cond-mat0409689/BFC.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: % Molecule -- Fermionic pair coupling in a Finite System
3: % Since May 9
4: % Last Update Sep 24
5: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6: \documentclass[aps,pra,twocolumn,showpacs,groupedaddress]{revtex4}
7: \usepackage{graphicx}
8: %----------------------------
9: \def\bra{\langle}
10: \def\ket{\rangle}
11: \def\pd{\partial}
12: \def\su{\uparrow}
13: \def\sd{\downarrow}
14: \def\bs{\bar{s}}
15: %----------------------------
16: \begin{document}
17: 
18: \title{Boson-Fermion coherence in a spherically symmetric harmonic trap}
19: \author{Takahiko Miyakawa and Pierre Meystre}
20: \affiliation{Optical Sciences Center,
21: The University of Arizona, Tucson, AZ 85721}
22: 
23: \date{\today}
24: 
25: \begin{abstract}
26: We consider the photoassociation of a low-density gas of
27: quantum-degenerate trapped fermionic atoms into bosonic molecules
28: in a spherically symmetric harmonic potential. For a dilute system
29: and the photoassociation coupling energy small
30: compared to the level separation of the trap, only those fermions
31: in the single shell with Fermi energy are coupled to the bosonic
32: molecular field.
33: Introducing a collective pseudo-spin operator formalism we show
34: that this system can then be mapped onto the Tavis-Cummings
35: Hamiltonian of quantum optics, with an additional pairing
36: interaction. By exact diagonalization of the Hamiltonian, we
37: examine the ground state and low excitations of the Bose-Fermi
38: system, and study the dynamics of the coherent coupling between
39: atoms and molecules. In a semiclassical description of the system,
40: the pairing interaction between fermions is shown to result in a
41: self-trapping transition in the photoassociation, with a sudden
42: suppression of the coherent oscillations between atoms and
43: molecules. We also show that the full quantum dynamics of the
44: system is dominated by quantum fluctuations in the vicinity of the
45: self-trapping solution.
46: \end{abstract}
47: 
48: \pacs{03.75.Lm, 03.75.Ss, 42.50.Ar} \maketitle
49: %
50: \section{Introduction}
51: 
52: The formation of ultracold diatomic molecules from Feshbach
53: resonances and photoassociation has witnessed spectacular
54: developments in recent years. Early demonstrations of molecule
55: formation using two-photon Raman photoassociation \cite{Wynar00}
56: and a Feshbach resonance \cite{Inouye98} were dominated by the
57: molecular losses due to processes such as inelastic decay to lower
58: energy molecular vibrational states \cite{Yurovsky99}, so that the
59: existence of the molecules could only be inferred from the
60: decrease in the number of atoms. The first unambiguous coherent
61: conversion of atoms into molecules was performed by Donley {\it et
62: al.} \cite{Donley02}, who exploited a Feshbach resonance in a
63: $^{85}$Rb Bose-Einstein Condensate (BEC). In subsequent experiments starting from an
64: atomic condensate of $^{87}$Rb, Rempe and coworkers used adiabatic
65: rapid passage to create the molecules \cite{Durr04}. Because the
66: molecules and atoms have different magnetic moments, they could be
67: spatially separated from each other using a magnetic field
68: gradient via the Stern-Gerlach effect. Similar work has been
69: conducted by Xu {\it et al.} \cite{Xu03}. Starting from a sodium
70: BEC, they used resonant laser light to blast away the remaining
71: atoms in the sample and isolated the molecules. Unfortunately, the
72: conversion efficiency was limited by inelastic losses very close
73: to the resonance so that molecular yields were $\lesssim 10\%$.
74: 
75: For fermionic atoms close to a Feshbach resonance the inelastic
76: collision rate for relaxation  to lower energy vibrational states
77: of the molecules scales like $a(B)^{-2.55}$ whereas for bosons it
78: scales like $a(B)$ where $a(B)$ is the scattering length near the
79: resonance \cite{Petrov04}. This is because close to resonance the
80: effective size of the molecules is of the order $a(B)$, which is
81: comparable to the interparticle spacing. In order for a molecule
82: to decay to a more deeply bound vibrational state with radius
83: $R_e\ll a(B)$, the atoms comprising the molecule along with an
84: additional atom must all collide within a distance $\sim R_e$.
85: Since two of the three atoms are necessarily identical, the
86: collision rate is suppressed for fermions. Taking advantage of
87: this consequence of the Pauli Exclusion Principle, Greiner {\it et
88: al.} \cite{Greiner03} achieved the first molecular BEC by starting from a quantum degenerate spin mixture of
89: $^{40}$K using adiabatic rapid passage through a Feshbach
90: resonance with a conversion efficiency of $\sim 80\%$. At about
91: the same time Zwierlein {\it et al.} \cite{Zwierlein03} and Jochim
92: {\it et al.} \cite{Jochim03} succeeded in producing a BEC of
93: $^{6}$Li$_2$ dimers by evaporatively cooling the atoms at a
94: constant magnetic field just below a resonance where $a(B)$ is
95: large and positive. Two recent experiments have led to the
96: observation of heteronuclear Feshbach resonances in Bose-Fermi
97: mixtures of $^6$Li and $^{23}$Na \cite{Stan04} in one case, and of
98: $^{87}$Rb and $^{40}$K in the other \cite{Inouye04}. We also
99: mention recent experiments by Kerman {\it et al.} \cite{Kerman04},
100: who produced metastable RbCs molecules in their lowest triplet
101: state starting from a laser-cooled mixture of $^{85}$Rb and
102: $^{133}$Cs by photoassociation. Major current experimental and
103: experimental efforts are directed towards the exploration of the
104: crossover between the Bardeen-Cooper-Schrieffer (BCS) state of
105: fermionic atoms and BEC of bosonic
106: molecules as the system crosses a Feshbach resonance
107: \cite{ex_BECBCS}. In the vicinity of these resonances, the system
108: enters a strongly interacting regime that offers a challenge for
109: many-body theories \cite{Timmermans,th_BECBCS}.
110: 
111: It is known that the pairing properties of finite-size systems can
112: be significantly different from those of the bulk material, due to
113: the discrete energy spectrum of the particles involved. The
114: detailed role of the shell structure has been explored both in the
115: nuclei~\cite{Mottelson}  and of superconductor
116: grains~\cite{SCgrain}, in which a collective character of pairs
117: plays an important role. In this paper we consider the
118: photoassociation of a dilute quantum-degenerate gas of fermionic
119: atoms trapped in a spherically symmetric harmonic trap into
120: molecular dimers, including the pairing interaction between
121: fermions. By introducing a pseudo-spin formalism for the time
122: reversal pairing operator~\cite{Anderson,Kerman61}, this problem
123: can be mapped onto an extension of the Tavis-Cummings
124: model~\cite{Tavis,Bogoliubov} that describes the coupling of $N$
125: two-level atoms and a single mode of the electromagnetic field,
126: and has found applications in the study of superradiance in
127: quantum optics~\cite{Dicke,Bonifacio}. This analogy allows us to
128: study in detail the ground state and lower excited states of the
129: system, as well as the coherent dynamics of atom-molecule coupling
130: in the trap. We show that for appropriate conditions, only those
131: fermions on the last energy shell of the trap participate in the
132: photoassociation process, and discuss the impact of the filling of
133: that shell on molecule formation. We find that depending on the
134: detuning $\delta$ of the photoassociation laser from the energy
135: difference between the molecules and atom pairs, the nature of the
136: ground state changes from being predominantly atomic to
137: predominantly molecular in nature. We study the crossover between
138: these two regions in detail, and quantify its property via the
139: joint coherence of the atomic and molecular fields and the
140: entanglement entropy of the system.
141: 
142: The coherent conversion of fermions into bosons has been studied
143: in the homogeneous case by several
144: authors~\cite{Timmermans,dy_BECBCS,Vardi}. Trapped systems, in
145: addition to being closer to the experimental situation, present
146: several unique characteristics: First, the discreteness of the
147: energy levels eliminates many of the difficulties associated with
148: a continuum. In addition, the high degeneracy of spherically
149: symmetric harmonic potentials simplifies significantly the study
150: of coherent quantum dynamics. Indeed, the problem resembles then
151: the dynamics of a bosonic Josephson Junction~\cite{BJJ1,BJJ2},
152: although the nonlinear coupling between fermionic atoms and
153: bosonic molecules leads to considerably richer dynamics. Moreover
154: the additional pairing interaction between fermions is shown to
155: result in a self-trapping transition~\cite{BJJ1,Scott}, with a
156: sudden suppression of the coherent oscillations.
157: 
158: This paper is organized as follows. Section II discusses our model
159: and formulates it in terms of a pseudo-spin formalism. Section III
160: presents results of the static problem where the many-body states
161: are classified by number of unpaired fermions known as {\it
162: seniority}~\cite{Racah} in nuclear physics. We show that for a
163: fixed number of atoms, the ground state of the system is always
164: the state of minimum seniority. We examine the crossover behavior
165: of the ground state as the detuning parameter $\delta$ is varied,
166: as a function of the ratio of the total number of fermionic pairs
167: and molecules to the degenerate number of the Fermi level. The
168: entanglement between fermions and bosons is evaluated and found to
169: be reduced as the pairing interaction becomes stronger. In section
170: IV, we analyze the coherent dynamics of the nonlinear
171: atom-molecule coupling. Using a semiclassical factorization
172: ansatz, we show the appearance of a self-trapping transition in
173: the presence of pairing interaction. An exact quantum solution
174: shows that around that transition point the dynamics is
175: characterized by large quantum fluctuations. Finally, section V is
176: a summary and outlook. Calculational details are relegated to an
177: appendix.
178: 
179: %
180: \section{Model}
181: 
182: We consider a trapped dilute gas of two-component fermionic atoms
183: in hyperfine states of spin $\sigma=\su,\sd$ at zero temperature
184: and coupled to a single-mode gas of bosonic molecules via a
185: two-photon Raman transition. The trap is assumed to be harmonic
186: and spherically symmetric, described by the potential
187:     \begin{equation}
188:     V_f=\frac{1}{2}m_f \omega_{\rm ho}^2 {\bf r}^2
189:     \end{equation}
190: for the atoms, and similarly with $f \rightarrow b$ for the
191: bosonic molecules.
192: 
193: In the absence of interactions between particles, the trap levels
194: have the energies
195:     \begin{equation}
196:     {\cal E}_{n}=\left(n+\frac{3}{2}\right)\hbar\omega_{\rm
197:     ho}.
198:     \end{equation}
199: where the principal quantum number $n$ is positive or zero. In
200: order to deal with the high degree of degeneracy of this
201: potential, it is convenient to introduce the (integer) radial and
202: angular quantum numbers $n_r$ and $l$, which are positive or zero,
203: with \cite{CT}
204:     \begin{equation}
205:     n= 2n_{r} + l.
206:     \end{equation}
207: Each pair $(n_r, l)$ corresponds to a radial wave function
208: $R_{n,l}(r)$ and hence $(2l+1)$ common eigenfunctions of $V_f(r)$,
209: ${\bf L}^2$ and $L_z$,
210:     \begin{equation}
211:     \phi_{n,l,m}({\bf r})= R_{n,l}(r)
212:     Y_{lm}(\theta,\phi).
213:     \end{equation}
214: Taking into account the magnetic quantum number
215:     \begin{equation}
216:     -l \le m \le +l,
217:     \end{equation}
218: the degeneracy of each level ${\cal E}_n$ is therefore
219:     \begin{equation}
220:     \Omega_n = \frac{1}{2} (n + 1)(n +2),
221:     \end{equation}
222: and the total number of states up to the shell $n_F$
223: corresponding to the Fermi energy
224:     $$
225:     {\cal E}_F = \left(n_F + \frac{3}{2}\right) \hbar \omega_{\rm ho}
226:     $$
227: is
228: \[
229:     N_{n_F}=2\times
230:     \sum_{n=0}^{n_F}\Omega_{n}=\frac{1}{3}(n_F+1)
231:     (n_F+2)(n_F+3),
232: \]
233: where the factor of 2 accounts for the two hyperfine spin states
234: of the atoms. In the following, it will be necessary to include
235: the attractive interaction responsible for the pairing between
236: fermions. Since as we show later on this interaction splits the
237: degeneracy of the trap levels, we keep the angular momentum index
238: in the labelling of the atomic energies, ${\cal E}_{n,l}$. The
239: fermionic atoms are therefore described by the annihilation
240: operator $c_{n;l,m,\sigma}$, where $\sigma$ labels the hyperfine
241: spin state of the atom with single-particle wave function
242: $\phi_{n,l,m}({\bf r})$ and eigenenergy ${\cal E}_{n,l}$.
243: 
244: Assuming that the atom-molecule photoassociation energy is smaller
245: than the trap energy spacing, $\hbar\omega_{\rm ho}$, and in
246: addition that the system is sufficiently dilute that the
247: attractive interaction between fermions is likewise less than
248: $\hbar\omega_{\rm ho}$, it is possible to tune the frequency of
249: the photoassociation laser so as to only couple fermions in the
250: shell $n_F$ with Fermi energy ${\cal E}_F$ to molecules in the ground
251: state of the harmonic trap. We can then ignore all shells other than the
252: $n_F$-shell for the fermions, and all trap states above the ground state
253: for the molecules, which are then described in terms of the ground
254: state bosonic annihilation operator $b$ with single particle
255: energy ${\cal E}_0$.
256: 
257: Both atomic pairing and photoassociation involve the creation and
258: annihilation of pairs of atoms, hence it is convenient to
259: introduce pseudo-spin operators $S_l$~\cite{Kerman61}
260: for atoms of angular momentum $l$ in the $n_F$-shell as
261: \begin{eqnarray}
262:     S_l^+&=&\sum^l_{m=-l}(-1)^{l-m} c_{lm\su}^\dagger
263:     c_{l-m\sd}^\dagger,\\
264:     S_l^-&=&\sum^l_{m=-l}(-1)^{l-m} c_{l-m\sd}
265:     c_{lm\su},\\
266:     S_l^z&=&\frac{1}{2}\left[\sum_{m,\sigma}c_{lm\sigma}^\dagger
267:     c_{lm\sigma}-\Omega_l\right]=\frac{1}{2}\left(\hat{n}_l-\Omega_l\right),
268:     \end{eqnarray}
269: where $\Omega_l = 2l+1$ and $\hat{n}_l$ is number operator in each
270: level $l$. (Here and hereafter we have omitted the label $n_F$
271: from the fermionic operator $c_{n_F;l,m,\sigma}$.) They are easily
272: seen to obey the SU(2) algebra
273:     \begin{equation}
274:     [S_l^+,S_{l'}^-]=2S_l^z \delta_{l,l'}, \,\,\,\,\,\,\,
275:     [S_l^z,S_{l'}^{\pm}]=\pm S_l^\pm \delta_{l,l'}.
276:     \end{equation}
277: 
278: Since we need only consider atoms in the Fermi level, the total
279: number of relevant particles in the system is
280:     \begin{equation}
281:     N= n_p + n_b + \nu \equiv M + \nu,
282:     \label{Ntotal}
283:     \end{equation}
284: where
285:     \begin{equation}
286:     M= n_p + n_b
287:     \end{equation}
288: is the number of molecules ($n_b$) and atomic pairs ($n_p$)
289: in the $n_F$-shell, or loosely speaking the number of pairs,
290: and $\nu$ is the number of unpaired atoms in the $n_F$-shell.
291: In that reduced Hilbert space, a complete set of
292: states is given by
293:     \begin{eqnarray}
294:     &&|n_{l},n_{l^\prime},\cdots,n_{l^{\prime\prime}},n_b;\nu\ket\nonumber\\
295:     &&\quad =\frac{1}{\sqrt{\mathcal N}}
296:     (S_l^+)^{n_l}(S_{l^\prime}^+)^{n_{l^\prime}}
297:     \cdots(S_{l^{\prime\prime}}^+)^{n_{l^{\prime\prime}}}
298:     (b^\dagger)^{n_b}|\nu\ket,
299:     \end{eqnarray}
300: where $\mathcal N$ is a normalization constant.
301: 
302: For a given angular momentum $l$, the possible number of atomic
303: pairs $n_l$ in the Fermi level is
304:     $$
305:     0\leq n_l \leq \Omega_l = 2l+1,
306:     $$
307: while the number of molecules is $0\leq n_b \leq N$.
308: 
309: The pseudo-spin operator $S_l^{-}$ annihilates atoms in pairs,
310: hence any state
311: $|\nu\ket\equiv|\nu_l,\nu_{l^\prime},\cdots,\nu_{l^{\prime\prime}}\ket$
312: of unpaired fermions and zero molecules clearly satisfies
313:     \begin{equation}
314:     S_l^{-}|\nu\ket=0,\,\,\,\,\,\, \hat{n}_b|\nu\ket=0,
315:     \label{nu-definition1}
316:     \end{equation}
317: with
318:     \begin{equation}
319:     \hat{n}_l|\nu\ket=\nu_l|\nu\ket,
320:     \label{nu-definition2}
321:     \end{equation}
322: see Eq. (\ref{Ntotal}). The number operator of bosonic molecules has been
323: defined by $\hat{n}_b=b^\dagger b$. $\nu_l$ is the number of unpaired fermions of
324: angular momentum $l$ in each level and is
325: referred to as {\it seniority}~\cite{Racah} in nuclear physics.
326: 
327: With this formal development at hand, we now turn to the
328: discussion of the fermionic pairing and of the photoassociation of
329: atoms into molecules. It is described by the effective Hamiltonian
330: \begin{equation}
331: \label{modelH} H=(\hbar \delta+{\cal E}_0) b^\dagger
332: b+\sum_{l,m,\sigma}{\cal E}_{n_F,l} c_{lm\sigma}^\dagger
333: c_{lm\sigma} +V_{p} +V_{am},
334: \end{equation}
335: where $\delta$ is the two-photon detuning between the Raman lasers
336: and the internal energy difference between atomic pairs and
337: molecules, $V_p$ describes atomic pairing and $V_{am}$ accounts
338: for the photoassociation of atoms into molecules. Before
339: discussing these two interaction Hamiltonians in detail, we first
340: evaluate the mean-field lifting of the single-particle energy
341: degeneracy, ${\cal E}_{n_F} \rightarrow {\cal E}_{n_F,l}$.
342: 
343: In the $s$-wave scattering approximation, valid at $T=0$, atoms of
344: opposite spin interact via the two-body interaction potential
345:     \begin{equation}
346:     V({\bf r}_1 - {\bf r}_2) = \frac{4 \pi \hbar^2 a}{m_f}
347:     \delta({\bf r}_1 - {\bf r}_2),
348:     \end{equation}
349: where $a <0$ is the scattering length, negative for attractive
350: interactions. In the Thomas-Fermi limit, this results in the atoms
351: being subjected to the mean-field potential
352: \begin{equation}
353: V(r)=\frac{2\pi\hbar^2a}{m_f}\rho(r),
354: \end{equation}
355: where the density $\rho(r)$ is given by
356: \begin{equation}
357: \rho(r)\simeq \rho_0(1-r^2/R^2_{TF})^{3/2},
358: \end{equation}
359: for
360:     $$
361:     r\leq R_{\rm TF}=a_{\rm ho}\sqrt{2n_F+3}
362:     $$
363: and is zero otherwise. Here $a_{\rm ho}=\sqrt{\hbar/m_f
364: \omega_{\rm ho}}$ is the oscillator length and
365: $\rho_0=(2n_F+3)^{3/2}/3\pi^2a_{\rm ho}^3$. The resulting mean-field
366: energy splitting of the $l$-states within the $n_F$ manifold is
367: then \cite{Heiselberg}
368: \begin{eqnarray}
369: \label{MFshift}
370: {\cal E}_{n_F,l}&-&{\cal E}_{n_F}\nonumber\\&=&\int\,dr\, r^2V(r)|R_{n_F,l}|^2
371: \\
372: &\simeq&\frac{2}{3\pi}\frac{a}{a_{\rm
373: ho}}(2n_F+3)^{3/2}\hbar\omega_{\rm ho}\left[
374: \frac{4}{3\pi}-\frac{1}{4\pi}\frac{l(l+1)}{n_F^2}\right]\nonumber,
375: \end{eqnarray}
376: where we have used the WKB limit of the radial harmonic oscillator
377: wave function which is valid for $n_F\gg 1$.
378: For an atomic system to be dilute, the mean-field shift Eq.(\ref{MFshift})
379: should be less than the unperturbed energy ${\cal E}_{n}$. This
380: implies that
381: \begin{equation}
382: n_F^{1/2}\frac{|a|}{a_{\rm ho}}\ll 1,
383: \end{equation}
384: which is equivalent to the familiar diluteness condition
385: $\rho_0|a|^3\ll 1$.
386: 
387: It is known that for attractive short range interactions, the potential
388: $V({\bf r}_1-{\bf r}_2)$ favors the creation of a time reversal
389: state and lets pairing take place. The Hamiltonian $V_p$ describes
390: this pairing correlation by including only the correlations for time reversal
391: pair states
392:     \begin{equation}
393:     |L=0,M=0;ll\ket=\sum_{m=-l}^l (l m ,l -m|0 0)|l,m\ket|l,-m\ket,
394:     \end{equation}
395: where
396:     \begin{equation}
397:     (l m,l -m|0 0)=(-1)^{l-m}(2l+1)^{-1/2}
398:     \label{CB-coeff}
399:     \end{equation}
400: is a Clebsch-Gordan coefficient. Assuming for simplicity that the
401: radial part $g$ of the pairing interaction is independent of the
402: angular momenta $l$ and $l'$ of the atomic pairs involved in the
403: interaction, $V_p$ reads explicitly
404:     \begin{equation}
405:     \label{pinteraction}
406:     V_{p}=-g\sum_{l,m}\sum_{l^\prime,m^\prime}
407:     (-1)^{l-m+l^\prime-m^\prime}c_{lm\su}^\dagger c_{l-m\sd}^\dagger
408:     c_{l^\prime m^\prime\sd}c_{l^\prime
409:     -m^\prime\su},
410:     \end{equation}
411: where the terms $(-1)^{l-m}$ and $(-1)^{l^\prime-m^\prime}$ come
412: from the Clebsch-Gordan coefficients (\ref{CB-coeff}) and hence
413: account for the angular part of the wave functions.
414: Strictly speaking, the coupling constant $g_{l,l'}$ should be
415: determined by the spatial integral
416: \begin{equation}
417: g_{l,l^\prime}=\frac{4\pi\hbar^2 |a|}{m_f}\int dr\,r^2\,
418: R_{n_F,l}^2(r)R_{n_F,l^\prime}^2(r). \label{g-spatial}
419: \end{equation}
420: However for our present purpose it is sufficient to estimate $g$ by
421: replacing the spatial integral in Eq. (\ref{g-spatial}) by $\bra
422: r_{n_F}^2 \ket^{-3/2}$, where for a pure harmonic oscillator state
423: in the $n_F$-shell the mean square radius $\bra r_{n_F}^2 \ket$ is
424: given by $(2n_F+3)a_{\rm ho}^2$ from the virial theorem. This
425: gives
426: \begin{equation}
427: g\sim \frac{4\pi\hbar^2 |a|}{m_f} (2n_F+3)^{-3/2}a_{\rm ho}^{-3}
428: \sim \frac{|a|}{a_{\rm ho}}n_F^{-3/2}\hbar\omega_{\rm ho}.
429: \end{equation}
430: 
431: In this model, the strongest pairing occurs for degenerate
432: energies, ${\cal E}_{l}={\cal E}_{l^\prime}$ for all $l,l^\prime$.
433: We will see that all fermions in the $n_F$-shell are then
434: coherently paired, the pairing energy being proportional to the
435: degeneracy factor $\Omega_{n_F}$. In the dilute gas limit, which
436: is equivalent to $g\Omega_{n_F}\sim g n_F^2 \ll \hbar\omega_{\rm
437: ho}$, the pairing takes then place on a single shell, see
438: Ref.~\cite{Heiselberg} for more details.
439: 
440: Turning finally to the photoassociation of atomic pairs into
441: molecules and the reverse process of photodissociation of
442: molecules back into atoms, we note that it is possible to neglect
443: all processes involving atoms other than those in the $n_F$-shell
444: provided that their characteristic frequency, the product $\chi
445: \sqrt{ \langle n_b \rangle}$ of the photoassociation coupling
446: constant, $\chi$, and the square root of the mean number of
447: molecules, remains small compared to the frequency separation
448: between neighboring shells of the trap. This condition is well
449: fulfilled for typical laser strengths and trap depths, in which
450: case $\chi \sqrt{ \langle n_b \rangle} \simeq 10^2 n_F^{-1/2}
451: \,{\rm s}^{-1} \ll \omega_{\rm ho} \simeq 10^3 \,{\rm s}^{-1}$.
452: In this case, the atom-molecule coupling can be approximated by the Hamiltonian
453: \begin{equation}
454: \label{aminteraction}
455: V_{am}=\sum_{l,m}(-1)^{l-m}\left[\chi
456: c_{lm\su}^\dagger c_{l-m\sd}^\dagger b +\chi^* b^\dagger
457: c_{l-m\sd} c_{lm\su}\right],
458: \end{equation}
459: where we only include the coupling between time-reversal atomic
460: pair and bosonic molecules in the ground state of the harmonic
461: trap. The photoassociation coupling constant $\chi$ is
462: proportional to the far off-resonant two-photon Rabi frequency
463: associated with two nearly co-propagating lasers with frequencies
464: $\omega_1$ and $\omega_2$~\cite{Wynar00},
465: \begin{eqnarray}
466: \label{coupling} \chi&=&\frac{\chi_0}{\sqrt{4\pi}} \int\, dr\,
467: r^2\, R^{(b)}_{0,0}(r) R_{n_F,l}^2(r)
468: \nonumber\\
469: &\sim& \frac{\chi_0}{n_F^{3/2} a_{\rm ho}^{3/2}},
470: \end{eqnarray}
471: where $R^{(b)}_{0,0}$ is a ground-state wave function of molecules
472: of mass of $2m_f$ with a spatial width about $a_{ho}$,
473: and in the second equality we
474: have again replaced the radial wave function of the atoms by 
475: $\bra r^2_{n_F} \ket^{-3/2}$.
476: From the coupling constant between $^{87}$Rb$_2$ molecules and $^{87}$Rb
477: atoms of Ref.~\cite{superchemistry} we estimate the
478: photoassociation coupling constant to be of the order of
479: $\chi_0=7.6 \times 10^{-7}$ m$^{3/2}$s$^{-1}$.
480: 
481: In terms of the pseudo-spin operators, and with Eqs.
482: (\ref{pinteraction}) and (\ref{aminteraction}), the total model
483: Hamiltonian Eq.~(\ref{modelH}) finally reads
484: \begin{eqnarray}
485: \label{spinH} H&=&(\hbar \delta+{\cal E}_0) b^\dagger b+\sum_l
486: 2{\cal E}_{n_F,l}(S_l^z+\Omega_l/2)\nonumber\\
487: &+&\chi\sum_l (S_l^+ b+b^\dagger
488: S_l^-)-g\sum_{l,l^\prime}S_l^+S_{l^\prime}^-.
489: \end{eqnarray}
490: This Hamiltonian clearly conserves the spin operators ${\bf
491: S}_l^2$. Applying this operator on the state $|\nu\rangle$ we have
492: \begin{eqnarray*}
493: {\bf S}_l^2|\nu\ket&=&S_l(S_l+1)|\nu\ket \\
494: &=& \{S_l^+S_l^-+S_l^z(S_l^z-1)\}|\nu\ket \\
495: &=& (\Omega_l/2-\nu_l/2)(\Omega_l/2-\nu_l/2+1)|\nu\ket,
496: \end{eqnarray*}
497: where we have used the identity ${\bf
498: S}_l^2=S_l^{+}S_l^{-}+S_l^{z}(S_l^{z}-1)$ and Eqs.
499: (\ref{nu-definition1}) and (\ref{nu-definition2}), so that
500:     $$
501:     S_l=\frac{1}{2}(\Omega_l-\nu_l).
502:     $$
503: This allows us to identify the operator
504:     \begin{equation}
505:     \hat{M}=\sum_l \left(S^z_l+S_l\right)+\hat{n}_b,
506:     \end{equation}
507: as the operator giving the total number of fermionic pairs and
508: molecules, or loosely speaking the ``pair number'' operator, which
509: is easily seen to also be a conserved quantity.
510: 
511: In the limit $g \to 0$ the Hamiltonian~(\ref{spinH}) reduces to
512: the Tavis-Cummings Model~\cite{Tavis,Bogoliubov} of quantum
513: optics, while for $\chi \to 0$, it becomes the Pairing Model
514: ~\cite{Richardson,Dukelsky}, for which Richardson first gave an
515: exact solution in the context of nuclear physics. We also mention
516: a recent family of exactly solvable models of atom-molecule
517: proposed in Ref.~\cite{AMHamiltonian}.
518: 
519: The discussion of the following sections concentrates specifically
520: in that situation where the mean-field energy shift in
521: single-particle energies is smaller than the photoassociation
522: coupling. In this degenerate model, the single-particle energies
523: of all atoms in the $n_F$-shell can then be taken to be equal,
524: ${\cal E}_{n_F,l}= {\cal E}_F$, and the Hamiltonian (\ref{spinH})
525: simplifies to
526:     \begin{equation}
527:     \label{DGmodel}
528:     H=-\hbar \omega\left(S^z+S\right)
529:     +\hbar\chi\left(S^{+}b+b^\dagger S^{-}\right)-\hbar gS^{+}S^{-},
530:     \end{equation}
531: where we have introduced the total spin operator
532:     \begin{equation}
533:     {\bf S}=\sum_l {\bf S}_l,
534:     \end{equation}
535: $\omega=\delta+({\cal E}_0-2{\cal E}_F)/\hbar$, $\Omega_{n_F}=\sum_l \Omega_l$, $\nu=\sum_l \nu_l$, so that the total spin is
536:     \begin{equation}
537:     S= \sum_l S_l=\frac{1}{2}\left (\Omega_{n_F} - \nu \right ).
538:     \label{total spin}
539:     \end{equation}
540: In Eq.~(\ref{DGmodel}), we have neglected constant terms
541: proportional to the conserved quantity $M$.
542: 
543: Assuming an oscillator length $a_{\rm ho} =3.2\times 10^{-6}$ m
544: for the mass of $^6$Li, $a=-114$ nm for its scattering length, and
545: $\omega_{\rm ho}=1000$ s$^{-1}$, then $\chi n_F^{3/2} \simeq 10^{2}$
546: s$^{-1}$.  The validity of the degenerate model, $|{\cal
547: E}_{n_F,n_F}- {\cal E}_{n_F,l=0}| < \chi n_F$ requires $n_F
548: \lesssim 10$, which corresponds to a total number of fermions 
549: $N_{n_F} \lesssim 10^3$.
550: 
551: \section{Ground State}
552: 
553: In this section, we discuss the dependence of the ground state of
554: the model Hamiltonian (\ref{DGmodel}) on the ratio of number of
555: paired fermions and molecules, $M=n_p+n_b,$ to the degeneracy
556: $\Omega$ of the Fermi level (we drop the subscript `$n_F$' for
557: notational clarity from now on). We identify qualitatively
558: different types of ground states, a pair-dominated ground state
559: and a molecular-dominated one, as a function of the parameter
560:     \begin{equation}
561:     \kappa \equiv \frac{\omega}{\chi \sqrt{\Omega}},
562:     \end{equation}
563: where $\omega = \delta +({\cal E}_0-2 {\cal E}_F)/\hbar$ is the
564: photoassociation frequency detuning.
565: 
566: \subsection{Pairing Model}
567: 
568: For a total number of particles $N=2M+\nu \leq \Omega$ , which
569: corresponds to the Fermi energy shell less than half-filled since
570: there are two hyperfine atomic states involved, the seniority
571: $\nu$ can take the values
572: \begin{equation}
573:   \nu = \cases{0,2,4,\dots,N & ({\em N} \, even) \cr
574:              1,3,5,\dots,N & ({\em N} \, odd), \cr}
575: \end{equation}
576: while for $\Omega < N \leq 2\Omega$, corresponding to a shell more
577: than half filled, the permissible values of $\nu$ are
578: \begin{equation}
579:   \nu = \cases{0,2,4,\dots,2\Omega-N & ({\em N} \, even) \cr
580:              1,3,5,\dots,2\Omega-N & ({\em N} \, odd). \cr}
581: \end{equation}
582: 
583: The degenerate Pairing Model of Refs.~\cite{Schuck,Walecka} is
584: obtained by neglecting the photoassociation coupling, $\chi=0$, in
585: the Hamiltonian~(\ref{DGmodel}). In this case, the total energy is
586: given as a function of seniority $\nu$ and the total number $N$ of
587: fermions by
588: \begin{equation}
589: E_{\rm pm}(N,\nu)=-\omega N -\frac{g}{4}(N-\nu)(2\Omega-\nu-N+2),
590: \end{equation}
591: and it is minimized for $\nu=0$ or $1$.
592: 
593: For an even particle number, the ground state corresponds
594: therefore to all pseudo-spins aligned, $S = \Omega/2$,
595:     \begin{equation}
596:     E_{\rm pm}(N,\nu=0)=-\omega N -gN(2\Omega -N + 2).
597:     \end{equation}
598: The first excited state corresponds to $\nu=2$ unpaired atoms, or
599: $S=\Omega/2-1$, and its energy is
600: \begin{equation}
601: \label{gapenergy} E_{\rm pm}(N,\nu=2)-E_{\rm pm}(N,\nu=0)=g\Omega.
602: \end{equation}
603: Thus, the energy needed to break up an atomic pair into two
604: un-paired fermions is independent of the number of fermions in the
605: $n_F$-shell. This energy is consistent to Bogoliubov
606: quasi-particle energy based on BCS variational state except for
607: corrections of relative order $1/\Omega$.
608: 
609: \subsection{Photoassociation}
610: 
611: In the presence of photoassociation, $\chi \neq 0$, the
612: eigenstates of the system consist of a coherent superposition of
613: atoms and molecules. For fixed $N$, we can classify them by the
614: total spin $S=(\Omega-\nu)/2$, each manifold consisting of $(M+1)$
615: eigenstates, where $M=(N-\nu)/2 = n_p + n_b$ is the total number
616: of pairs.
617: 
618: For positive detuning, $\omega>0$, the ground state is a pure
619: fermionic state, i.e., $N=2n_p+\nu$. On the other hand, for a
620: large negative detuning energy, $\omega<0$, and even particle
621: number, the ground state is reached when all particles are
622: molecules, corresponding to a zero seniority, or maximum spin
623: state.
624: 
625: Figure~\ref{gsfig1} shows the energies $E_\nu$ of a few seniority
626: states relative to that of the state $\nu=0$ as a function of
627: $\kappa = \omega/\chi\sqrt{\Omega}$. This example is for $\eta=
628: g\sqrt{\Omega}/\chi=1$ and a half-filled Fermi energy shell,
629: $N=\Omega=120$, corresponding to $n_F=14$, and results from a
630: direct numerical diagonalization of the Hamiltonian. It shows that
631: over the wide range of $\kappa$ considered here, the ground state
632: is always the maximum spin manifold $\nu=0$. For large negative
633: detunings, the energy differences $E_\nu-E_0$ are $\nu/2$ times
634: the single-particle energy difference between a molecule and two
635: fermions, and increase with $|\omega|$. For large positive
636: detunings, on the other hand, the energies of the excited states
637: approach the values $\nu/2$ times $g\Omega$, see
638: Eq.~(\ref{gapenergy}). In the crossover region where the nature of
639: the ground state changes from atomic to molecular, the ground
640: state is a coherent superposition of atoms and molecules that is
641: likewise the maximum spin state. We have verified that in the
642: absence of pairing interaction, $g=0$, the ground state is
643: likewise the state of maximum spin (not shown in figure).
644: 
645: \begin{figure}
646: %  \begin{minipage}{0.45\textwidth}
647:     \begin{center}
648:       \includegraphics[width=7.5cm,clip]{gsfig1.eps}
649:     \end{center}
650:     \caption{Excitation energies, in units of $\chi \sqrt{\Omega}$,
651:      relative to the ground-state energy
652:      for the seniority states
653:      $\nu=2$ (solid line), $\nu=4$ (dashed line), and $\nu=6$
654:      (dash-dotted line), as a function of the dimensionless
655:      parameter $\kappa=\omega/\chi\sqrt{\Omega}$
656:      for $\eta=g\sqrt{\Omega}/\chi=1$ and $N=\Omega=120$.}
657:     \label{gsfig1}
658: \end{figure}
659: %
660: 
661: In the following we concentrate on the state of maximum total
662: spin, $|S=\Omega/2\ket$, for an even number of fermions, $N=2M$.
663: The eigenstates of the atom-molecule system in the spin manifold
664: $S=\Omega/2$ have the general form
665: \begin{eqnarray}
666: |\Phi^\lambda_S\ket&=&\sum_{n_b=0}^{M} C_{\lambda}(n_p)
667: |S=\Omega/2,S^z=-S+n_p\ket_f \nonumber\\&&\otimes |n_b=M-n_p\ket_b,
668: \end{eqnarray}
669: where $n_p$ denotes a number of fermionic pairs and
670: $\lambda=0,1,\cdots M$ represent eigenmodes of the system with
671: eigenenergies $E_\lambda$. We have found these eigenstates and the
672: associated eigenenergies by direct diagonalization of the
673: Hamiltonian~(\ref{DGmodel}).
674: 
675: \begin{figure}
676:   \begin{minipage}{0.45\textwidth}
677:     \begin{center}
678:       \includegraphics[width=7.5cm,clip]{gsfig2.eps}
679:     \end{center}
680:     \caption{Normalized average number of molecules in the ground state
681:     as a function of $\kappa$ and $M$ for $\Omega=120$ and $g=0$.}
682:     \label{gsfig2}
683:   \end{minipage}
684: \hspace{4mm}
685: %
686:   \begin{minipage}{0.45\textwidth}
687:     \begin{center}
688:       \includegraphics[width=7.5cm,clip]{gsfig3.eps}
689:     \end{center}
690:     \caption{Normalized joint coherence function, $|G_{am}|$,
691:       as functions of $\kappa$ and $M$ for $\Omega=120$ and for $g=0$.}
692:     \label{gsfig3}
693:   \end{minipage}
694: \end{figure}
695: 
696: The average number of molecules in the ground state is shown in
697: Fig.~\ref{gsfig2} as a function of
698: $\kappa=\omega/\chi\sqrt{\Omega}$ and $M$ for $g=0$ and
699: $\Omega=120$. As expected, for large positive detuning $\omega>0$,
700: the ground state population consists almost exclusively of atoms,
701: while it is mostly made up of molecules for large negative
702: detunings. In the crossover region around $\omega=0$, the ground
703: state consists of a coherent superposition state between molecules
704: and atomic pairs.
705: 
706: It is possible to characterize this superposition in terms of the
707: normalized joint coherence function
708:     \begin{equation}
709:     G_{am} = \frac{2 \langle S^+ b\rangle}{M\sqrt{\Omega}},
710:     \end{equation}
711: which is shown in Fig.~\ref{gsfig3} as a function of $\kappa$ and
712: the number of pairs $M$. The joint coherence of the atomic and
713: molecular fields shows a remarkable enhancement in the crossover
714: region as well as a change in shape as a function of $M$, due to
715: the nonlinearity of the atom-molecule
716: coupling~(\ref{aminteraction}). This dependence on the filling
717: factor $M/\Omega$ can be understood more quantitatively by
718: considering the two limiting cases $M/\Omega\ll 1$ and $M/\Omega
719: \simeq 1$.
720: 
721: \subsubsection{$M/\Omega \ll 1$ --- mapping on a linear coupled-boson
722: system}
723: 
724: In the limit of small filling factors, it is convenient to
725: describe the system in terms of the Holstein-Primakoff mapping
726: ~\cite{HP40} of the $SU(2)$ generators
727: $S^+=(S^{-})^\dagger$ and $S^{z}$ in terms of bosonic operators.
728: According to this mapping, the Hilbert space of the group $SU(2)$
729: is carried by a subspace of the bosonic Fock space given by the
730: bosonic vacuum $|0\rangle_d$ and the bosonic operators $d$ and
731: $d^\dagger$, with
732:     \begin{equation}
733:     [d,d^\dagger]=1, \qquad d|0\ket_d=0,
734:     \end{equation}
735: The bosonic space being spanned by the $n_d-$boson states
736:     \[
737:     |n_d\ket_d=\frac{1}{\sqrt{n_d !}}(d^\dagger)^{n_d}|0\ket_d \qquad
738:     {\rm for}\,\, n_d=0,1,2,\dots,M.
739:     \]
740: Since we restrict our considerations to the subspace characterized
741: by the angular momentum quantum number $S=\Omega/2$, we can map
742: the operators $S^{\pm}$ and $S^{z}$ as
743: \begin{eqnarray}
744:     \label{HPmapping}
745:     S^{+} &\to& \sqrt{\Omega} d^\dagger
746:     \sqrt{1-\frac{d^\dagger d}{\Omega}},\nonumber \\
747:     S^{-} &\to& \sqrt{\Omega} \sqrt{1-\frac{d^\dagger d}{\Omega}} d,\nonumber \\
748:     S^z &\to& -\frac{\Omega}{2}+d^\dagger d.
749:     \end{eqnarray}
750: 
751: In the limit $M/\Omega\ll 1$, we only need a lowest-order of the
752: operators (\ref{HPmapping}), that is, replace the square roots by
753: 1. The Hamiltonian~(\ref{DGmodel}) reduces then to that of a
754: linearly coupled two-mode boson system,
755: \begin{equation}
756:     \label{linearboson}
757:     H\to H_{\rm linear}=-(\omega+g\Omega)d^\dagger
758:     d +\chi\sqrt{\Omega}(d^\dagger b +b^\dagger d),
759:     \end{equation}
760: with the constraint $M=n_d+n_b$.
761: 
762: This Hamiltonian can be diagonalized by the transformation
763:     \begin{eqnarray}
764:     c_{-}^\dagger&=&\cos{\theta}d^\dagger-\sin{\theta}b^\dagger\nonumber \\
765:     c_{+}^\dagger&=&\sin{\theta}d^\dagger+\cos{\theta}b^\dagger \nonumber,
766:     \end{eqnarray}
767: with $\cot{2\theta}=(\omega+g\Omega)/2\chi\sqrt{\Omega}$,
768: ,$0\le\theta\le 2\pi$, to give
769: \begin{equation}
770: \label{lineardiagonal} H_{\rm linear}=\epsilon_{-}c_-^\dagger
771: c_-+\epsilon_{+}c_+^\dagger c_+,
772: \end{equation}
773: with energies
774:     \begin{eqnarray}
775:     \epsilon_-&=&-(\omega+g\Omega)\cos^2{\theta}
776:     -2\chi\sqrt{\Omega}\cos{\theta}\sin{\theta}, \nonumber \\
777:     \epsilon_+&=&-(\omega+g\Omega)\sin^2{\theta}
778:     +2\chi\sqrt{\Omega}\cos{\theta}\sin{\theta}. \nonumber
779:     \end{eqnarray}
780: Since $\epsilon_-<\epsilon_+$ for $\chi > 0$, the ground state is
781: $(1/\sqrt{M !})(c_-^\dagger)^M|0\ket_d|0\ket_b$.
782: 
783: \subsubsection {$M/\Omega \simeq 1$ --- Mapping on a binary
784: atomic-molecular condensate}
785: 
786: Tikhonenkov and Vardi~\cite{Vardi} showed for in the homogeneous
787: case that when the total number of pairs is equal to the available
788: momentum states of the fermions, the system of fermionic atoms and
789: bosonic molecules can be mapped onto a two-mode atomic-molecular
790: BEC system~\cite{AMBEC1,AMBEC2,Zhou}. The corresponding situation
791: in our case occurs for $M=\Omega$, with an additional
792: atom-molecule two-body collision term required in addition.
793: 
794: To show how this works we introduce the two-mode Hamiltonian of a
795: two-component condensate of atoms and molecules,
796: \begin{equation}
797: \label{amhamiltonian}
798: H_{am}=-\omega b_m^\dagger b_m + \frac{\chi}{2}(b_m^\dagger b_a b_a+h.c.)
799: -\frac{g}{2}b^\dagger_m b_m b^\dagger_a b_a,
800: \end{equation}
801: where $b_a^\dagger$ and $b_m^\dagger$ are the bosonic creation
802: operators for the atomic and molecular modes, respectively.
803: Clearly the total number of particles $N_{am}=n_a+2n_m$ is
804: conserved, where $n_a$ and $n_m$ are the number of atoms and
805: molecules, respectively. For $N_{am}$ even, a general state of the
806: system can be expressed as
807: \begin{eqnarray}
808: |\Phi_{am}\ket&=&\sum_{m_a=0}^{M}C(m_a)|n_a=2m_a\ket\otimes
809: |n_m=M-m_a\ket \nonumber \\
810: &=& \sum_{m_a=0}^{M}C(m_a)|2m_a,n_m \ket,
811: \end{eqnarray}
812: with $M=N_{am}/2$. In this representation, the matrix form of the
813: Hamiltonian~(\ref{amhamiltonian}) is
814:      \begin{eqnarray*}
815:        \label{amMH}
816:        &&\bra 2m_a;n_m |H_{am}| 2m_a;n_m\ket\\
817:        &&\qquad =-\omega(M-m_a)-gm_a(M-m_a)\\
818:        &&\bra 2(m_a-1);n_m+1 |H_{am}| 2m_a;n_m\ket\\
819:        &&\qquad=\bra2m_a;n_m|H_{am}|2(m_a-1);n_m+1\ket \\
820:        &&\qquad=\chi \sqrt{m_a(m_a-1/2)(M-m_a)}.
821:      \end{eqnarray*}
822: Similarly, for our model the matrix form of the
823: Hamiltonian~(\ref{DGmodel}) in the ``pair number'' representation
824: is
825:     \begin{eqnarray*}
826:       \label{fbMH}
827:       &&\bra n_b;n_p|H|n_b;n_p\ket\\
828:       &&\qquad =-\omega(M-n_b)-g (n_b+\Delta M+1)(M-n_p)\\
829:       &&\bra n_b-1;n_p+1 |H| n_p;n_b\ket\\
830:       &&\qquad =\bra n_p;n_b |H|n_b-1;n_p+1\ket\\
831:       &&\qquad =\chi \sqrt{n_b(n_b+\Delta M+1)(M-n_b)}.
832:     \end{eqnarray*}
833: where
834:     \begin{equation}
835:     \Delta M=\Omega-M ,
836:     \end{equation}
837: with $\Delta M>0$. In the limit $M\simeq \Omega$, that is, when we
838: can neglect $\Delta M/M \ll 1$ and other terms of order of
839: $1/\Omega$, these two Hamiltonians are same under the
840: transformations $m_a\leftrightarrow n_b$ and $n_m \leftrightarrow
841: n_p$.
842: 
843: \subsubsection{Intermediate regime}
844: 
845: The behavior of the system apart from these two limiting cases
846: deviates from both models. The most striking difference between
847: these regimes appears in the energy, $\Delta$, of the first
848: excited state. Figure~\ref{gsfig4} shows
849: $\Delta/\chi\sqrt{\Omega}$ in the absence of pairing interaction
850: for $M/\Omega=1$ (solid line), $M/\Omega=0.5$ (dashed line), and
851: $M/\Omega=0.1$ (dash-dotted line) as a function of $\kappa =
852: \omega/\chi\sqrt{\Omega}$ for fixed $M=120$. In the case of
853: $M/\Omega=0.1$, $\Delta$ agrees well with the one-particle energy
854: difference $\epsilon_+-\epsilon_-=\sqrt{\omega^2+ 4\chi^2\Omega}$
855: of the linear coupled-boson model~(\ref{lineardiagonal}) (shown as
856: asterisks in the figure). Increasing $M/\Omega$ shifts the
857: location of the minimum of the energy gap and reduces the value of
858: its minimum. For $M=\Omega$, finally, the minimum gap approaches
859: zero at $\kappa\simeq 2$, consistently with a transition point in
860: the atom-molecule condensate system~\cite{AMBEC1,Zhou}. This
861: result is indicative of the appearance of a quantum phase
862: transition in the limit $M \to \infty$.
863: 
864: \begin{figure}
865:     \begin{center}
866:       \includegraphics[width=7.5cm,clip]{gsfig4.eps}
867:     \end{center}
868:     \caption{Lowest excitation energies as a function of
869:       $\kappa$ for $M/\Omega=1$ (solid line)
870:       and $M/\Omega=0.5$ (dashed line), and $M/\Omega=0.1$ (dash-dotted line),
871:       at fixed pair number $M=120$, where $g=0$. The asterisks correspond
872:       to a linear approximation for small $M/\Omega$ (see text).}
873:     \label{gsfig4}
874: \end{figure}
875: 
876: \subsection{The role of the pairing interaction}
877: 
878: We now examine in more detail the ground-state statistics of the
879: molecular field in the presence of pairing interaction $V_{p}$.
880: Since the cases of $M\ll \Omega$ and $M\simeq \Omega$ can be
881: mapped onto relatively well-known systems, we present results for
882: the situation of a half-filled shell, $M/\Omega=0.5$, only. Figures
883: \ref{gsfig5} and \ref{gsfig6} show the probability $P(n_b)$ of
884: having $n_b$ molecules in the trap, the molecule statistics, for
885: $2M=\Omega=120$ as a function of the dimensionless parameter
886: $\kappa$, with $g=0$ and $\eta=g\sqrt{\Omega}/\chi=10$,
887: respectively. There are several quantitative differences between
888: the two cases: the transition from an atomic to a molecular ground
889: state is shifted by the pairing interaction, and the width of the
890: crossover region in detuning space is significantly broader.
891: \begin{figure}
892:   \begin{minipage}{0.45\textwidth}
893:     \begin{center}
894:       \includegraphics[width=7.5cm,clip]{gsfig5.eps}
895:     \end{center}
896:     \caption{Molecule statistics $P(n_b)$ as a function of
897:       the dimensionless parameter $\kappa$ in the absence of
898:       pairing interaction, $g=0$, for $2M=\Omega=120$.}
899:     \label{gsfig5}
900:   \end{minipage}
901: \hspace{4mm}
902: %
903:   \begin{minipage}{0.45\textwidth}
904:     \begin{center}
905:       \includegraphics[width=7.5cm,clip]{gsfig6.eps}
906:     \end{center}
907:     \caption{Molecule statistics $P(n_b)$ as a function of
908:     the dimensionless parameter $\kappa$ in the presence of
909:       pairing interaction, $\eta=g\sqrt{\Omega}/\chi=10$, for $2M=\Omega=120$.}
910:     \label{gsfig6}
911:   \end{minipage}
912: \end{figure}
913: 
914: This behavior can be understood by noting that the pairing
915: interaction gives rise to an additional detuning effect depending
916: on the number of molecules, as shown by the diagonal terms of the
917: Hamiltonian~(\ref{fbMH}) with $M=\Delta M=\Omega/2$. This
918: interaction leads to number-dependent energy shifts and dephasing
919: between different molecular number states. A similar effect has
920: been studied in the context of a Jaynes-Cummings-like description
921: of photoassociation and has been referred as ``nonlinear detuning''
922: ~\cite{search}.
923: 
924: We can estimate the position of the ``resonance'' point
925: $\omega_{\rm res}(g)$ where the ground state goes from being
926: molecular to atomic in nature by taking $\bra
927: n_b=M/2+1;n_p=M/2-1|H|n_b=M/2+1;n_p=M/2-1\ket =\bra
928: n_b=M/2;n_p=M/2|H|n_b=M/2;n_p=M/2\ket$. This gives
929: \[
930: \omega_{\rm res}(g)=-g\frac{\Omega}{2},
931: \]
932: where we have neglected a term of order $1/\Omega$.
933: 
934: The shift in $\omega_{\rm res}(g)$  due to the pairing interaction
935: is further illustrated in Fig. \ref{gsfig7}, which shows the
936: magnitude of the joint coherence function, $|G_{am}|$, as a
937: function of $\kappa=\omega/\chi\sqrt{\Omega}$ for three values of
938: the pairing coefficient $g$. Finally, Fig. \ref{gsfig8} shows the
939: entanglement entropy $E(\rho_b)$ of the ground state, obtained
940: from the von Neumann entropy of the molecular reduced density
941: operator~\cite{Hines}
942: \[
943: \rho_b=Tr_{f}(\rho),
944: \]
945: as
946: \begin{equation}
947: E(\rho_b)=-\sum_{n_b=0}^{M} |C(n_b)|^2 \log{|C(n_b)|^2},
948: \end{equation}
949: where the logarithm is taken in base $2$, for three values of the
950: pairing interaction strengths, $\eta=0, 5, 10$. The entanglement
951: in Figure~ \ref{gsfig8} is divided by a maximum entanglement,
952: $\log{M}$. Consistently with the results of of Figs. \ref{gsfig5}
953: and \ref{gsfig6} for the molecular statistics, the pairing
954: interaction reduces the entanglement.
955: %
956: \begin{figure}
957:   \begin{minipage}{0.45\textwidth}
958:     \begin{center}
959:       \includegraphics[width=7.5cm,clip]{gsfig7.eps}
960:     \end{center}
961:     \caption{Normalized joint coherence, $|G_{am}|$, of the ground state
962:     as a function of $\kappa$
963:     for three values of the pairing interaction strengths
964:     $\eta=g\sqrt{\Omega}/\chi$,
965:     indicated in the insert.}
966:     \label{gsfig7}
967:   \end{minipage}
968: \hspace{4mm}
969: %
970:   \begin{minipage}{0.45\textwidth}
971:     \begin{center}
972:       \includegraphics[width=7.5cm,clip]{gsfig8.eps}
973:     \end{center}
974:     \caption{Entanglement entropy, $E(\rho_b)$, of the ground
975:     state, in units of the maximum entanglement $\log M$,
976:      as a function of $\kappa$ for three values of the pairing
977:      interaction strengths $\eta=g\sqrt{\Omega}/\chi$ indicated in the insert.}
978:     \label{gsfig8}
979:   \end{minipage}
980: \end{figure}
981: 
982: 
983: \section{Dynamics}
984: 
985: In the absence of  pairing interaction our model is equivalent to
986: the Tavis-Cummings model, whose dynamics has been studied in
987: detail in the context of coherent spontaneous emission from a
988: system of $N$ two-level atoms interacting with a quantized
989: radiation field~\cite{Dicke,Bonifacio}. The main purpose of this
990: section is to extend this work to the study of the system dynamics
991: in the presence of $V_p$. One important result is that the
992: nonlinearity of this interaction produces a self-trapping
993: transition.
994: 
995: We assume that the system consists initially either of atomic
996: pairs or of molecules only, corresponding to a maximal spin state.
997: A photoassociation beam is applied from $t=0$ on, and we examine
998: the subsequent coherent dynamics of the system. Note that although
999: this problem resembles the dynamics of a bosonic Josephson
1000: Junction ~\cite{BJJ1,BJJ2} in an asymmetric trap for the initial
1001: imbalance in populations, the nonlinear coupling term in our model
1002: leads to considerably richer dynamics. Since the total spin is a
1003: constant of motion of the Hamiltonian~(\ref{DGmodel}), we confine
1004: our discussion to the maximal spin state $|S=\Omega/2\ket$.
1005: 
1006: \subsection{Semiclassical approximation}
1007: 
1008: Before proceeding with a full quantum analysis, we first consider
1009: the semiclassical approximation. Our approach is very similar to
1010: that taken in Ref.~\cite{AMBEC2}. Introducing the two operators
1011: \begin{eqnarray}
1012: J_+&\equiv& S^+b,\nonumber \\
1013:  J_-&\equiv&  b^\dagger S^-,
1014: \end{eqnarray}
1015: results in the Heisenberg equations of motion
1016: \begin{eqnarray}
1017: \label{heisenbergeq}
1018: i\frac{d}{dt}J_{+}&=&\omega J_{+}
1019:  +2\chi b^\dagger b S^z+\chi S^+ S^- -2gJ_{+}S^z , \\
1020: i\frac{d}{dt}J_{-}&=&-\omega J_{-}
1021: -2\chi b^\dagger b S^z
1022: -\chi S^+ S^- +2g S^zJ_-, \\
1023: i\frac{d}{dt}S^z&=&\chi (J_+ - J_-).
1024: \end{eqnarray}
1025: Here we note the following relations between operators and conserved
1026: quantities,
1027: \[
1028: S^+ S^- = S(S+1)-S^z(S^z-1), \quad b^\dagger b = M-S-S^z.
1029: \]
1030: 
1031: As usual we introduce a semiclassical approximation by factorizing
1032: the mean values of the various operators that appear in the
1033: Heisenberg equations of motion, such as $\bra S_z J_+\ket=\bra S^z
1034: \ket\bra J_+ \ket$ and $\bra J_- S_z\ket=\bra J_-\ket\bra S_z
1035: \ket$, etc. Introducing the $c-$number functions $s_z\equiv\bra
1036: S_z\ket$, $j_x\equiv \bra (J_{+} + J_{-})\ket/2$, and $j_y\equiv
1037: \bra (J_{+} - J_{-})\ket/2i$, and neglecting corrections of order
1038: $1/\Omega$, that is, setting $\bra S^{+}S^{-}\ket=S^2-s_z^2$,
1039: we obtain the semiclassical equations of motion
1040: \begin{eqnarray}
1041: \label{cdynamics}
1042: \frac{d}{dt}j_x&=&\omega j_y-2gj_ys_z\\
1043: \frac{d}{dt}j_y&=&-\omega j_x+2gj_xs_z-\chi h(s_z)\\
1044: \frac{d}{dt}s_z&=&2\chi j_y,
1045: \end{eqnarray}
1046: where $h(s_z)=-3s_z^2+2(M-S)s_z+S^2$. Noting the additional
1047: conserved quantity
1048: \[
1049: \frac{d}{dt}\left(j_x-\frac{\omega}{2\chi}s_z+\frac{g}{2\chi}s_z^2\right)=0,
1050: \]
1051: we find that the coupled equations~(\ref{cdynamics}) are
1052: equivalent to the classical Newtonian equation of motion
1053: \begin{equation}
1054: \label{Newtoneq} \frac{d^2}{dt^2}s_z=-\frac{d}{ds_z}U(s_z),
1055: \end{equation}
1056: where the potential $U(s_z)$ is determined by the initial conditions
1057: for $j_x(0)$ and $s_0\equiv s_z(0)$. It is sufficient for our
1058: purpose to assume an initial Fock state, so that $j_x(0)=0$ and
1059: $U(s_z)$ is given by
1060: \begin{eqnarray}
1061: \label{scdpotential}
1062: U(s_z)&=&\frac{g^2}{2}s_z^4-(g\omega+2\chi^2)s_z^3\nonumber\\
1063: &+&\left[\frac{\omega^2}{2}+g\omega
1064: s_0-g^2(s_0)^2+2\chi^2(M-S)\right]s_z^2 \nonumber \\
1065: &+&\left[2\chi^2 S^2-\omega^2s_0+\omega gs_0^2 \right]s_z.
1066: \end{eqnarray}
1067: Since the potential has a quartic form of $s_z$,
1068: Eq.~(\ref{Newtoneq}) can be solved analytically in terms of
1069: Jacobian elliptic functions. The derivation of the general
1070: solutions, which is straightforward but lengthy, is given in
1071: Appendix~A.
1072: 
1073: \subsection{Coherent dynamics, $g=0$}
1074: 
1075: We first examine a typical behavior of the semiclassical dynamics
1076: in the absence of pairing interaction, $g=0$, for $M/\Omega\ll 1$,
1077: $M/\Omega \sim 0.5$, and $M/\Omega\simeq 1$.
1078: 
1079: From the semiclassical solutions~(\ref{apg0eq3})
1080: in the case of $g=0$ and on the exact resonance $\omega=0$,
1081: the population imbalance between fermionic pairs and molecules,
1082: $\bra \hat{n}_p\ket - \bra \hat{n}_b \ket=2S-M+2s_z$, and the coherence
1083: function $j_y$ is given by
1084: \begin{eqnarray}
1085: \label{scdresonance1}
1086: \frac{\bra n_p\ket-\bra n_b\ket}{M}&=&2\,{\rm sn}^2(\chi\sqrt{\Omega}t+\phi;k)-1,\\
1087: \label{scdresonance2}
1088: \frac{2j_y}{M\sqrt{\Omega}}&=&2\,{\rm sn}(\chi\sqrt{\Omega}t+\phi;k)
1089: \times{\rm cn}(\chi\sqrt{\Omega}t+\phi;k)\nonumber\\
1090: &\times&{\rm dn}(\chi\sqrt{\Omega}t+\phi;k),
1091: \end{eqnarray}
1092: where {\rm sn}, {\rm cn}, and {\rm dn} are Jacobian elliptic
1093: functions~\cite{elliptic}. Here, $k=\sqrt{M/\Omega}$ is the
1094: elliptic modulus and $\phi$ is a phase factor determined by the
1095: initial conditions. It is equal to $\phi=K$, corresponding to the
1096: complete elliptic integral of the first kind, for an initial
1097: fermionic Fock state $|n_p=M\ket$, and to $\phi=0$ for an initial
1098: the entire population being in the molecular mode. We note that
1099: $j_x(t)$ is zero for all time on the exact resonance.
1100: 
1101: \subsubsection{$M/\Omega \ll 1$ --- Linear coupled-bosons regime}
1102: 
1103: In this case, the elliptic modulus $k\ll 1$ (see Appendix~A) so that
1104: the elliptic functions in Eqs.~(\ref{scdresonance1}) and (\ref{scdresonance2})
1105: can be approximated by ${\rm sn}(u,k)\to \sin{u}$, ${\rm cn}(u,k)\to
1106: \cos{u}$, and ${\rm dn}(u,k)\to 1$, respectively. The imbalance
1107: in atomic and molecular populations which undergoes Rabi
1108: oscillations at the frequency $2\chi\sqrt{\Omega}$ is given by
1109:     \[
1110:         (\bra\hat{n}_p\ket-\bra\hat{n}_b\ket)/M=\cos{(2\chi\sqrt{\Omega}t)}
1111:     \]
1112: for an initial fermionic pair state,
1113:     \[
1114:     (\bra\hat{n}_p\ket-\bra\hat{n}_b\ket)/M=-\cos(2\chi \sqrt{\Omega}t)
1115:     \]
1116: for an initial molecular state. These results are equivalent to
1117: those obtained directly from the Hamiltonian~(\ref{linearboson}).
1118: We have compared these solutions with the exact quantum mechanical
1119: dynamics obtained numerically, and checked that the linear
1120: approximation agrees with the quantum results for
1121: $M/\Omega\lesssim 0.2$ and for times shorter than $t\sim
1122: \pi/\chi\sqrt{\Omega}$.
1123: 
1124: \subsubsection{$M/\Omega \sim 0.5$ --- Intermediate regime}
1125: 
1126: Figure~\ref{dyfig1}a shows the normalized population difference
1127: $(\bra\hat{n}_p\ket-\bra\hat{n}_b\ket)/M$, and Fig.~\ref{dyfig1}b
1128: shows the normalized coherence function $2j_y/M\sqrt{\Omega/2}$,
1129: as a function of the dimensionless time $\tau=\chi\sqrt{\Omega}t$
1130: for a system initially either in a pure atomic state or a pure
1131: molecular state and for $M/\Omega=0.5$. The circles correspond to
1132: the semiclassical description, while the lines are the results of a
1133: full quantum-mechanical analysis. The anharmonicity due to
1134: the nonlinear atom-molecule coupling is clearly apparent, and also
1135: shows that the semiclassical dynamics approximate the quantum
1136: dynamics very well.
1137: 
1138: %
1139: \begin{figure}
1140:     \begin{center}
1141:       \includegraphics[width=7.5cm,clip]{dyfig1.eps}
1142:     \end{center}
1143:     \caption{Comparison of the semiclassical and quantum dynamics
1144:         for $2M=\Omega=120$, $\omega=0$, and $g=0$.
1145:         Figure (a) shows the population difference between fermionic pairs and
1146:         molecules $(\bra\hat{n}_p\ket-\bra\hat{n}_b\ket)/M$. Figure (b) plots
1147:         the coherence function $2j_y/M\sqrt{\Omega}$
1148:         as a function of the dimensionless time $\tau=\chi\sqrt{\Omega}t$.
1149:         The solid and dashed lines give the quantum results for an
1150:         initial fermionic pair sate and a pure molecular state, respectively.
1151:         The corresponding semiclassical results are indicated by filled
1152:         and open circles, respectively.}
1153:     \label{dyfig1}
1154: \end{figure}
1155: 
1156: We note that the atomic pair state in the half-filled shell
1157: corresponds to a Dicke superradiant state~\cite{Dicke}, which is
1158: known from quantum optics to give rise to the strongest collective
1159: enhancement of transition probabilities. This enhancement is
1160: proportional to the product of the number of particle pairs $M$
1161: and the number of hole pairs $\Omega-M$, and is maximum for
1162: $M=\Omega/2$. From Eq.~(\ref{scdresonance1}), we have that for
1163: sufficient short times $\tau=t/\chi\sqrt{\Omega}\ll 1$ the average
1164: number of molecules builds up as
1165: \begin{eqnarray*}
1166: \bra n_b \ket&=&{\rm cn}^2(\chi\sqrt{\Omega}t+\phi)\\
1167: &\rightarrow_{\tau\ll 1}& M\left(\Omega-M\right)(\chi t)^2
1168: =\frac{\chi^2}{4}\Omega^2t^2.
1169: \end{eqnarray*}
1170: %
1171: 
1172: \subsubsection{$M/\Omega \simeq 1$ --- Binary atomic-molecular BEC}
1173: 
1174: As shown in Ref.~\cite{Vardi}, the coherent dynamics in this
1175: regime is qualitatively very similar to that of binary condensate
1176: of atoms and molecules. For $M =\Omega$, corresponding to $k=1$,
1177: the population imbalance between atomic pairs and molecules is
1178: given in the semiclassical approximation by
1179: \begin{equation}
1180: \frac{\bra n_p\ket-\bra
1181: n_b\ket}{M}=2\,\tanh{^2(\chi\sqrt{\Omega}t+\phi)-1},
1182: \end{equation}
1183: indicating that the point $\bra n_p\ket=M$ is stationary. However,
1184: the system is dynamically unstable against small fluctuations ,
1185: see Ref.~\cite{AMBEC2} for a detailed discussion in the context of
1186: binary condensates of atoms and molecules and Ref.~\cite{Walls} in
1187: the context of second-harmonic generation.
1188: 
1189: %
1190: \subsection{\label{selftrap}Self-trapping transition and quantum dynamics}
1191: 
1192: To conclude, we discuss the coherent dynamics of the system in the
1193: presence of pairing interaction, $g\neq 0$, considering only
1194: the case of half-filling for simplicity. We consider specifically
1195: the example $2M=\Omega=120$, which corresponds to $n_F=14$, and
1196: take an initial state as the Fock state $|n_p=M;n_b=0\ket$.
1197: Figure~\ref{appfig1} of the appendix shows the phase diagram of
1198: the semiclassical dynamics in the $\kappa-\eta$ space,
1199: illustrating that a self-trapping transition~\cite{BJJ1,Scott}
1200: takes place when varying the dimensionless detuning frequency
1201: $\kappa=\omega/\chi\sqrt{\Omega}$, provided that the pairing
1202: interaction strength $\eta=g\sqrt{\Omega}/\chi$ exceeds a critical
1203: strength $\eta_c$. For the case of half-filling case, we find
1204: $\eta_c=5.0302$.
1205: 
1206: %
1207: \begin{figure}
1208:     \begin{center}
1209:       \includegraphics[height=4.5cm,width=7.5cm,clip]{dyfig2.eps}
1210:     \end{center}
1211:     \caption{Schematic potential curves, $U(s_z)-U(s_0)$, around
1212:      a self-trapping transition point $\kappa_0$. These curves are
1213:      at (a) above $\kappa_0$, (b) at transition point
1214:      $\kappa=\kappa_0$, and (c) below $\kappa_0$. The filled circle
1215:      indicates an initial position of classical particle.}
1216:     \label{dyfig2}
1217: \end{figure}
1218: %\begin{widetext}
1219: \begin{figure*}
1220:     \begin{center}
1221:       \includegraphics[height=12cm,width=18cm,clip]{dyfig3.eps}
1222:     \end{center}
1223:     \caption{Population imbalance versus the rescaled time
1224:       $\tau=\chi\sqrt{\Omega}t$ for $\eta=6$. Quantum (solid line) and
1225:       semiclassical
1226:       (dashed line) solutions are shown, respectively, for the detuning
1227:       parameters (a) $\kappa=0$ and (b) $\kappa=-3.1$, and
1228:       (c) $\kappa=-3.3$, and (d) $\kappa=-4.0$.
1229:       The dotted line in (c) corresponds to the semiclassical solution
1230:       at the transition point $\kappa_0=-3.2307$.}
1231:     \label{dyfig3}
1232: \end{figure*}
1233: %
1234: \begin{figure*}
1235:   \begin{minipage}{0.45\textwidth}
1236:     \begin{center}
1237:       \includegraphics[height=5cm,width=7.5cm,clip]{dyfig4.eps}
1238:     \end{center}
1239:     \caption{Time-averaged population imbalance as a function of
1240:       $\kappa$ for $\eta=2.0$.
1241:     Quantum result (cross) and semiclassical result (open circle).}
1242:     \label{dyfig4}
1243:   \end{minipage}
1244:   \hspace{4mm}
1245: %
1246:   \begin{minipage}{0.45\textwidth}
1247:     \begin{center}
1248:       \includegraphics[height=5cm,width=7.5cm,clip]{dyfig5.eps}
1249:     \end{center}
1250:     \caption{Time-averaged population imbalance as a function of
1251:       $\kappa$ for $\eta=6.0$.
1252:     Quantum result (cross) and semiclassical result (open circle).}
1253:     \label{dyfig5}
1254:   \end{minipage}
1255: \end{figure*}
1256: 
1257: This transition can be interpreted physically from the motion of
1258: ``classical particle''  in the ``potential'' (\ref{scdpotential}).
1259: Figure~\ref{dyfig2} displays schematic potential curves,
1260: $U(s_z)-U(s_0)$ as a function of $s_z$ in the vicinity of a
1261: self-trapping transition point $\kappa_0$. For our specific
1262: initial conditions the particle ``velocity'' is initially zero,
1263: $ds_z(0)/dt\propto j_y(0)=0$. For $\kappa > \kappa_0$,
1264: Fig.~\ref{dyfig2}-(a), the classical particle oscillates
1265: periodically in the potential. As $\kappa$ approaches $\kappa_0$,
1266: one additional potential barrier appears, and at the transition
1267: point its height equals the initial potential energy of the
1268: particle (see Fig.~\ref{dyfig2}-(b)). At that point, the particle
1269: rests on the potential maximum after reaching it.  Below the
1270: critical point, the barrier confines the particle in a narrow
1271: range, as shown in Fig.~\ref{dyfig2}-(c). The self-trapping effect
1272: provides a sudden suppression of the amplitude of coherent
1273: oscillations. Since the key factor in achieving this transition is
1274: the quartic term in the potential~(\ref{scdpotential}), it
1275: disappears in the absence of pairing interaction.
1276: 
1277: Fig.~\ref{dyfig3} shows the time evolution of the population
1278: difference $(\bra \hat{n}_{p}\ket-\bra \hat{n}_b\ket)/M$ for
1279: $\eta=6$ and for several detuning energies. The semiclassical
1280: solutions (dashed lines) clearly show the self-trapping as a
1281: sudden suppression of coherent oscillations from just above to
1282: just below the transition detuning. The dotted line in
1283: Fig.~\ref{dyfig3}-(c) corresponds to the semiclassical solution
1284: for the threshold detuning $\kappa_0=-3.2307$. The solid lines
1285: show the exact quantum solutions. Apart from the transition point,
1286: the quantum and semiclassical dynamics are similar at least for
1287: short enough times. However, the oscillations of the quantum
1288: solution deviate from those of the semiclassical solution near the
1289: transition point, Fig.~\ref{dyfig3}(b-c). Since in the semiclassical
1290: picture the height of the potential barrier near the transition
1291: point is just below or above the initial potential energy of a
1292: particle, the quantum motion of the particle is very sensitive to
1293: fluctuations and hence deviates significantly from its classical
1294: counterpart. The initial Fock state provides fluctuations of
1295: coherence, and large quantum fluctuations of the population
1296: imbalance arise as a result. We have verified numerically that the
1297: number fluctuations near the transition point are enhanced by an
1298: order of magnitudes as compared to those far away from that point.
1299: 
1300: Figs.~\ref{dyfig4} and \ref{dyfig5} compare the quantum and
1301: semiclassical time-averaged population imbalance as a function of
1302: $\kappa$ for $\eta=2.0$, and for $\eta=6.0$, respectively. In contrast to the
1303: second case, there is no self-trapping transition in the first
1304: case. Hence the semiclassical time-averaged population imbalance
1305: (cross) is a smooth function of $\kappa$, and agrees well
1306: with the quantum results. For the strong pairing coupling of
1307: Fig.~\ref{dyfig5}, in contrast, an abrupt jump of the
1308: semiclassical time-averaged value occurs when varying $\kappa$, a
1309: signature of the self-trapping transition. Due to the large
1310: quantum fluctuations, it differs markedly from the time-averaged
1311: quantum result near a transition point.
1312: 
1313: %
1314: %\begin{figure}
1315: %  \begin{minipage}{0.45\textwidth}
1316: %    \begin{center}
1317: %      \includegraphics[width=7.5cm,clip]{dyfig2a.eps}
1318: %    \end{center}
1319: %    \caption{Dynamics for $g/\chi=1$ and for $\omega/\chi\Omega=0$.}
1320: %    \label{dyfig2a}
1321: %  \end{minipage}
1322: %  \hspace{4mm}
1323: %
1324: %  \begin{minipage}{0.45\textwidth}
1325: %    \begin{center}
1326: %      \includegraphics[width=7.5cm,clip]{dyfig2b.eps}
1327: %    \end{center}
1328: %    \caption{Dynamics for $g/\chi=1$ and for $\omega/\chi\Omega=-0.4$}
1329: %    \label{dyfig2b}
1330: %  \end{minipage}
1331: %\end{figure}
1332: %
1333: %\begin{figure}
1334: %  \begin{minipage}{0.45\textwidth}
1335: %    \begin{center}
1336: %      \includegraphics[width=7.5cm,clip]{dyfig2c.eps}
1337: %    \end{center}
1338: %    \caption{Dynamics for $g/\chi=1$ and for $\omega/\chi\Omega=-0.47$.}
1339: %    \label{dyfig2c}
1340: %  \end{minipage}
1341: %  \hspace{4mm}
1342: %
1343: %  \begin{minipage}{0.45\textwidth}
1344: %    \begin{center}
1345: %      \includegraphics[width=7.5cm,clip]{dyfig2d.eps}
1346: %    \end{center}
1347: %    \caption{Dynamics for $g/\chi=1$ and for $\omega/\chi\Omega=-0.6$}
1348: %    \label{dyfig2d}
1349: %  \end{minipage}
1350: %\end{figure}
1351: 
1352: \section{Summary}
1353: 
1354: In this paper, we have considered the coherent photoassociation of
1355: fermionic atoms into bosonic molecules trapped in a spherically
1356: symmetric harmonic trap. We showed that under a realistic set of
1357: conditions this system can be mapped onto a Tavis-Cummings
1358: Hamiltonian with an additional paring interaction using
1359: pseudo-spin operators. We carried out an exact numerical
1360: diagonalization of the Hamiltonian to determine the ground state
1361: of the system, investigating the crossover from a predominantly
1362: atomic to a predominantly molecular state. We also investigated
1363: the joint coherence and the quantum entanglement between the
1364: atomic and molecular fields, and found that the atomic pairing
1365: interaction suppresses the entanglement between fermions and
1366: bosons. We then analyzed the coherent dynamics of photoassociation
1367: due to the nonlinear atom-molecule coupling. Using a semiclassical
1368: factorization ansatz, we showed the appearance of a self-trapping
1369: transition in the presence of pairing interaction. An exact
1370: quantum solution illustrated the important role of quantum
1371: fluctuations in the neighborhood of that transition point. Future
1372: work will extend this study to a detailed analysis of the
1373: non-degenerate model and to multi-well superradiant systems. For
1374: instance, preparing an atomic Fermi gas in a Josephson-type
1375: configuration and applying a photoassociation beam should lead to
1376: the efficient production of spatially correlated molecules.
1377: 
1378: \acknowledgements This work is supported in part by the US Office
1379: of Naval Research, the National science Foundation, the US Army
1380: Research Office, NASA, and the Joint Services Optics Program.
1381: 
1382: \appendix
1383: \section{Analytic solutions in terms of Jacobian Elliptic
1384: functions}
1385: 
1386: In this appendix, we obtain analytic solutions of semiclassical
1387: dynamics that obey the Newtonian equation of motion
1388: Eq.~(\ref{Newtoneq}), and show the phase diagram of
1389: the semiclassical dynamics based on those solutions.
1390: 
1391: The solution of Eq.~(\ref{Newtoneq}) has the general form
1392: \begin{equation}
1393: \label{integral}
1394: t=\int^{s_z}_{s_0}\frac{ds}{\sqrt{2[U(s_0)-U(s_z)}]},
1395: \end{equation}
1396: with the potential $U(s_z)$ given by Eq.~(\ref{scdpotential}).
1397: 
1398: We analyze the solution for the two cases $g=0$ and $g \neq 0$,
1399: separately, and for simplicity we take the initial state as an
1400: atomic state in the maximum spin manifold $S=\Omega/2$. The
1401: extension to other initial states is straightforward.
1402: 
1403: \subsection{Case $g=0$}
1404: In this case, the ``potential'' $U(s_z)$ is a cubic function of
1405: $s_z$,
1406: \begin{equation} U(s_z)=-2\chi^2
1407: s_z^3+\left[\frac{\omega^2}{2}+2\chi^2 s_0\right]s_z^2
1408: +\left[2\chi^2S^2-\omega^2s_0\right]s_z.
1409: \end{equation}
1410: Here we have used the initial condition $s_0=-S+M$. By introducing
1411: the normalized quantities
1412:    \[
1413:       \bs_z=s_z/2S\quad (-1/2\leq \bs_z\leq 1/2), \qquad
1414:       \bs_0=s_0/2S,
1415:    \]
1416: we obtain the explicit form of the denominator of the right-hand
1417: side of Eq. (\ref{integral}),
1418: \begin{eqnarray}
1419: \varphi(s_z)&\equiv&2[U(s_0)-U(s_z)]\nonumber\\
1420: &=&4\chi^2(2S)^3(\bs_z-\bs_0)(\bs_z-\bs_+)(\bs_z-\bs_-)\\
1421: \bs_{\pm}&=&\frac{\kappa^2}{8}\pm\frac{1}{8}
1422: \sqrt{\kappa^4-16\bs_0\kappa^2+16},
1423: \end{eqnarray}
1424: where $\kappa=\omega/\chi\sqrt{2S}$. We note that the variables
1425: $\bs_\pm$ are always real-valued for any magnitudes of $\kappa^2$,
1426: because $-1/2\le \bs_0 \le 1/2$, and then $\bs_+\ge \bs_0\ge
1427: \bs_z(t) \ge \bs_-$. The solution can be obtained by integrating
1428: the form
1429: \begin{widetext}
1430: \begin{equation}
1431: \label{apg0eq0}
1432: t= \int^{s_z}_{s_0}\frac{ds}{\sqrt{\varphi(s_z)}}
1433: =\frac{1}{\chi\sqrt{2S(\bs_{+} -\bs_{-})}}
1434: \left\{\int^{\phi}_{0}-\int^{\pi/2}_{0} \right\}
1435: \frac{d\theta}{\sqrt{1-k^2 \sin^2{\theta}}},
1436: \end{equation}
1437: \end{widetext}
1438: where,
1439: \[
1440:   \phi(t)=\arcsin{\sqrt{\frac{(\bs_+-\bs_-)(\bs_0-\bs_z)}{(\bs_0-\bs_-)(\bs_+-\bs_z)}}}.
1441: \]
1442: The integral that appears in that equation is an elliptic integral
1443: of the first kind. Noting that the integration within $0\le
1444: \theta\le \pi/2$ gives rise to a complete elliptic integral of the
1445: first kind, $K$, we find
1446: \begin{equation}
1447: \chi\sqrt{2S(\bs_+ -\bs_-)} t+K={\rm sn}^{-1}\left(\sin{\phi(t)};k\right)
1448: \end{equation}
1449: where the function ${\rm sn}^{-1}$ is the inverse of the Jacobian
1450: elliptic function and $k=\sqrt{(\bs_0-\bs_-)/(\bs_+-\bs_-)}$
1451: denotes the elliptic modulus. This gives the evolution of
1452: $\bs_z(t)$
1453: \begin{equation}
1454:     \label{apg0eq1}
1455:     \bs_z(t)=\bs_- +
1456:     (\bs_0-\bs_-){\rm sn}^2\left(\chi\sqrt{2S(\bs_+ - \bs_-)}t+K;k
1457:     \right).
1458: \end{equation}
1459: At the exact resonance, $\omega=0$, we have that  $s_+=1/2$,
1460: $s_-=-1/2$, and $k=\sqrt{M/2S}$ so that this expression reduces to
1461: \begin{equation}
1462: \label{apg0eq3} s_z(t)=-S+M{\rm sn}^2(\chi\sqrt{2S}t+K;k).
1463: \end{equation}
1464: In terms of $s_z(t)$ the coherence functions $j_x(t)$ and $j_y(t)$
1465: are given by
1466: \begin{equation}
1467: \label{apg0eq2}
1468: j_x(t)=\frac{\omega}{2\chi}s_z(t), \quad j_y(t)=\frac{1}{2\chi}
1469: \frac{ds_z(t)}{dt}.
1470: \end{equation}
1471: This results in the expressions for the difference in atomic and
1472: molecular populations $\bra n_{p} \ket-\bra n_b\ket$, and the
1473: coherence function $j_y$ of Eqs.~(\ref{scdresonance1}) and
1474: (\ref{scdresonance2}), respectively.
1475: 
1476: \subsection{Case $g\neq 0$}
1477: The presence of pairing interaction renders the potential
1478: quartic in $s_z$, see Eq.~(\ref{scdpotential}) where we have again
1479: assumed that the initial state is an atomic state of maximum spin,
1480: $|S,-S+M\ket$. It is convenient to introduce the function
1481: $f(\bs_z)$ as
1482: \begin{equation}
1483: \varphi(s_z)=-g^2(2S)^4(\bs_z-\bs_0)\cdot f(\bs_z),
1484: \end{equation}
1485: where
1486: \begin{eqnarray}
1487: f(\bs_z)&=&\bs_z^3+\alpha \bs_z^2+\beta \bs_z+\gamma \nonumber\\
1488: &=&(\bs_z-\bs_1)(\bs_z-\bs_2)(\bs_z-\bs_3),
1489: \end{eqnarray}
1490: $\alpha=\bs_0-2\kappa/\eta-4/\eta^2$,
1491: $\beta=-\bs_0^2+\kappa^2/\eta^2$, and
1492: $\gamma=-\bs_0^3+2\kappa\bs_0^2/\eta-\kappa^2\bs_0/\eta^2+1/\eta$
1493: with $\eta=g\sqrt{\Omega}/\chi$.
1494: 
1495: The variables $\bs_1$, $\bs_2$, and $\bs_3$, which correspond to
1496: the roots of the cubic equation, $f(s_j)=0$, are obtained by
1497: ``Cardano's formula''. With $\xi\equiv e^{i2\pi/3}$,
1498: $p=-\alpha^2/3+\beta$, and $q=2\alpha^3/27-\alpha\beta/3+\gamma$,
1499: and also $D=-(4p^3+27q^2)$, those roots are given by
1500: \begin{eqnarray}
1501: s_j&=&-\frac{\alpha}{3}+\xi^{j-1}\left[-\frac{q}{2}+\frac{1}{6}\sqrt{-\frac{D}{3}}\right]^{1/3}\nonumber\\
1502: &&\qquad+\xi^{1-j}\left[-\frac{q}{2}-\frac{1}{6}\sqrt{-\frac{D}{3}}\right]^{1/3},
1503: \end{eqnarray}
1504: where $j=1,2,3$.
1505: 
1506: The numbers of real and complex roots are determined by the sign
1507: of the polynomial discriminant $D$. If (a) $D > 0$, all three
1508: roots are real and unequal. If (b) $D<0$, one root is real and two
1509: are complex conjugates. If (c) $D=0$, two roots are equal for
1510: $q\neq 0$, and all roots are equal for $q=p=0$.
1511: 
1512: 
1513: \subsubsection{Case $D > 0$}
1514: 
1515: Suppose that $\bs_0\ge \bs_z\ge \bs_a$, and $\bs_a>\bs_b>\bs_c$ or
1516: $\bs_b>\bs_c>\bs_0$, where each $s_{a,b,c}$ corresponds to one of
1517: the roots $s_j$'s. The solution of ~Eq.(\ref{integral}) reads then
1518: \begin{widetext}
1519: \begin{equation}
1520: t=\int^{s_z}_{s_0}\frac{ds}{\sqrt{\varphi(s)}}
1521: =-\frac{1}{gS\sqrt{(\bs_0-\bs_b)(\bs_a-\bs_c)}}\int^\phi_0
1522: \frac{d\theta}{\sqrt{1-k^2\sin^2{\theta}}},
1523: %={\rm sn}^{-1}\left(\sin{\phi(t)}\right),
1524: \end{equation}
1525: where
1526: \begin{equation}
1527: \phi(t)=\arcsin{\sqrt{\frac{(\bs_a-\bs_c)(\bs_0-\bs_z)}{(\bs_0-\bs_a)(\bs_z-\bs_c)}}},
1528: \qquad
1529: k=\sqrt{\frac{(\bs_0-\bs_a)(\bs_b-\bs_c)}{(\bs_0-\bs_b)(\bs_a-\bs_c)}},
1530: \end{equation}
1531: so that
1532: %\begin{widetext}
1533: \begin{equation}
1534: \label{apgeq1}
1535: \bs_z(t)=\frac{\bs_0(\bs_a-\bs_c)+\bs_c(\bs_0-\bs_a){\rm sn}^2\left(-gS\sqrt{(\bs_0-\bs_b)(\bs_a-\bs_c)}t;k\right)}
1536: {(\bs_a-\bs_c)+(\bs_0-\bs_a){\rm sn}^2\left(-gS\sqrt{(\bs_0-\bs_b)(\bs_a-\bs_c)}t;k\right)}.
1537: \end{equation}
1538: \end{widetext}
1539: 
1540: \subsubsection{Case $ D<0$}
1541: 
1542: Letting $\bs_a$ label the real root and with
1543: $\bs_b=\bs_c^*=\bs_R+i \bs_I$, we have that
1544: $\varphi(s_z)=-g^2(2S)^4(\bs_z-\bs_0)(\bs_z-\bs_a)
1545: \left\{(\bs_z-\bs_R)^2+\bs_I^2\right\}$. With the change of
1546: variable from $\bs_z$ ($\bs_0\ge \bs_z \ge \bs_a$) to
1547: \begin{equation}
1548: \frac{\bs_0-\bs_z}{\bs_z-\bs_a}=\frac{A}{B}\frac{1-\cos{\phi}}{1+\cos{\phi}},
1549: \end{equation}
1550: where $A=\sqrt{(\bs_0-\bs_R)^2+\bs_I^2}>0$ and
1551: $B=\sqrt{(\bs_a-\bs_R)^2+\bs_I^2}>0$, and taking the elliptic
1552: modulus as
1553: \begin{equation}
1554: k=\sqrt{\frac{(\bs_0-\bs_a)^2-(A-B)^2}{4AB}},
1555: \end{equation}
1556: the integral Eq.~(\ref{integral}) can then be replaced by
1557: \begin{widetext}
1558: \begin{equation}
1559: t=\int^{s_z}_{s_0}\frac{ds}{\sqrt{\varphi(s)}}
1560: =-\frac{1}{2gS\sqrt{AB}}\int^{\phi}_0\frac{d\theta}{\sqrt{1-k^2\sin^2{\theta}}}.
1561: %={\rm sn}^{-1}(\sin{\phi(t})),\\
1562: \end{equation}
1563: The semiclassical solution is then given by
1564: \begin{equation}
1565: \label{apgeq2}
1566: \bs_z(t)=\frac{\bs_aA+\bs_0B-(\bs_aA-\bs_0B){\rm cn}\left(-2gS\sqrt{AB}t;k\right)}
1567: {A+B-(A-B){\rm cn}\left(-2gS\sqrt{AB}t;k\right)}.
1568: \end{equation}
1569: \end{widetext}
1570: 
1571: \subsubsection{Case $ D=0$}
1572: 
1573: In this case, the solution of Eq.~(\ref{integral}) can be
1574: expressed in terms of elementary functions. For $q\neq 0$ the
1575: solutions are equivalent to Eq.~(\ref{apgeq1}) in which the
1576: elliptic functions are replaced by trigonometric functions for
1577: $k=0$ ($\bs_b=\bs_c$), or hyperbolic functions for $k=1$
1578: ($\bs_0>\bs_a=\bs_b$). If $q=p=0$, the function $f(\bs_z)$ has
1579: triple degenerate roots at a point $\bs_a=\bs_b=\bs_c$, and the
1580: corresponding solution is obtained by
1581: \begin{equation}
1582: \label{apgeq3}
1583: \bs_z(t)=\bs_a+\frac{\bs_0-\bs_a}{1-\left\{-gS(\bs_0-\bs_a)t\right\}^2}.
1584: \end{equation}
1585: 
1586: \subsection{\label{phasediagram}Phase diagram of semiclassical dynamics}
1587: 
1588: In this subsection, we discuss the structure of the semiclassical
1589: dynamics in $\kappa-\eta$ parameter space for the specific case of
1590: a half-filled shell, $2M=\Omega$, by calculating the elliptic
1591: modulus of the semiclassical solution.
1592: 
1593: \begin{figure}
1594:     \begin{center}
1595:       \includegraphics[width=7.5cm,clip]{appfig1.eps}
1596:     \end{center}
1597:     \caption{Elliptic modulus in $\kappa-\eta$ space
1598:       for the half-filled shell: $2M=\Omega$.}
1599:     \label{appfig1}
1600: \end{figure}
1601: 
1602: For $g=0$, Eq.~(\ref{apg0eq1}) describe all possible dynamics for
1603: arbitrary detuning energy $\omega$, while in the presence of a
1604: pairing interaction the dynamics is given by
1605: solutions~(\ref{apgeq1}) and (\ref{apgeq2}), depending on the sign
1606: of $D$. The dynamics is described by Eq.~(\ref{apgeq3}) at a
1607: single singular point in $\kappa-\eta$ space, as we shall see
1608: later.
1609: 
1610: Figure~\ref{appfig1} shows the elliptic modulus of the
1611: semiclassical solution in $\kappa-\eta$ space. The two regions $D>
1612: 0$ and $D<0$ are separated by $D=0$ lines that correspond to the
1613: specific values of elliptic modulus $k=0$ and $k=1$.
1614: Eqs.~(\ref{apgeq1}) and (\ref{apgeq2}) coincide for $k=0$, which
1615: corresponds to the white lines in Fig.~\ref{appfig1}. Hence these
1616: two solutions connect continuously when crossing that line. For
1617: the black line $k=1$, on the other hand, solutions corresponding
1618: to Eq.~(\ref{apgeq1}) differ from Eq.~(\ref{apgeq2}). This
1619: discontinuity gives rise to the self-trapping transition discussed
1620: in subsection~\ref{selftrap}.
1621: 
1622: Figure~\ref{appfig1} shows that the black line and one of the
1623: white lines intersect at the critical point $(\kappa_c, \eta_c)$,
1624: where $q=p=0$, given explicitly by
1625: \begin{eqnarray}
1626: \eta_c&=&2\sqrt{\frac{2}{3}} (45+26\sqrt{3})^{1/4}\simeq 5.0302\\
1627: \kappa_c&=&-\frac{4}{\eta_c}(2+\sqrt{3})\simeq -2.9677.
1628: \end{eqnarray}
1629: Form this result, we conclude that the self-trapping transition
1630: appears by varying the detuning parameter $\kappa$ only for
1631: $\eta>\eta_c$ .
1632: %
1633: %
1634: \begin{references}
1635: \bibitem{Wynar00} R.~Wynar {\it et al.}, Science {\bf 287}, 1016 (2000).
1636: %
1637: \bibitem{Inouye98} S.~Inouye {\it et al.}, Nature (London) {\bf 392}, 151 (1998).
1638: %
1639: \bibitem{Yurovsky99} V.~A.~Yurovsky, A.~Ben-Reuven, P.~S.~Julienne,
1640: and C.~J.~Williams, Phys.~Rev.~A {\bf 60}, R765 (1999).
1641: %
1642: \bibitem{Donley02} E.~A.~Donley, N.~R.~Claussen, S.~T.~Thompson,
1643: and C.~E.~Wieman, Nature (London) {\bf 417}, 529 (2002).
1644: %
1645: \bibitem{Durr04} S.D{\"u}rr, T.~Volz, A.~Marte, and G.Rempe,
1646: Phys.~Rev.~Lett. {\bf 92}, 020406 (2004).
1647: %
1648: \bibitem{Xu03} K.~Xu, T.~Mukaiyama, {\it et al.}, Phys.~Rev.~Lett.
1649: {\bf 91}, 210402 (2003).
1650: %
1651: \bibitem{Petrov04} D.~S.~Petrov, C.~Salomon, and G.~V.~Shlyapnikov,
1652: Phys.~Rev.~Lett. {\bf 93}, 090404 (2004).
1653: %
1654: \bibitem{Greiner03} M.~Greiner, C.~A.~Regal, and D.~S.~Jin, Nature (London)
1655: {\bf 426}, 537 (2003).
1656: %
1657: \bibitem{Zwierlein03} M.~W.~Zwierlein {\it et al.}, Phys.~Rev.~Lett.
1658: {\bf 91}, 250401 (2003).
1659: %
1660: \bibitem{Jochim03} S.~Jochim {\it et al.}, Scince {\bf 301}, 2101 (2003).
1661: %
1662: \bibitem{Stan04} C.~A.~Stan {\it et al.}, cond-mat/0406129.
1663: %
1664: \bibitem{Inouye04} S.~Inouye {\it et al.}, cond-mat/0406208.
1665: %
1666: \bibitem{Kerman04} A.~J.~Kerman, J.~M.~Sage, S.~Sainis,
1667: T.~Bergeman, and D.~DeMille Phys.~Rev.~Lett.
1668: {\bf 92}, 033004 (2004).
1669: %
1670: \bibitem{ex_BECBCS} C.~A.~Regal, M.~Greiner, and D.~S.~Jin,
1671: Phys.~Rev.Lett. {\bf 92}, 040403; M.~Bartenstein {\it et al.},
1672: Phys.~Rev.~Lett. {\bf 92}, 120401 (2004); M.~Zwierlein {\it et al.},
1673: Phys.~Rev.~Lett. {\bf 92}, 120403 (2004).
1674: %
1675: \bibitem{Timmermans} E.~Timmermans, K.~Furuya, P.~W.~Milonni, and A.~K.~Kerman, Phys.~Lett. A {\bf 285}, 288 (2001).
1676: %
1677: \bibitem{th_BECBCS} M.~Holland, S.~J.~J.~M.~F. Kokkelmans, M.~L.~Chiofalo,
1678: and R.~Walser, Phys.~Rev.~Lett. {\bf 87}, 120406 (2001); Y.~Ohashi and
1679: A.~Griffin, Phys.~Rev.~Lett. {\bf 89}, 130402 (2002).
1680: %
1681: \bibitem{Mottelson} A.~Bohr and B.~R.~Mottelson, ``Nuclear Structure''
1682: (Benjamin, New York, 1975), Vols. I and II.
1683: %
1684: \bibitem{SCgrain} C.~T.~Black, D.~C.~Ralph, and M.~Tinkham, Phys.~Rev.~Lett.
1685: {\bf 76}, 688 (1996).
1686: %
1687: \bibitem{Anderson} P.~W.~Anderson, Phys.~Rev.~, {\bf 112}, 1900 (1958).
1688: %
1689: \bibitem{Kerman61} A.~Kerman, Ann.~Phys. (N.Y.), {\bf 12}, 300 (1961).
1690: %
1691: \bibitem{Tavis} M.~Tavis and F.~W.~Cummings, Phys.~Rev. {\bf 170}, 379 (1968).
1692: %
1693: \bibitem{Bogoliubov} N.~M.~Bogoliubov, R.~K.~Bullough, and J.~Timonen,
1694: J.~Phys.~A {\bf 29}, 6305 (1996).
1695: %
1696: \bibitem{Dicke} R.~H.~Dicke, Phys.~Rev. {\bf 93}, 99 (1954).
1697: %
1698: \bibitem{Bonifacio} R.~Bonifatio and G.~Preparata, Phys.~Rev.~A {\bf 2},
1699: 336 (1970).
1700: %
1701: \bibitem{dy_BECBCS} J.~Javanainen {\it et al.}, Phys.~Rev.~Lett. {\bf 92}, 200402 (2004);
1702: A.~V.~Andreev, V.~Gurarie, and L.Radzihovsky, Phys.~Rev.~Lett. {\bf 93},
1703: 130402 (2004); R.~A.~Barankov and L.~S.~Levitov, Phys.~Rev.~Lett. {\bf 93},
1704: 130403 (2004).
1705: %
1706: \bibitem{Vardi} I.~Tikhonenkov and A.~Vardi, cond-mat/0407424.
1707: %
1708: \bibitem{BJJ1} G.~J.~Milburn, J.~Corney, E.~M.~Wright, and, D.~F.~Walls,
1709: Phys.~Rev.~A {\bf 55}, 4318 (1997); A.~Smerzi, S.~Fantoni, S~Giovanazzi,
1710: and S.~R.Shenoy, Phys.~Rev.~Lett. {\bf 79}, 4950 (1997).
1711: %
1712: \bibitem{BJJ2} J.~Javanainen and M.~Y.~Ivanov, Phys.Rev.~A {\bf 60},
1713: 2351 (1999); A.~J.~Leggett, Rev.~Mod.~Phys. {\bf 73}, 307 (2001);
1714: J.~R.~Anglin and A.~Vardi, Phys.~Rev.~A {\bf 64}, 013605 (2001).
1715: %
1716: %
1717: \bibitem{Scott} J.~C.~Eilbeck, P.~S.~Lomdahl, and A.~C~.Scott,Physica D
1718: {\bf 16}, 318 (1985); A.~C.~Scott and J.~C.~Eilbeck, Phys.~Lett. A {\bf 119},
1719: 60 (1986).
1720: %
1721: \bibitem{Racah} G.~Racah, Phys.~Rev. {\bf 63}, 367 (1943).
1722: %
1723: \bibitem{CT} see e.g. C. Cohen-Tannoudji, B. Diu and F. Lalo{\"e},
1724: ``Quantum Mechanics'', Vol.~I Complement $B_{VII}$ (A Wiley-Interscience
1725: Publication, 1977).
1726: %
1727: \bibitem{Heiselberg} H.~Heiselberg and B.~Mottelson, Phys.~Rev.~Lett.
1728: {\bf 88}, 190401 (2002); G.~M.~Bruun and H.~Heiselberg, Phys.~Rev.~A {\bf 65}, 053407 (2002);
1729: H.~Heiselberg, Phys.~Rev.~A {\bf 68}, 053616 (2003).
1730: %
1731: \bibitem{superchemistry} D.~J.~Heinzen, R.~Wynar, P.~D.~Drummond, and
1732: K.~V.~Kheruntsyan, Phys.~Rev.~Lett. {\bf 84}, 5029 (2000).
1733: %
1734: \bibitem{Richardson} R.~W.~Richardson, Phys.~Lett. {\bf 3} 277 (1963);
1735: R.~W.~Richardson and N.~Sherman, Nucl.~Phys. {\bf 52}, 221 (1964).;
1736: R.~W.~Richardson, J.~Math.~Phys. (N.Y.) {\bf 6}, 1034 (1965).
1737: %
1738: \bibitem{Dukelsky}
1739: J.~Dukelsky, C.~Esebbag, and P.~Schuck, Phys.~Rev.~Lett. {\bf 87},
1740: 066403 (2001).
1741: %
1742: \bibitem{AMHamiltonian}
1743: J.~Dukelsky, G.~G.~Dussel, C.~Esebbag, and S.~Pittel, Phys.~Rev.~Lett.
1744: {\bf 93}, 050403 (2004).
1745: %
1746: \bibitem{Schuck} P.~Ring and P.~Schuck, "The Nuclear Many-Body Problem"
1747: (Springer-Verlag, New York, 1980).
1748: %
1749: \bibitem{Walecka} A.~L.~Fetter and J.~D.~Walecka, "Quantum Theory of
1750: Many-Particle Systems" (Dover, Mineola, NY, 2003).
1751: %
1752: \bibitem{HP40} T.~Holstein and Primakoff, Phys.~Rev. {\bf 58}, 1098 (1940).
1753: %
1754: \bibitem{AMBEC1} J.~Javanainen and M.~Mackie, Phys.~Rev.~A {\bf 59},
1755: R3186 (1999).
1756: %
1757: \bibitem{AMBEC2} A.~Vardi, V.~A.~Yurovsky, and J.~R.~Anglin,
1758: Phys.~Rev.~A {\bf 64}, 063611 (2001).
1759: %
1760: \bibitem{Zhou} H.~Q.~Zhou, J.~Links, and R.~H.~McKenzie,
1761: cond-mat/0207540.
1762: %
1763: \bibitem{Hines} A.~P.~Hines, R.~H.~McKenzie, and G.~J.~Milburn,
1764: Phys.~Rev.~A {\bf 67}, 013609 (2003).
1765: %
1766: \bibitem{search} C.~P.~Search, W.~Zhang, and P.~Meystre, Phys.~Rev.~Lett.
1767: {\bf 91}, 190401 (2003); C.~P.~Search, T.~Miyakawa, and P.~Meystre,
1768: Optics Express {\bf 12}, 3318 (2004).
1769: %
1770: \bibitem{elliptic} H.~Hancock: {\it Elliptic Integrals} (Dover Publications,
1771: Inc., New York, 1917); {\it Handbook of Mathematical Functions}
1772: Eddited by M.~Abramowitz and I.~A.~Stegun (Dover Publications, Inc.,
1773: New York, 1972).
1774: %
1775: \bibitem{Walls} D.~F.~Walls, Phys.~Lett. {\bf 32A}, 476, (1970);
1776: D.~F.~Walls and C.~T.~Tindle, J.~Phys.~A {\bf 5}, 534 (1972).
1777: %
1778: \end{references}
1779: \end{document}
1780: