeezed field with $|\tilde{M}|=\tilde{N}$,
5053: $\bar{\mu} \perp \bar{r}_{12}$ and different $\tilde{N}$:
5054: $\tilde{N}=0.05$ (solid line), $\tilde{N}=0.5$ (dashed line),
5055: $\tilde{N}=5$ (dashed-dotted line). }
5056: \label{ftfig22}
5057: \end{figure}
5058: %
5059: The visibility is negative indicating that the squeezing
5060: correlations are mostly stored in the antisymmetric state.
5061: In Fig.~\ref{ftfig22}, we plot the visibility ${\cal{V}}$ as a function
5062: of the interatomic separation for a quantum squeezed field with
5063: $|\tilde{M}|^{2}=\tilde{N}(\tilde{N}+1)$, Fig.~\ref{ftfig22}(a), and a
5064: classical squeezed field with $|\tilde{M}|=\tilde{N}$,
5065: Fig.~\ref{ftfig22}(b). An interference pattern with a dark center is
5066: observed for small interatomic separations $(r_{12}<\lambda/2 )$ and
5067: with the quantum squeezed field the visibility attains
5068: the maximal negative value of
5069: ${\cal{V}}\approx -0.7$ for $r_{12}< 0.3\lambda$. According to
5070: Eq.~(\ref{t222}), at these interatomic separations the antisymmetric
5071: state is the most populated state of the system. The value
5072: ${\cal{V}}= -0.7$ compared to the possible negative
5073: value ${\cal{V}}=-1$ indicates that $70\%$ of the squeezing
5074: correlations are stored in the antisymmetric state.
5075: In Fig.~\ref{ftfig22}(b), we show the visibility for a classical
5076: squeezed field with $|\tilde{M}|=\tilde{N}$. The visibility is much
5077: smaller than that in the quantum squeezed field and vanishes when
5078: $\tilde{N}\rightarrow \infty$. In contrast, for the quantum squeezed
5079: field $V$ approaches $-1/2$ when $\tilde{N}\rightarrow \infty$. Thus,
5080: the visibility can provide the information about the degree of
5081: nonclassical correlations stored in the entangled state $\ket a$.
5082:
5083:
5084: \subsection{Mapping of the field fluctuations}\label{ftsec1042}
5085:
5086:
5087: The fluctuations of the electric field are determined by the normally
5088: ordered variance of the field operators as~\cite{park,fd1,fd2}
5089: \begin{equation}
5090: \langle :\left(\Delta E_{\theta}\right)^{2}:\rangle =
5091: \sum_{\vec{k}s}E_{k}\left( 2\langle \hat{a}_{\vec{k}s}^{\dagger}
5092: \hat{a}_{\vec{k}s}\rangle +\langle \hat{a}_{\vec{k}s}\hat{a}_{\vec{k}s}
5093: \rangle e^{2i\theta}
5094: + \langle \hat{a}_{\vec{k}s}^{\dagger}\hat{a}_{\vec{k}s}^{\dagger} \rangle
5095: e^{-2i\theta}\right) \ .\label{t229}
5096: \end{equation}
5097: Using the correlation functions~(\ref{t17}) of the three-dimensional
5098: squeezed vacuum field and choosing $\theta
5099: =\pi/2$, the variance of the incident squeezed vacuum field can be
5100: written as
5101: \begin{equation}
5102: \left\langle :\left(\Delta E^{in}_{\pi/2}\right)^{2}:\right\rangle =
5103: 2E_{0}\left(\tilde{N} -|\tilde{M}|\right) \ ,\label{t230}
5104: \end{equation}
5105: where $E_{0}$ is a constant. Since
5106: $|\tilde{M}|=\sqrt{\tilde{N}(\tilde{N}+1)}> \tilde{N}$, the
5107: variance~(\ref{t230}) is negative indicating that the incident field
5108: is in a squeezed state.
5109:
5110: On the other hand, the normally ordered variance of the emitted
5111: fluorescence field can be expressed in terms of the density matrix
5112: elements of the two-atom system as
5113: \begin{equation}
5114: \left\langle :\left(\Delta E^{out}_{\theta}\right)^{2}:\right\rangle =
5115: E_{0}\left(2\rho_{ss}+2\rho_{ee} +|\rho_{u}|\cos 2\theta \right)
5116: \ .\label{t231}
5117: \end{equation}
5118: Using the steady-state solutions~(\ref{t210}) and choosing $\theta
5119: =\pi/2$, we find
5120: \begin{equation}
5121: \left\langle :\left(\Delta E^{out}_{\pi/2}\right)^{2}:\right\rangle =
5122: 2E_{0}\frac{\left(\tilde{N} -|\tilde{M}|\right)}{2\tilde{N}+1}
5123: \ .\label{t232}
5124: \end{equation}
5125: Thus, at low intensities of the squeezed vacuum field $(\tilde{N}\ll 1)$
5126: the fluctuations in the incident field are perfectly
5127: mapped onto the atomic system. For large intensities $(\tilde{N}>1)$, the
5128: thermal fluctuations of the atomic dipoles dominate over the squeezed
5129: fluctuations resulting in a reduction of squeezing in the fluorescence
5130: field.
5131:
5132: The idea of mapping the field fluctuations on the collective system of
5133: two atoms have been extended to multi-level atoms. For example,
5134: Kozhekin {\it et al.}~\cite{koz} proposed a method of mapping of
5135: quantum states onto an atomic system based on the stimulated Raman
5136: absorption of propagating quantum light by a cloud of three-level atoms.
5137: Hald {\it et al.}~\cite{hal1,hal2} have experimentally observed
5138: the squeezed spin states of trapped three-level atoms in the $\Lambda$
5139: configuration, irradiated by a squeezed field. The observed squeezed
5140: spin states have been generated via entanglement exchange with the
5141: squeezed field that was completely absorbed by the atoms. The exchange
5142: process was, however, accomplished by spontaneous emission and only
5143: a limited amount of spin squeezing was achieved.
5144: Fleischhauer {\it et al.}~\cite{flei} have considered
5145: a similar system of three-level atoms and have found that quantum
5146: states of single-photon fields can be mapped onto entangled states
5147: of the field and the collective states of the atoms. This effect
5148: arises from a substantial reduction of the group velocity of the field
5149: propagating through the atomic system, which results in a temporary
5150: storage of a quantum state of the field in atomic spins.
5151: These models are two examples of the continuing fruitful investigation
5152: of entanglement and reversible storage of information in collective
5153: atomic systems.
5154:
5155:
5156: \section{Conclusions}\label{ftsec11}
5157:
5158: In this paper, we have reviewed the recent work on entanglement and
5159: nonclassical effects in two-atom systems. We have discussed different
5160: schemes for generation of nonclassical states of light and preparation
5161: of two interacting atoms in specific entangled state. In particular,
5162: we have presented different methods of preparing two atoms in the
5163: antisymmetric state which is an example of a decoherence-free entangled
5164: state. The ability to prepare two-atom system in the decoherence-free
5165: state represents the ultimate quantum control of a physical system and
5166: opens the door for a number of applications ranging from quantum
5167: information, quantum computing to high-resolution spectroscopy.
5168: However, the practical implementation of entanglement in information
5169: processing and quantum computation requires coherent manipulation of a
5170: large number of atoms, which is not an easy task. Although the two-atom
5171: systems, discussed in this review, are admittedly elementary models, they
5172: offers some advantages over the multiatom problem. Because of their
5173: simplicity, we have obtained detailed and almost exact solutions
5174: that can be easily interpreted physically, and thus provide insight into
5175: the behavior of more complicated multiatom systems. Moreover, many results
5176: discussed in this review is analogous to phenomena that one would expect
5177: in multiatom systems. For example, the nonexponential decay of the
5178: total radiation intensity from two nonidentical atoms is an elementary
5179: example of superradiant pulse formation, and a manifestation of the
5180: presence of coherences between the collective entangled states.
5181: A number of theoretical studies have been performed recently on
5182: entanglement and irreversible dynamics of a large number of
5183: atoms~\cite{bbtk,koz,hal1,hal2,flei,sm01,b1,b2,nat1,nat2,nat3}.
5184: These studies, however, have been limited to the Dicke model that
5185: ignores antisymmetric states of a multi-atom system. Nevertheless, the
5186: calculations have shown that population (information) can be stored in
5187: the collective atomic states or in the so called
5188: dark-state polaritons~\cite{fl00,fg02}, which are quasiparticles
5189: associated with electromagnetically induced transparency in
5190: multi-level atomic systems.
5191:
5192:
5193: \section*{Acknowledgments}
5194: This work was supported by the Australian Research Council.
5195:
5196:
5197: \begin{thebibliography}{9}
5198:
5199:
5200: \bibitem{pv} M.B. Plenio and V. Vedral, Contemp. Phys. {\bf 39}, 431
5201: (1998).
5202:
5203: \bibitem{schr} E. Schr\"{o}dinger, Naturwissenschaften {\bf 23},
5204: 807 (1935).
5205:
5206: \bibitem{beke} A. Barenco and A. Ekert, J. Mod. Opt. {\bf 42}, 1253
5207: (1995).
5208:
5209: \bibitem{pok} S.F. Pereira, Z.Y. Ou, and H.J. Kimble, Phys. Rev. A
5210: {\bf 62}, 042311 (2000).
5211:
5212: \bibitem{ek} A. Ekert, Phys. Rev. Lett. {\bf 67}, 661 (1991).
5213:
5214: \bibitem{bar} A. Barenco, Contemp. Phys. {\bf 37}, 375 (1996).
5215:
5216: \bibitem{gro} L.K. Grover, Phys. Rev. Lett. {\bf 79}, 325 (1997).
5217:
5218: \bibitem{boll1} J.J. Bollinger, W.M. Itano, D.J. Wineland, and D.J.
5219: Heinzen, Phys. Rev. A {\bf 54}, R4649 (1996).
5220:
5221: \bibitem{boll2} S.F. Huelga, C. Macchiavello, T. Pellizzari, A.K. Ekert,
5222: M.B.~Plenio, and J.I. Cirac, Phys. Rev. Lett. {\bf 79}, 3865 (1997).
5223:
5224: \bibitem{dic} R.H. Dicke, Phys. Rev. {\bf 93}, 99 (1954).
5225:
5226: \bibitem{ftk87} Z. Ficek, R. Tana\'s, and S. Kielich, Physica {\bf
5227: 146A}, 452 (1987).
5228:
5229: \bibitem{buz} V. Buzek, Phys. Rev. A {\bf 39}, 2232 (1989).
5230:
5231: \bibitem{eich} U. Eichmann, J.C. Bergquist, J.J. Bollinger, J.M. Gilligan,
5232: W.M. Itano, D.J. Wineland, and M.G. Raizen, Phys. Rev. Lett. {\bf 70}, 2359
5233: (1993).
5234:
5235: \bibitem{deb} R.G. DeVoe and R.G. Brewer, Phys. Rev. Lett. {\bf 76},
5236: 2049 (1996).
5237:
5238: \bibitem{ber} D.J. Berkeland, in {\it Laser Spectroscopy} edited by
5239: R. Blatt, J. Eschner, D. Leibfried and F. Schmidt-Kaler (World
5240: Scientific, Singapore, 1999), p. 352.
5241:
5242: \bibitem{tos} D. Riesch, K. Abich, W. Neuhauser, Ch. Wunderlich, and
5243: P.E. Toschek, Phys. Rev. A {\bf 65}, 053401 (2002).
5244:
5245: \bibitem{ag70} G.S. Agarwal, Phys. Rev. A {\bf 2}, 2038 (1970).
5246:
5247: \bibitem{bbtk} A. Beige, D. Braun, B. Tregenna, and P.L. Knight, Phys.
5248: Rev. Lett. {\bf 85}, 1762 (2000).
5249:
5250: \bibitem{afs} U. Akram, Z. Ficek, and S. Swain, Phys. Rev. A {\bf 62},
5251: 013413 (2000).
5252:
5253: \bibitem{bcjd} G.K. Brennen, C.M. Caves, P.S. Jessen, and I.H.
5254: Deutsch, Phys. Rev. Lett. {\bf 82}, 1060 (1999).
5255:
5256: \bibitem{bdj} G.K. Brennen, I.H. Deutsch, and P.S. Jessen, Phys. Rev.
5257: A {\bf 61}, 062309 (2000).
5258:
5259: \bibitem{tbk} B. Tregenna, A. Beige, and P.L. Knight, Phys. Rev. A
5260: {\bf 65}, 032305 (2002).
5261:
5262: \bibitem{mas} S.M. Barnett and M.A. Dupertuis, J. Opt. Soc. Am. B
5263: {\bf 4}, 505 (1987).
5264:
5265: \bibitem{pas1} G.M. Palma and P.L. Knight, Phys. Rev. A {\bf 39}, 1962
5266: (1989).
5267:
5268: \bibitem{pas2} A.K. Ekert, G.M. Palma, S.M. Barnett, and P.L. Knight,
5269: Phys. Rev. A {\bf 39}, 6026 (1989).
5270:
5271: \bibitem{pas3} G.S.~Agarwal and R.R. Puri, Phys. Rev. A
5272: {\bf 41}, 3782 (1990).
5273:
5274: \bibitem{pas4} Z. Ficek, Phys. Rev. A {\bf 42}, 611 (1990).
5275:
5276: \bibitem{pas5} Z. Ficek, Phys. Rev. A {\bf 44}, 7759 (1991).
5277:
5278: \bibitem{park} A.S. Parkins, in {\it Modern Nonlinear Optics, Part II},
5279: eds. M. Evans and S. Kielich (Wiley, New York, 1993), p.607.
5280:
5281: \bibitem{fd1} Z. Ficek and P.D. Drummond, Physics Today (September
5282: 1997), p.34.
5283:
5284: \bibitem{fd2} B.J. Dalton, Z. Ficek, and S. Swain, J. Mod. Opt.
5285: {\bf 46}, 379 (1999).
5286:
5287: \bibitem{tur} Q.A. Turchette, C.S. Wood, B.E. King, C.J. Myatt, D.
5288: Leibfried, W.M. Itano, C.~Monroe, and D.J. Wineland, Phys. Rev. Lett.
5289: {\bf 81}, 3631 (1998).
5290:
5291: \bibitem{hag} E. Hagley, X. Maitre, G. Nogues, C. Wunderlich, M.
5292: Brune, J.M. Raimond, and S. Haroche, Phys. Rev. Lett. {\bf 79}, 1
5293: (1997).
5294:
5295: \bibitem{rbh} J.M. Raimond, M. Brune, and S. Haroche, Rev. Mod.
5296: Phys. {\bf 73}, 565 (2001).
5297:
5298: \bibitem{kmb} A. Kuzmich, L. Mandel, and N.P. Bigelow, Phys. Rev.
5299: Lett. {\bf 85}, 1594 (2000).
5300:
5301: \bibitem{osna} S. Osnaghi, P. Bertet, A. Auffeves, P. Maioli, M.
5302: Brune, J.M. Raimond, and S. Haroche, Phys. Rev. Lett. {\bf 87},
5303: 037902 (2001).
5304:
5305: \bibitem{ymbrh} F. Yamaguchi, P. Milman, M. Brune, J.M. Raimond,
5306: and S. Haroche, Phys. Rev. A {\bf 66}, 010302(R) (2002).
5307:
5308: \bibitem{mb1} T. Pellizzari, Phys. Rev. Lett. {\bf 79}, 5242 (1997).
5309:
5310: \bibitem{mb2} A. Sorensen and K. Molmer, Phys. Rev. A {\bf 58},
5311: 2745 (1998).
5312:
5313: \bibitem{mb3} S.J. van Enk, H.J. Kimble, J.I. Cirac, and P. Zoller,
5314: Phys. Rev. A {\bf 59}, 2659 (1999).
5315:
5316: \bibitem{mb4} A.S. Parkins and H.J. Kimble, Phys. Rev. A {\bf 61},
5317: 052104 (2000).
5318:
5319: \bibitem{mb5} S. Mancini and S. Bose, Phys. Rev. A {\bf 64},
5320: 032308 (2001).
5321:
5322: \bibitem{prsg} I.E. Protsenko, G. Reymond, N. Schlosser, and
5323: P. Grangier, Phys. Rev. A {\bf 65}, 052301 (2002).
5324:
5325: \bibitem{fg1} H. Pu and P. Meystre, Phys. Rev. Lett. {\bf 85},
5326: 3987 (2000).
5327:
5328: \bibitem{fg2} L.M. Duan, A. Sorensen, J.I. Cirac, and P. Zoller,
5329: Phys. Rev. Lett. {\bf 85}, 3991 (2000).
5330:
5331: \bibitem{schl} N. Schlosser, G. Reymond, I. Protsenko, and P.
5332: Grangier, Nature {\bf 411}, 1024 (2001).
5333:
5334: \bibitem{leh} R.H. Lehmberg, Phys. Rev. A {\bf 2}, 883; {\bf 2}, 889
5335: (1970).
5336:
5337: \bibitem{lui} W.H. Louisell, {\it Statistical Properties of Radiation},
5338: (Wiley, New York, 1973).
5339:
5340: \bibitem{ag74} G.S. Agarwal, {\it Quantum Statistical Theories of Spontaneous
5341: Emission and their Relation to other Approaches}, edited by G. H\"{o}hler,
5342: Springer Tracts in Modern Physics, Vol. 70, (Springer-Verlag , Berlin, 1974).
5343:
5344: \bibitem{car93} H. Carmichael, {\it An Open Systems Approach to
5345: Quantum Optics}, Lecture Notes in Physics, Vol. {\bf 18}, (Springer-Verlag ,
5346: Berlin, 1993).
5347:
5348: \bibitem{dcm92} J. Dalibard, Y. Castin, and K. Molmer, Phys. Rev. Lett.
5349: {\bf 68}, 580 (1992).
5350:
5351: \bibitem{pk98} M.B. Plenio and P.L. Knight, Rev. Mod. Phys. {\bf 70},
5352: 101 (1998).
5353:
5354: \bibitem{pg89} A.S. Parkins and C.W. Gardiner, Phys. Rev. A {\bf 40},
5355: 2534 (1989).
5356:
5357: \bibitem{kimb1} Q.A. Turchette, N.Ph. Georgiades, C.J. Hood, H.J. Kimble,
5358: and A.S. Parkins, Phys. Rev. A {\bf 58}, 4056 (1998).
5359:
5360: \bibitem{kimb2} N.Ph. Georgiades, E.S. Polzik and H.J. Kimble, Phys. Rev.
5361: A {\bf 59}, 123 (1999).
5362:
5363: \bibitem{bw} M. Born and E. Wolf, {\it Principles of Optics} (Macmillan,
5364: New York, 1964), Chap.7.
5365:
5366: \bibitem{yar} A. Yariv, {\it Quantum Electronics} (Wiley, New York, 1989).
5367:
5368: \bibitem{fde} Z. Ficek and P.D. Drummond, Europhys. Lett. {\bf 24}, 455
5369: (1993).
5370:
5371: \bibitem{fd91} Z. Ficek and P.D. Drummond, Phys. Rev. A {\bf 43}, 6247
5372: (1991); {\bf 43}, 6258 (1991).
5373:
5374: \bibitem{ae} L. Allen and J.H. Eberly, {\it Resonance Fluorescence and
5375: Two-Level Atoms}, (Wiley, New York, 1975).
5376:
5377: \bibitem{mk74} P.W. Milonni and P.L. Knight, Phys. Rev. A {\bf 10},
5378: 1096 (1974).
5379:
5380: \bibitem{mk75} P.W. Milonni and P.L. Knight, Phys. Rev. A {\bf 11},
5381: 1090 (1975).
5382:
5383: \bibitem{sz97} M.O. Scully and M.S. Zubairy, {\it Quantum Optics},
5384: (Cambridge University Press, Cambridge, 1997), p. 13.
5385:
5386: \bibitem{mil} G.J. Milburn, Phys. Rev. A {\bf 34}, 4882 (1986).
5387:
5388: \bibitem{oc} G.W. Ford and R.F. O'Connell, J. Opt. Soc. Am. B {\bf 4},
5389: 1710 (1987).
5390:
5391: \bibitem{pk89} G.M. Palma and P.L. Knight, Optics Commun. {\bf 73},
5392: 131 (1989).
5393:
5394: \bibitem{st64} M.J. Stephen, J. Chem. Phys. {\bf 40}, 669 (1964).
5395:
5396: \bibitem{hh64} A. Hutchinson and H.F. Hameka, J. Chem. Phys. {\bf
5397: 41}, 2006 (1964).
5398:
5399: \bibitem{kbr87} G. Kurizki and A. Ben-Reuven, Phys. Rev. A {\bf 36},
5400: 90 (1987).
5401:
5402: \bibitem{bgz00} I.V. Bargatin, B.A. Grishanin, and V.N. Zadkov, Phys.
5403: Rev. A {\bf 61}, 052305 (2000).
5404:
5405: \bibitem{bg02} S.U. Addicks, A. Beige, M. Dakna, and G.C. Hegerfeldt,
5406: Eur. Phys. J. D {\bf 15}, 393 (2001).
5407:
5408: \bibitem{ryf} T. Rudolph, I. Yavin, and H. Freedhoff,
5409: quant-ph/0206067.
5410:
5411: \bibitem{lm88} M. Lewenstein and T.W. Mossberg, Phys. Rev. A {\bf
5412: 37}, 2048 (1988).
5413:
5414: \bibitem{kky} G. Kurizki, A.G. Kofman, and V. Yudson, Phys. Rev. A
5415: {\bf 53}, R35 (1996).
5416:
5417: \bibitem{fwd97} Z. Ficek, B. J. Dalton and M. R. B. Wahiddin, J. Mod.
5418: Opt. {\bf 44}, 1005 (1997).
5419:
5420: \bibitem{mtf99} A. Messikh, R. Tana\'s, and Z. Ficek, Phys. Rev. A {\bf
5421: 61}, 033811 (2000).
5422:
5423: \bibitem{lambro} G.M. Nikolopoulos and P. Lambropoulos, J. Mod.
5424: Opt. {\bf 49}, 61 (2002).
5425:
5426: \bibitem{dkw} H.T. Dung, L. Kn\"{o}ll, and D.G. Welsch,
5427: quant-ph/0205056.
5428:
5429: \bibitem{hel87} H.S. Freedhoff, J. Phys. B {\bf 22}, 435 (1989).
5430:
5431: \bibitem{bh98} A. Beige and G.C. Hegerfeldt, Phys. Rev. A {\bf 58},
5432: 4133 (1998).
5433:
5434: \bibitem{cabr} C. Cabrillo, J.I. Cirac, P. Garcia-Fernandez, and P.
5435: Zoller, Phys. Rev. A {\bf 59}, 1025 (1999).
5436:
5437: \bibitem{sb} C. Sch\"{o}n and A. Beige, Phys. Rev. A {\bf 64}, 023806
5438: (2001).
5439:
5440: \bibitem{ftk81} Z. Ficek, R. Tana\'s, and S. Kielich, Optics
5441: Commun. {\bf 36}, 121 (1981).
5442:
5443: \bibitem{ftk83} Z. Ficek, R. Tana\'s, and S. Kielich, Optica Acta {\bf
5444: 30}, 713 (1983).
5445:
5446: \bibitem{hsf82} H.S. Freedhoff, Phys. Rev. A {\bf 26}, 684 (1982).
5447:
5448: \bibitem{mw} L.~Mandel and E.~Wolf, {\it Optical Coherence and Quantum
5449: Optics}, (Cambridge, New York, 1995).
5450:
5451: \bibitem{ac78} A.S.J. Amin and J.G. Cordes, Phys. Rev. A {\bf 18},
5452: 1298 (1978).
5453:
5454: \bibitem{ftk81a} Z. Ficek, R. Tana\'s, and S. Kielich, in LASERS-80,
5455: ed. C.B.Collins (STS Press, McLean, Virginia, 1981) p.800.
5456:
5457: \bibitem{ftk86} Z. Ficek, R. Tana\'s, and S. Kielich, Optica Acta {\bf
5458: 33}, 1149 (1986).
5459:
5460: \bibitem{fs} Z. Ficek and B.C. Sanders, Phys. Rev. A {\bf 41}, 359
5461: (1990).
5462:
5463: \bibitem{rfd} T.G. Rudolph, Z. Ficek, and B.J. Dalton, Phys. Rev. A
5464: {\bf 52}, 636 (1995).
5465:
5466: \bibitem{lm93} G. Lenz and P. Meystre, Phys. Rev. A {\bf 48}, 3365
5467: (1993).
5468:
5469: \bibitem{bl} R. Bonifacio and L.A. Lugiato, Phys. Rev. A {\bf 11},
5470: 1507 (1975).
5471:
5472: \bibitem{gh} M. Gross and S. Haroche, Phys. Rep. {\bf 93}, 301
5473: (1982).
5474:
5475: \bibitem{er} J.H. Eberly and N.E. Rehler, Phys. Rev. A {\bf 2}, 1607
5476: (1970).
5477:
5478: \bibitem{cf78} B. Coffey and R. Friedberg, Phys. Rev. A {\bf 17},
5479: 1033 (1978).
5480:
5481: \bibitem{rich81} Th. Richter, Ann. Phys. {\bf 38}, 106 (1981).
5482:
5483: \bibitem{bla} H. Blank, M. Blank, K. Blum, and A. Faridani, Phys.
5484: Lett. {\bf A105}, 39 (1984).
5485:
5486: \bibitem{dea} E. De Angelis, F. De Martini, and P. Mataloni,
5487: J. Opt. B: Quantum Semiclass. Opt. {\bf 2}, 149 (2000).
5488:
5489: \bibitem{zs98} P. Zhou and S. Swain, J. Opt. Soc. Am. B {\bf 15},
5490: 2593 (1998).
5491:
5492: \bibitem{cw76} H.J. Carmichael and D.F. Walls, J. Phys. B {\bf 9},
5493: 1199 (1976).
5494:
5495: \bibitem{km76} H.J. Kimble and L. Mandel, Phys. Rev. A {\bf 13}, 2123
5496: (1976).
5497:
5498: \bibitem{wal79} D.F. Walls, Nature {\bf 280}, 451 (1979).
5499:
5500: \bibitem{lo80} R. Loudon, Rep. Progr. Phys. {\bf 43}, 913 (1980).
5501:
5502: \bibitem{per80} J. Perina, in {\it Progress in Optics}, edited by
5503: E. Wolf, Vol. XVIII, (North Holland, Amsterdam, 1980), p. 128.
5504:
5505: \bibitem{paul82} H. Paul, Rev. Mod. Phys. {\bf 54}, 1061 (1982).
5506:
5507: \bibitem{wz81} D.F. Walls and P. Zoller, Phys. Rev. Lett. {\bf 47},
5508: 709 (1981).
5509:
5510: \bibitem{man82} L. Mandel, Phys. Rev. Lett. {\bf 49}, 136 (1982).
5511:
5512: \bibitem{kdm77} H.J. Kimble, M. Dagenais, and L. Mandel, Phys. Rev.
5513: Lett. {\bf 39}, 691 (1977).
5514:
5515: \bibitem{kdm78} H.J. Kimble, M. Dagenais, and L. Mandel, Phys. Rev.
5516: A {\bf 18}, 201 (1978).
5517:
5518: \bibitem{cre} J.D. Cresser, J. Hager, G. Leuchs, M.S. Rateike, and
5519: H. Walther, in {\it Dissipative Systems in Quantum Optics}, edited
5520: by R. Bonifacio, (Springer, Berlin, 1982).
5521:
5522: \bibitem{dw87} F. Diedrich and H. Walther, Phys. Rev. Lett. {\bf 58},
5523: 203 (1987).
5524:
5525: \bibitem{fost00} G.T. Foster, S.L. Mielke, and L.A. Orozco, Phys.
5526: Rev. A {\bf 61}, 053821 (2000).
5527:
5528: \bibitem{slush} R.M. Slusher, L.W. Hollberg, B. Yurke, J.C. Mertz
5529: and J.F. Valley, Phys. Rev. Lett. {\bf 55}, 2409 (1985).
5530:
5531: \bibitem{jmo} Special issue of J. Mod. Optics {\bf 34}, nos. 6/7
5532: (1987).
5533:
5534: \bibitem{josab} Special issue of J. Opt. Soc. Am. B {\bf 4}, no. 10
5535: (1987).
5536:
5537: \bibitem{dodo} V.V. Dodonov, J. Opt. B: Quantum Semiclass. Opt. {\bf
5538: 4}, R1 (2002).
5539:
5540: \bibitem{lax} M. Lax, Phys. Rev. {\bf 172}, 350 (1968).
5541:
5542: \bibitem{aga79} G.S. Agarwal, L.M. Narducci, D.H. Feng, and R.
5543: Gilmore, Phys. Rev. Lett. {\bf 42}, 1260 (1979).
5544:
5545: \bibitem{kil80} S.J. Kilin, J. Phys. B {\bf 13}, 2653 (1980).
5546:
5547: \bibitem{has82} S.S. Hassan, G.P. Hildred, R.R. Puri, and S.V.
5548: Lawande, J. Phys. B {\bf 15}, 1029 (1982).
5549:
5550: \bibitem{cor82} J.G. Cordes, J. Phys. B {\bf 15}, 4349 (1982).
5551:
5552: \bibitem{cor87} J.G. Cordes, J. Phys. B {\bf 20}, 1433 (1987).
5553:
5554: \bibitem{fr79} H.S. Freedhoff, Phys. Rev. A {\bf 19}, 1132 (1979).
5555:
5556: \bibitem{hds} G.C. Hegerfeldt and D. Seidel, J. Opt. B: Quantum
5557: Semiclass. Opt. {\bf 4}, 245 (2002).
5558:
5559: \bibitem{law85} S.V. Lawande, B.N. Jagatap, and R.R. Puri,
5560: J. Phys. B {\bf 18}, 1711 (1985).
5561:
5562: \bibitem{rich1} T. Richter, Optica Acta {\bf 29}, 265 (1982).
5563:
5564: \bibitem{rich2} T. Richter, Optica Acta {\bf 30}, 1769 (1983).
5565:
5566: \bibitem{ftk84} Z. Ficek, R. Tana\'s, and S. Kielich, Phys. Rev. A
5567: {\bf 29}, 2004 (1984).
5568:
5569: \bibitem{ftk87s} Z. Ficek, R. Tana\'s, and S. Kielich,
5570: J. Physique {\bf 48}, 1697 (1987).
5571:
5572: \bibitem{ftk84a} Z. Ficek, R. Tana\'s, and S. Kielich,
5573: J. Opt. Soc. Am. B {\bf 1}, 882 (1984).
5574:
5575: \bibitem{bk88} S.M. Barnett and P.L. Knight, Phys. Script. {\bf 21},
5576: 5 (1988).
5577:
5578: \bibitem{dfk94} B.J. Dalton, Z. Ficek, and P.L. Knight, Phys. Rev.
5579: A {\bf 50}, 2646 (1994).
5580:
5581: \bibitem{rich3} T. Richter, Optica Acta {\bf 31}, 1045 (1984).
5582:
5583: \bibitem{ft94} Z. Ficek and R. Tana\'s, Quantum Opt. {\bf 6}, 95 (1994).
5584:
5585: \bibitem{va92} G.V. Varada and G.S. Agarwal, Phys. Rev. A {\bf 45},
5586: 6721 (1992).
5587:
5588: \bibitem{bsp} A. Beige, C. Sch\"{o}n, and J. Pachos, Fortschr.
5589: Phys. {\bf 50}, 594 (2002).
5590:
5591: \bibitem{rich4} T. Richter, Optics Commun. {\bf 80}, 285 (1991).
5592:
5593: \bibitem{man65} A. Javan, E.A. Ballik, and W.L. Bond, J. Opt. Soc. Am.
5594: {\bf 52}, 96 (1962).
5595:
5596: \bibitem{man65a} M.S.~Lipsett and L.~Mandel, Nature {\bf 199}, 553
5597: (1963).
5598:
5599: \bibitem{mand1} R. Gosh and L. Mandel, Phys. Rev. Lett. {\bf 59}, 1903
5600: (1987).
5601:
5602: \bibitem{mand2} C.K. Hong, Z.Y. Ou, and L.~Mandel, Phys. Rev. Lett.
5603: {\bf 59}, 2044 (1987).
5604:
5605: \bibitem{mand3} Z.Y. Ou and L. Mandel, Phys. Rev. Lett. {\bf 62},
5606: 2941 (1989).
5607:
5608: \bibitem{rich90} Th. Richter, Phys. Rev. A {\bf 42}, 1817 (1990).
5609:
5610: \bibitem{sckd} M.O. Scully and K. Dr\"{u}hl, Phys. Rev. A {\bf 25},
5611: 2208 (1982).
5612:
5613: \bibitem{wong} T. Wong, S.M. Tan, M.J. Collett and D.F. Walls,
5614: Phys. Rev. A {\bf 55}, 1288 (1997).
5615:
5616: \bibitem{fr} T. Rudolph and Z. Ficek, Phys. Rev. A {\bf 58},
5617: 748 (1998).
5618:
5619: \bibitem{kochan} P. Kochan, H.J. Carmichael, P.R. Morrow and M.G. Raizen,
5620: Phys. Rev. Lett. {\bf 75}, 45 (1995).
5621:
5622: \bibitem{du} H.T. Dung and K. Ujihara, Phys. Rev. Lett. {\bf 84}, 254
5623: (2000).
5624:
5625: \bibitem{ftk88} Z. Ficek, R. Tana\'s, and S. Kielich, J. Mod. Opt.
5626: {\bf 35}, 81 (1988).
5627:
5628: \bibitem{sko} C. Skornia, J. von Zanthier, G.S. Agarwal, E. Werner,
5629: and H. Walther, Phys. Rev. A {\bf 64}, 063801 (2001).
5630:
5631: \bibitem{sko1} G.S. Agarwal, J. von Zanthier, C. Skornia, and H.
5632: Walther, Phys. Rev. A {\bf 65}, 053826 (2002).
5633:
5634: \bibitem{man64} L. Mandel, Phys. Rev. {\bf 134}, A10 (1964).
5635:
5636: \bibitem{be} A. Beige, S.F. Huelga, P.L. Knight, M.B. Plenio, and
5637: R.C. Thompson, J. Mod. Opt. {\bf 47}, 401 (2000).
5638:
5639: \bibitem{phbk} M.B. Plenio, S.F. Huelga, A. Beige, and P.L. Knight,
5640: Phys. Rev. A {\bf 59}, 2468 (1999).
5641:
5642: \bibitem{yzm} G.J. Yang, O. Zobay, and P. Meystre, Phys. Rev. A {\bf
5643: 59}, 4012 (1999).
5644:
5645: \bibitem{pb} S.J.D. Phoenix and S.M. Barnett, J. Mod. Opt. {\bf 40},
5646: 979 (1993).
5647:
5648: \bibitem{kk} I.K. Kudrayvtsev and P.L. Knight, J. Mod. Opt. {\bf 40},
5649: 1673 (1993).
5650:
5651: \bibitem{cz} J.I. Cirac and P. Zoller, Phys. Rev. A {\bf 50}, R2799
5652: (1994).
5653:
5654: \bibitem{ge} C.C. Gerry, Phys. Rev. A {\bf 53}, 2857 (1996).
5655:
5656: \bibitem{wan01} J. Wang, PhD Thesis, The University of Queensland,
5657: 2001.
5658:
5659: \bibitem{bmk} A. Beige, W.J. Munro, and P.L. Knight, Phys. Rev. A {\bf
5660: 62}, 052102 (2000); {\bf 64}, 049901 (2001).
5661:
5662: \bibitem{fhb} S. Franke, G. Huyet, and S.M. Barnett, J. Mod. Opt.
5663: {\bf 47}, 145 (2000).
5664:
5665: \bibitem{gy1} G.C. Guo and C.P. Yang, Physica {\bf A260}, 173 (1998).
5666:
5667: \bibitem{gy2} C.P. Yang and G.C. Guo, Physica {\bf A273}, 352 (1999).
5668:
5669: \bibitem{meyer} G.M. Meyer and G. Yeoman, Phys. Rev. Lett. {\bf 79}, 2650
5670: (1997).
5671:
5672: \bibitem{pk90} G.M. Palma and P.L. Knight, Phys. Rev. A {\bf 39},
5673: 1962 (1989).
5674:
5675: \bibitem{ap90} G.S. Agarwal and R.R. Puri, Phys. Rev. A {\bf 41},
5676: 3782 (1990).
5677:
5678: \bibitem{fw97} Z. Ficek and M.R.B. Wahiddin, Optics Commun. {\bf 134},
5679: 387 (1997).
5680:
5681: \bibitem{ft} Z. Ficek and R. Tana\'s, Optics Commun. {\bf 153}, 245
5682: (1998).
5683:
5684: \bibitem{koz} A.E. Kozhekin, K. Molmer, and E. Polzik, Phys. Rev. A
5685: {\bf 62}, 033809 (2000).
5686:
5687: \bibitem{hal1} J. Hald, J.L. Sorensen, L. Leich, and E.S. Polzik, Opt.
5688: Express {\bf 2}, 93 (1998).
5689:
5690: \bibitem{hal2} J. Hald, J.L. Sorensen, C. Schori, and E.S.
5691: Polzik, Phys. Rev. Lett. {\bf 83}, 1319 (1999).
5692:
5693: \bibitem{flei} M. Fleischhauer, S.F. Yelin, and M.D. Lukin, Optics
5694: Commun. {\bf 179}, 395 (2000).
5695:
5696: \bibitem{sm01} S. Schneider and G.J. Milburn, Phys. Rev. A {\bf 65},
5697: 042107 (2002).
5698:
5699: \bibitem{b1} A. Beige, S. Bose, D. Braun, S.F. Huelga, P.L. Knight,
5700: M.B. Plenio, and V. Verdal, J. Mod. Opt. {\bf 47}, 2583 (2000).
5701:
5702: \bibitem{b2} C.A. Sackett, D. Kielpinski, B.E. King, C. Langer,
5703: V. Meyer, C.J. Myatt, M. Rowe, Q.A. Turchette, W.M. Itano, D.J.
5704: Wineland, and I.C. Monroe, Nature {\bf 404}, 256 (2000).
5705:
5706: \bibitem{nat1} M.D. Lukin, M. Fleischhauer, R. Cote, L.M. Duan, D.
5707: Jaksch, J.I. Cirac, and P. Zoller, Phys. Rev. Lett. {\bf 87}, 037901
5708: (2001).
5709:
5710: \bibitem{nat2} M. Fleischhauer and C. Mewes, Lecture notes of the
5711: Enrico-Fermi School, Varenna, 2001.
5712:
5713: \bibitem{nat3} X. Wang and K. Molmer, Eur. Phys. J. D {\bf 18}, 385
5714: (2002).
5715:
5716: \bibitem{fl00} M. Fleischhauer and M.D. Lukin, Phys. Rev. A {\bf 65},
5717: 022314 (2002).
5718:
5719: \bibitem{fg02} M. Fleischhauer and S. Gong, Phys. Rev. Lett. {\bf 88},
5720: 070404 (2002).
5721:
5722:
5723: \end{thebibliography}
5724:
5725: \end{document}
5726: