cond-mat0207540/am.tex
1: 
2: 
3: 
4: \documentstyle[aps,prl,multicol,epsfig]{revtex}
5: %\documentstyle[amsfonts,floats,aps,preprint,tighten]{revtex}
6: %\usepackage{amsfonts,latexsym} 
7: %\usepackage{epsf}
8: \begin{document}
9: %\renewcommand{\thefootnote}{\fnsymbol{footnote}}
10: \draft
11: \title{Exact solution, scaling behaviour
12: and quantum dynamics of a model of an atom-molecule Bose-Einstein
13: condensate} 
14: 
15: \author { 
16: Huan-Qiang Zhou\cite{email0}, Jon Links and  Ross
17: H. McKenzie}
18: 
19: \address{Centre for Mathematical Physics, The University of Queensland,
20: 		     4072, Australia } 
21: 
22: 
23: 
24: \maketitle
25: 
26: \vspace{10pt}
27: 
28: \begin{abstract}
29: We study the exact solution for a two-mode model describing coherent
30: coupling between atomic and molecular Bose-Einstein condensates (BEC), in the
31: context of the Bethe ansatz. By combining an asymptotic and numerical
32: analysis, we identify the scaling behaviour of the model and determine
33: the zero temperature expectation value for the coherence and average atomic
34: occupation. The threshold coupling for production of the molecular BEC is
35: identified as the point at which the energy gap is minimum. 
36: Our numerical results indicate a parity effect for the
37: energy gap between ground and first excited state depending on whether
38: the total atomic number is odd or even. The numerical calculations for
39: the quantum dynamics reveals a smooth transition from the atomic to the
40: molecular BEC.
41: \end{abstract}
42: 
43: \pacs{PACS numbers: 03.75.Fi, 05.30.Jp} 
44: 
45:  
46:                      
47: %************************** Text Begins here ******************************
48:  
49: 
50: %  Greek letters
51: \def\aa{\alpha} 
52: \def\bb{\beta}
53: \def\a{\hat a}
54: \def\b{\hat b}
55: \def\d{\dagger}
56: \def\de{\delta} 
57: \def\e{\epsilon}
58: \def\g{\gamma}
59: \def\K{\kappa}
60: \def\ap{\approx}
61: \def\l{\lambda}
62: \def\o{\omega}
63: \def\t{\tilde{\tau}}
64: \def\s{S}
65: \def\D{\Delta}
66: \def\L{\Lambda}
67: \def\T{{\cal T}}
68: \def\TT{{\tilde{\cal T}}}
69: \def\E{{\cal E}} 
70: % Shorthands for \begin{equation} and the like
71: 
72: \def\beq{\begin{equation}}
73: \def\eeq{\end{equation}}
74: \def\bea{\begin{eqnarray}}
75: \def\eea{\end{eqnarray}}
76: \def\ba{\begin{array}}
77: \def\ea{\end{array}}
78: \def\no{\nonumber}
79: \def\le{\langle}
80: \def\re{\rangle}
81: \def\lt{\left}
82: \def\rt{\right}
83: \def\o{\omega}
84: \def\d{\dagger}
85: \def\nn{\nonumber}
86: \def\j{{ {\cal J}}}
87: \def\n{{\hat n}}
88: \def\N{{\hat N}}
89: \def\T{{\cal T}}
90: \def\TT{{\tilde {\cal T}}}
91: 
92: 
93: %\newcommand{\sect}[1]{\setcounter{equation}{0}\section{#1}}
94: %\renewcommand{\theequation}{\thesection.\arabic{equation}}
95: \newcommand{\reff}[1]{eq.~(\ref{#1})}
96: 
97: %\newpage
98: %\vskip.3in
99: \begin{multicols}{2}
100: 
101: 
102: After the experimental realization of a Bose-Einstein condensate
103: (BEC) in
104: dilute alkali gases \cite{and}, many physicists started to 
105: consider the possibility of producing a molecular Bose-Einstein
106: condensate from
107: photoassociation and/or the Feshbach resonance
108: of an atomic Bose-Einstein condensate
109: \cite{wynar,inouye}. As discussed by Zoller \cite{zoller}, this
110: tantalizing problem is now coming toward resolution. 
111: Donley {\it et al.} \cite{wieman} recently reported the creation of a BEC 
112: of coherent superpositions of atomic and molecular $^{85}$Rb states. 
113: This achievement is significant in that the entangled  
114: state is comprised of two chemically distinct components.   
115: 
116: In anticipation of this result, this novel area
117: has attracted  considerable attention from theoretical
118: physicists \cite{drum,java,timm,hein,abee,will,vardi,hope,holland}.
119: Drummond {\it et al.} \cite{drum,hein} emphasized the finite-dimensionality
120: of the system and the importance of quantum fluctuation.
121: Javanainen {\it et al.} \cite{java} systematically analysed the efficiency of
122: photoassociation of an atomic condensate into its molecular counterpart.
123: In Refs. \cite{java,timm,hein}, large-amplitude
124: coherent
125: oscillations between an atomic BEC and a molecular BEC were predicted
126: through the use of the Gross-Pitaevski (GP) mean-field theory (MFT).
127: Others have gone beyond the GP-MFT \cite{hope,holland}. 
128: Vardi {\it et al.} \cite{vardi} suggested that the large-amplitude atom-molecule
129: coherent oscillations are damped by the rapid growth
130: of fluctuations near the dynamically unstable molecular mode, which
131: contradicts
132: the MFT predictions. This has caused some disagreement regarding the
133: atom-molecule conversion and the nature of coherence 
134: \cite{wieman}. In order to clarify the 
135: controversies raised by these investigations, it is highly desirable to
136: extract
137: some rigorous and analytical results. 
138: 
139: The aim of this Letter is to show that a model Hamiltonian
140: to describe coherent coupling
141: between atomic and molecular BEC's described by single modes
142: is exactly solvable in the
143: context of
144: the algebraic Bethe ansatz \cite{kib}. This makes it 
145: feasible to
146: apply techniques well-established in the mathematical physics 
147: literature to
148: study the
149: physics behind this new  phenomenon. 
150: The Bethe ansatz equations are analysed and solved asymptotically in the 
151: limits of the stable molecular regime and the stable 
152: atomic regime. Numerical solutions are also obtained for the crossover
153: regime. We identify a scaling invariance for the model and  
154: show that the exact solution predicts a smooth transition from a
155: quasi-periodic and stable regime to a coherent, large-amplitude,
156: non-periodic oscillating regime. For finite particle number 
157: there are no stationary points in contrast to the  prediction by
158: Vardi {\it et al.} using a linearization scheme \cite{vardi}. Instead, the
159: onset of strong entanglement as the detuning is decreased is identified as the
160: point at which the energy gap to the first excited state 
161: takes the minimum value. 
162: 
163: The Hamiltonian takes the form
164: \beq
165: H= \frac {\delta}{2} {\hat a}^\dagger {\hat a} +\frac {\Omega}{2}
166:  ({\hat a}^\dagger {\hat a}^\dagger {\hat b}
167:   +{\hat b}^\dagger {\hat a}{\hat a}),
168:   \label{ham}
169:   \eeq
170:   where ${\hat a}^\dagger$ and ${\hat b}^\dagger$ denote the creation
171:   operators
172:   for atomic and molecular modes, respectively, $\Omega$ is a
173:   measure of the strength of the matrix element for creation and
174:   destruction of molecules, and $\delta$ is the molecular binding energy
175:   in the absence of coupling. A similar Hamiltonian was first used
176:   to describe optical second harmonic generation \cite{walls} and
177:   (with additional damping terms) photon squeezing experiments in a
178:   two-mode interferometer \cite{chat}.
179:   It was recently investigated by Vardi {\it et al.} \cite{vardi} as
180:   a model for atom-molecule BEC's. Note that the total atom number operator
181: $\N= {\hat n}_a+2{\hat n}_b$, where 
182: ${\hat n}_a={\hat a}^\dagger {\hat a},\,
183: {\hat n}_b=  {\hat b}^\dagger {\hat b}$, provides
184:   a good quantum number since $[H,\,{\hat N}]=0.$
185: 
186: {\it Bethe ansatz solution.} 
187: Following the procedure 
188: of the algebraic Bethe ansatz 
189: \cite{kib}, we have derived the Bethe ansatz 
190: equations (BAE) which solve the model (\ref{ham}). 
191: For given quantum number $M$ the BAE for the spectral parameters
192: $\{ v_i \}$ read 
193: \beq
194: \frac{\delta}{\Omega} - v_i + \frac {2k}{v_i} 
195: = 2 \sum ^M_{j \neq i} \frac {1}{v_j
196: -v_i}. \label {bae}
197: \eeq
198: Above, $k$ is a discrete
199: variable which takes values $1/4$ or $3/4$, dependent on whether the total
200: number of particles is even or odd respectively. More specifically we
201: have $N=2M+2(k-1/4)$. For each $M$ there are $2M+1$ families of solutions
202: $\{ v_i \}$ 
203: to (\ref{bae}). 
204: For a given solution,   
205: the corresponding energy eigenvalue is 
206: \bea E&=&\delta M+\delta(k-1/4)-\Omega\sum_{i=1}^M v_i.  
207: %\nn \\
208: %&=&\delta(k-1/4)-2k\Omega\sum_{i=1}^M\frac{1}{v_i}.  
209: \label{nrg} \eea 
210: Although we will not derive the BAE here, we remark that the model is
211: obtained through a product of the Lax operators for the $su(1,1)$ algebra
212: \cite{jurco,rktb} and the Heisenberg algebra \cite{kt89} in the quasi-classical limit.
213: The construction is similar to that used in the solution of the generalized
214: Tavis-Cummings model from quantum optics \cite{jurco,rktb}. The eigenstates are
215: also obtained through this procedure. Consider the following
216: class of states 
217: \beq \left|v_1,...,v_M\right>=\prod_{i=1}^MC(v_i)\left| 
218: \Psi\right> \label{estates} \eeq  
219: where $C(v)=(v\b^\d-\a^\d\a^\d/2)$, 
220: $\left|\Psi\right>=\left|0\right>$ for $k=1/4$ and
221: $\left|\Psi\right>=\a^\d\left|0\right>$ for $k=3/4$. 
222: In the case when the set of parameters $\{v_i\}$ satisfy the BAE then
223: (\ref{estates}) 
224: are precisely the eigenstates of the Hamiltonian. 
225: 
226: {\it Asymptotic analysis.} In the limit of large $|\de/\Omega|$ we can perform
227: an asymptotic analysis of the Bethe ansatz equations to determine the
228: asymptotic form of the energy spectrum.  
229: We choose the following ansatz for the Bethe roots
230: \bea v_i&\ap& \mu_i\Omega/\delta  
231: ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~i\leq m, \nn \\
232: v_i&\ap&\delta/\Omega +\e_i+\mu_i\Omega/\delta ~~~~~~~~~~~~~~~i>m. \nn \eea
233: Substituting into the BAE and solving for $\e_i,\,\mu_i$ leads to the
234: following asymptotic result for the energy spectrum (cf. \cite{zlmx})   
235: \bea
236: E_m&\ap&
237: \delta (m+k-1/4) \nn \\  
238: &&~~~~~~+
239: \Omega^2(3m^2-m+4km-2kM-2mM)/\delta.   \nn
240: \eea
241: This result can be confirmed by second-order perturbation theory. 
242: The level spacings 
243: $\Delta_m=E_m-E_{m-1}$ are found to be 
244: \bea \Delta_m&\ap& \delta-2\Omega^2(M+2-3m-2k)/\delta \nn
245: . \eea
246: 
247: We let $\E$ denote the ground state energy and $\Delta$ the
248: gap to the first excited state. Employing the Hellmann-Feynman theorem
249: we can determine the asymptotic form of the following zero temperature
250: correlations 
251: $$\left<{\hat n}_a\right>=2\frac{\partial\E}{\partial \delta}, ~~~~
252: \theta=2\frac{\partial \E}{\partial \Omega}  $$   
253: where $\theta \equiv -\left<{\hat a}^\dagger {\hat a}^\dagger {\hat b}
254:   +{\hat b}^\dagger {\hat a}{\hat a}\right>$ 
255:   is the {\it coherence correlator}. 
256: We then have for large $\delta/(\Omega N^{1/2})> 0$   
257: $$ 
258: \frac{\Delta}{\Omega N^{1/2}}\ap\frac{\delta}{\Omega N^{1/2}}
259: -\frac{\Omega N^{1/2}}{\delta},~~~     
260: \frac{\left<n_a\right>}{N}\ap 0, ~~~    
261: \frac{\theta}{N^{3/2}}\ap 0   $$       
262: while for large $\delta/(\Omega N^{1/2})<0$ 
263: $$ 
264: \frac{\Delta}{\Omega N^{1/2}}\ap -\frac{\delta}{\Omega N^{1/2}}
265: -2\frac{\Omega N^{1/2}}{\delta}, ~~    
266: \frac{\left<n_a\right>}{N}\ap 1-\frac{\Omega^2 N}{2\delta^2},$$
267: $$     
268: \frac{\theta}{N^{3/2}}\ap -\frac{\Omega N^{1/2}}{\delta}.  $$ 
269: The above results suggest that the model has scale invariance with respect
270: to the single variable $\Omega N^{1/2}/\delta$. They also 
271: suggest that the scaled gap $\Delta/(\Omega N^{1/2})$ will have a minimum
272: at some positive value of  $\Omega N^{1/2}/\delta$  of the order of unity. 
273: However, as the
274: above is an asymptotic result, we need to undertake a numerical analysis
275: to obtain a more precise picture for the region of small 
276: $|\delta/(\Omega N^{1/2})|$,
277: and to establish that the scaling is also valid
278: in this region. 
279: 
280: 
281: {\it Numerical analysis.} 
282: There is a convenient method to determine the energy spectrum without
283: solving the BAE. This is achieved by resorting to the functional Bethe
284: ansatz \cite{wiegmann}. Let us introduce the polynomial function
285: whose zeros are the roots of the BAE; viz.
286: $$G(u)=\prod_{i=1}^M(1-u/v_i). $$
287: It can be shown from the BAE that $G$
288: satisfies the differential equation
289: $$uG''-(u^2-\delta u-2k)G'+(Mu-E+\delta(k-1/4))G=0$$
290: subject to  the inital conditions
291: $G(0)=1,~~~~G'(0)={E-\delta(k-1/4)}/{2k}.$
292: By setting $G(u)=\sum^M_{n=0} g_n u^n $ the recurrence relation
293: $$ g_{n+1}=\frac{E-\de (n+k-1/4)}{\Omega (n+1)(n+2k)} g_n
294: +\frac{n-M-1}{(n+1)(n+2k)}
295: g_{n-1} $$ 
296: is readily obtained. It is clear from this relation that $g_n$ is a
297: polynomial in $E$ of
298: degree $n$. We also know that $G$ is a polynomial function of degree
299: $M$ and so we must have $g_{M+1}=0$. The $M+1$ roots of $g_{M+1}$
300: are precisely the energy levels $E_m$.
301: Moreover, the eigenstates (\ref{estates}) are expressible as (up to
302: overall normalisation)
303: \beq \left|v_1,...,v_M\right>=\sum_{n=0}^Mg_n(\b^\d)^{(M-n)}(\a^\d\a^\d/2)^n
304: \left|\Psi\right>. \label{gexp} \eeq 
305: 
306: We have implemented these results in order to numerically solve the energy
307: spectrum for various ranges of the coupling parameters. The numerical
308: results show that good  scaling behaviour holds for the entire range  of the
309: couplings, even for small particle numbers. 
310: A typical example for the coherence
311: correlator is shown below. 
312: We remark that the result for the atomic occupation 
313: shown in the inset is in qualitative agreement with
314: Fig. 1 of \cite{kmcj}.
315: 
316: \begin{figure}[h]
317: \centerline{
318: \psfig{file=fig1.eps,width=7.5cm}}
319: %\epsfxsize=0.45\textwidth
320: %\epsfbox{fig1.eps}
321: \caption{
322: Scaling behaviour of the coherence correlator for 
323: the ground state as a function of the scaled detuning  $\delta$.
324: The curves shown are for a total atom number
325: of $N=20,~30$ and 40.
326: The threshold coupling for formation of the
327: predominantly molecular BEC state is $\delta/(\Omega {\sqrt N}) \ap 1.4$,  
328: indicating
329: that there is a wide range of the scaled detuning $\delta$
330: below threshold for which the ground state consists
331: of a coherent superposition of the atomic and
332: molecular states. The inset shows the average fractional occupation 
333: of the atomic state in the ground state.
334: }
335: \label{fig1}
336: \end{figure}
337: 
338: 
339: Some striking features which we have
340: observed are that for fixed $N$ 
341: there are no level crossings in the energy spectrum
342: over the entire range of couplings, and 
343: the existence of a parity effect for the size of the gap 
344: dependent on whether the
345: total atomic number $N$ is odd or even, as 
346: illustrated in Fig. 2.  
347: 
348: The analysis of Vardi {\it et al.} \cite{vardi} in the limit of large $N$ 
349: has shown, for an entire population of molecular modes, 
350: the existence of two
351: stationary  points $|\delta/\Omega|=\sqrt{2N}$,    
352: and for $|\delta/\Omega|<\sqrt{2N}$
353: this state becomes unstable.  
354: Our numerical analysis for finite $N$ shows that a similar situation
355: occurs but the gap never vanishes.  
356: From Fig. 2 we see that the minimum value for
357: $\Delta/\delta$ occurs very close to $\delta=\Omega\sqrt{2N}$.   
358:  For $\delta/\Omega> \sqrt{2N}$ the state consisting entirely of molecules is
359: approximately the ground state which is stable due to the large gap to
360: the next excited level. (The same argument also applies to the negative
361: $\delta/\Omega$, which simply follows from the inherent symmetry in the
362: Hamiltonian (\ref{ham})).
363: 
364: \begin{figure}[h]
365: \centerline{
366: \psfig{file=fig2.eps,width=7.5cm}}
367: %\epsfxsize=0.45\textwidth
368: %\epsfbox{fig2.eps}
369: \caption {
370: Reduction in the energy gap to the first excited
371: state near the threshold coupling for formation of
372: the predominantly molecular BEC. In other words, near threshold
373: quantum fluctuations lead to a significant reduction
374: in the binding energy of the molecules.
375: However, note that the energy gap never vanishes for finite $N$,
376: in contrast to the treatment of Vardi {\it et al.}. 
377: \protect\cite{vardi}.
378: The inset shows a significant parity effect.
379: Near threshold the gap is larger for an odd number
380: of atoms than for an even number of atoms.
381: The solid (dash) curves shown are for $N=70$, 90, 110
382: (71, 91, 111).}
383: \label{fig2}
384: \end{figure}
385: 
386: \def\l{\left}
387: \def\r{\right}
388: {\it Quantum dynamics.} For an initial state $\l |\phi( 0)\r >$, the time
389: evolution is given by $\l|\phi(t)\r > = U(t)\l |\phi(0)\r>$, with 
390: $U(t) = \sum^M_{m=0} \left|m\r>\l<m\r| \exp (-i E_mt)$ where $\l|m\r>$ is
391: the eigenstate with energy $E_m$. We also let $g_n(m)$ denote the
392: co-efficients in (\ref{gexp}) for $\l|m\r>$. Setting
393: ${\hat s} \equiv (2 \b^\dagger b- \a^\dagger \a)/N$, and taking the 
394: initial state as the  
395:  pure molecular state; i.e, $\l|\phi(0)\r> =(1/{\sqrt M!}) (\b^\dagger)^M
396: \l|\psi\r>$, then we have
397: \bea 
398: \l<{\hat s} (t)\r> &=& 1 - (4k-1)/N \nn \\
399: &&-\frac{8}{N}\sum _{m \neq  m'} {\tilde g}_0 (m) {\tilde g}_0 (m')
400: \sin ^2 (\Delta_{mm'}t/2) \nn \\
401: &&\times \sum _{n=0}^M 
402: (n+k) {\tilde g}_n (m) {\tilde g}_n (m'),
403: \eea 
404: with
405: $$
406: {\tilde g}_n =\frac { 2^{-n} g_n \sqrt {(M-n)!(2n+2k-\frac {1}{2})! }}
407: {\sqrt { \sum _{m=0}^M 2^{-2m} (M-m)!(2m+2k -\frac {1}{2})! g^2_m}}
408: $$
409: and $\Delta_{mm'}=E_m-E_{m'}$.   
410: Similarly, if the initial state is a pure atomic state, then
411: \bea 
412: \l <{\hat s}(t)\r > &=& -1 - \frac{8}{N}  
413: \sum _{m \neq  m'} {\tilde g}_M (m) {\tilde g}_M (m')
414: \sin ^2 (\Delta_{mm'}t/2) \nn \\
415: &&\times \sum _{n=0}^M 
416: (M-n) {\tilde g}_n (m) {\tilde g}_n (m').
417: \eea   
418: 
419: 
420: From the expression for $\l<{\hat s}(t)\r>$, we see that the short time behavior is
421: quadratic rather than linear in $t$, due to the square of the sine
422: functions, which is qualitatively
423: consistent with the results of Vardi {\it et al.} \cite{vardi} using a
424: linearized model. As for the long time behavior, one can see that
425: (i) when $N=2$ and 3, there are only two levels, so it is always periodic.
426: (ii) when $\delta/\Omega$ is large and positive, 
427: the ground state and first excited state dominate the dynamics,  
428: since in this regime, the molecular state is approximately the ground state. 
429: As the energy levels are almost equidistant, the evolution is  
430: quasi-periodic (with period $T\approx 2\pi/\Delta$) and stable. A similar
431: situation prevails for largely negative $\delta/\Omega$. 
432: (iii) when $\delta/\Omega$ is small, all the eigenstates are involved in the 
433: evolution due to strong
434: entanglement. This means, all
435: $g_n (m)$  are of the same order
436: for all $\l|m\r>$. In this regime, the levels are not
437: uniformly distributed, so $\l<{\hat s}(t)\r>$ is  a non-trivial sum of  functions
438: $\sin^2(\Delta_{mm'}t)$ 
439: with coefficients depending on $g_n(m)$ and $g_n(m')$.  Therefore, the
440: evolution is not
441: quasi-periodic. 
442: 
443: \begin{figure}[h]
444: \centerline{
445: \psfig{file=fig3.eps,width=8.5cm}}
446: %\epsfxsize=0.45\textwidth
447: %\epsfbox{timed=6.eps}
448: \caption{
449: Time evolution of the average relative atom number
450: $\left<{\hat s}(\tau)\right>$,
451: where $\tau=\Omega \sqrt{N}t$ is the scaled time.   
452: The upper curve
453: depicts evolution of an initial molecular state for $N=40$ and
454: $|\delta/\Omega|=6$ (below threshold coupling). Clearly the dynamics are
455: non-periodic. The lower curve shows collapse and revival behaviour for
456: an initial atomic state for the same parameters.
457: }
458: \label{fig3}
459: \end{figure}
460: 
461: 
462: Vardi {\it et al.} \cite{vardi} claim that in the 
463: large $N$ limit the molecular state is
464: stationary when $\delta/\Omega=\sqrt{2N}$. Observe however that the molecular
465: state is never an eigenstate of the Hamiltonian for finite $\delta/\Omega$.
466: Hence for the molecular state to be stationary it is implied that the
467: gap between the ground and first excited state closes. 
468: For a large but finite $N$, the gap does not close (compare Fig. 2).
469: Our conclusion is that there is   
470: a smooth transition from the large amplitude, non-periodic coherent
471: oscillations
472: (small $\delta/\Omega$)
473: to the 
474: quasi-periodic and stable regime (large $\delta/\Omega$), supported by
475: Figs. 3 and 4.  Our results are in contrast to Fig. 2 in Ref.
476: \cite{vardi} where the relative population was shown for a sufficiently
477: short time interval that only the collapse was seen and on this basis it
478: was mistakenly suggested that this model can describe decoherence
479: effects.
480: 
481: \begin{figure}[h]
482: \centerline{
483: \psfig{file=fig4.eps,width=8.5cm}}
484: %\epsfxsize=0.45\textwidth
485: %\epsfbox{timed=10.eps}
486: \caption{
487: Time evolution of the average relative atom number for $N$=40 and
488: $|\delta/\Omega|=10$ (above threshold coupling). The upper curve for
489: an initial molecular state shows stable quasi-periodic behaviour. The lower
490: curve for an initial atomic state 
491: illustrates the transition from large amplitude collapse and revival 
492: to more stable, small amplitude oscillations, as
493: the detuning $\delta$ is increased.  
494: }
495: \label{fig4}
496: \end{figure}
497: 
498: 
499: {\it Conclusion.} 
500: We have shown that a two-mode model for an atom-molecule BEC
501: is exactly solvable via the Bethe ansatz. We have conducted both asymptotic
502: and numerical analysis to establish scaling invariance of the model, 
503: identified a parity effect in the energy spectrum 
504: and investigated the quantum dynamics. Our results indicate that the
505: transition between the atomic and molecular BEC regimes is smooth, in
506: contrast with Ref. \cite{vardi}. We believe
507: that the exact solution will help to clarify many facets of the
508: coherence properties of this model, in order to qualitatively  
509: compare with current experimental work \cite{wieman}. 
510: 
511: 
512: 
513: %\vskip.3in
514: %\acknowledgments
515: We thank Karen Dancer, Peter Drummond and Chris Tisdell  
516: for valuable discussions. 
517: This work was supported by the Australian Research Council.
518: %\newpage
519: %\vskip.3in
520: \begin{thebibliography}{99}
521: 
522: \bibitem[*]{email0}
523: E-mail: hqz@maths.uq.edu.au
524: \bibitem{and} M.H. Anderson et al., Science 269, 198 (1995).
525: \bibitem{wynar} R. Wynar et al., Science {\bf 287}, 1016 (2000).
526: \bibitem{inouye} S. Inouye et al., Nature {\bf 392},151;
527: J. Stenger et al, Phys. Rev. Lett. {\bf 85}, 4569 (1999);
528: S.L. Cornish et al., Phys. Rev. Lett. {\bf 85}, 1795 (2000).
529: \bibitem{zoller} P. Zoller, Nature {\bf 417}, 493 (2002).
530: \bibitem{wieman} E.A. Donley et al., Nature {\bf 417}, 529 (2002).
531: \bibitem{drum} P.D. Drummond, K.V. Kheruntsyan and H. He, Phys. Rev.
532: Lett. {\bf 81}, 3055 (1998);\\
533: K.V. Kheruntsyan and P.D. Drummond, Phys. Rev. {\bf A61}, 063816 (2000).
534: \bibitem{java} J. Javanainen and M. Mackie, Phys. Rev.
535: {\bf A59}, R3186 (1999).
536: \bibitem{timm} E. Timmermans et al., Phys. Rev. Lett.  {\bf 83}, 2691 (1999).
537: \bibitem{hein} D.J. Heinzen et al., Phys. Rev. Lett. {\bf 84}, 5029 (2000).
538: \bibitem{abee} F.A. van Abeelen and B.J. Verhaar, Phys. Rev. Lett.  
539: {\bf 83}, 1550 (1999).
540: \bibitem{will} C.J. Williams and P.S. Julienne, Science {\bf 287}, 986
541:       (2000).
542: \bibitem{vardi} A. Vardi, V.A. Yurovsky and J.R. Anglin,
543:       Phys. Rev. {\bf A64}, 063611 (2001). 
544: \bibitem{hope} J.J Hope, Phys. Rev. {\bf A64}, 053608 (2001).
545: \bibitem{holland} M.J. Holland, J. Park and R. Walser,
546:       Phys. Rev. Lett. {\bf 86}, 1915 (2001);
547:       S.J.J.M.F. Kokkelmans and M.J. Holland, cond-mat/0204504.
548: \bibitem{kib} V.E. Korepin,
549:       N.M. Bogoliubov, and A.G. Izergin
550:       \emph{Quantum inverse scattering method and correlation functions}
551:       (Cambridge University Press, 1993, Cambridge).
552: \bibitem{walls} 
553: D.F. Walls and R. Barakat, Phys Rev. A {\bf 1}, 446 (1970); \\
554: D.F. Walls and C.T. Tindle, J. Phys. A {\bf 5}, 534 (1972).
555: \bibitem{chat} P.D. Drummond, K. Dechoum and S. Chaturvedi,
556:   Phys. Rev. {\bf A65}, 033806 (2002) and references therein.  
557: \bibitem{jurco} B. Jurco, J. Math. Phys.
558:   {\bf 30}, 1739 (1989).  
559: \bibitem{rktb}  A. Rybin, G. Kastelewicz, J. Timonen and N. Bogoliubov,
560:       J. Phys. {\bf A31}, 4705 (1998).
561: \bibitem{kt89} V.B. Kuznetsov and A.V. Tsiganov, J. Phys. {\bf
562:       A22}, L73 (1989).
563: \bibitem{zlmx} H.-Q. Zhou, J. Links, R.H. McKenzie and X.-W. Guan,
564: cond-mat/0203009.
565: \bibitem{wiegmann} P.B. Wiegmann and A.V. Zabrodin, Nucl. Phys. {\bf
566:       B422}, 495 (1994).
567: \bibitem{kmcj} M. Ko$\check{\rm s}$trun, M. Mackie, R. C$\hat{\rm o}$t\'e 
568: and J. Javanainen, Phys.  Rev. A {\bf 62}, 063616 (2000).
569: 
570: 
571: 
572: 
573: 
574: \end{thebibliography} 
575: \end{multicols}
576: \end{document}
577: 
578: 
579: 
580: 
581: 
582: 
583: 
584: 
585: 
586: 
587: 
588: 
589: 
590: 
591: 
592: 
593: