cond-mat0605486/aa5.tex
1: 
2: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: %%
4: %%    Modelo AA - review
5: %%
6: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7: \documentclass[12pt]{article}
8: \usepackage{cite} 
9: \topmargin-1cm
10: \textwidth16cm
11: \textheight23cm
12: \oddsidemargin0cm
13: \usepackage{amsfonts,latexsym}
14: \usepackage{epsfig}
15: \begin{document}
16: \newcommand{\de}{\delta}\newcommand{\ga}{\gamma}
17: \newcommand{\e}{\epsilon} \newcommand{\ot}{\otimes}
18: \newcommand{\be}{\begin{equation}} \newcommand{\ee}{\end{equation}}
19: \newcommand{\ba}{\begin{array}} \newcommand{\ea}{\end{array}}
20: \newcommand{\beq}{\begin{equation}}\newcommand{\eeq}{\end{equation}}
21: \newcommand{\tmod}{{\cal T}}\newcommand{\amod}{{\cal A}}
22: \newcommand{\bemod}{{\cal B}}\newcommand{\cmod}{{\cal C}}
23: \newcommand{\dmod}{{\cal D}}\newcommand{\hmod}{{\cal H}}
24: \newcommand{\s}{\scriptstyle}\newcommand{\tr}{{\rm tr}}
25: \newcommand{\einsop}{{\bf 1}}
26: \def\R{\overline{R}} \def\doa{\downarrow}
27: \def\dag{\dagger}
28: \def\ve{\epsilon}
29: \def\si{\sigma}
30: \def\ga{\gamma}
31: \def\nn{\nonumber}
32: \def\le{\langle}
33: \def\re{\rangle}
34: \def\lt{\left}
35: \def\rt{\right}
36: \def\dwn{\downarrow}   
37: \def\up{\uparrow}
38: \def\j{\mathcal{J}}
39: \def\dag{\dagger}
40: \def\bea{\begin{eqnarray}}
41: \def\eea{\end{eqnarray}}
42: \def\p{\tilde{p}}
43: \def\q{\tilde{q}}
44: \def\H{\overline{H}}
45: \newcommand{\reff}[1]{eq.~(\ref{#1})}
46: 
47: 
48: 
49: \title{The two-site Bose--Hubbard model}
50: 
51: \author{ J. Links$^1$, A. Foerster$^2$,  A. Tonel$^2$ and  G. Santos$^2$
52: \vspace{1.0cm}\\
53: $^{1}$ Centre for Mathematical Physics, School of Physical Sciences \\ 
54: The University of Queensland, Queensland, 4072, Australia
55: \vspace{0.5cm}\\
56: $^{2}$ Instituto de F\'{\i}sica da UFRGS \\ 
57: Av. Bento Gon\c{c}alves 9500, Porto Alegre, RS - Brazil}
58: 
59: 
60: 
61: 
62: \maketitle
63: 
64: \begin{abstract}
65: The two-site Bose--Hubbard model is a simple model used to study
66: Josephson tunneling between two Bose--Einstein condensates. 
67: In this work we give an overview of some mathematical aspects of
68: this model. Using a classical analysis, 
69: we study the equations of motion and the level curves
70: of the Hamiltonian. Then, the quantum dynamics 
71: of the model is investigated   
72: using direct diagonalisation of the Hamiltonian. 
73: In both of these analyses, 
74: the existence of a threshold coupling between a delocalised 
75: and a self-trapped phase 
76: is evident, in qualitative agreement with experiments.
77: We end with a discussion of the exact solvability of the model via the
78: algebraic Bethe ansatz. 
79: 
80: \end{abstract}
81: 
82: PACS: 03.75.Lm, 32.80.Pj, 03.75.Kk
83: \vspace{1cm}
84: 
85: {\it This work is dedicated to the memory of Daniel Arnaudon} 
86: 
87: \vfil\eject
88: %%%%%%%%%%%%%%%%
89: 
90: \section{Introduction}
91: The phenomenon of Bose--Einstein condensation, while predicted
92: long ago \cite{bose,eins}, is nowadays responsible for many 
93: current perspectives on the potential applications of quantum
94: behaviour in mesoscopic systems. 
95: This point of view has arisen with the experimental
96: observation of condensation in systems of ultracold dilute alkali
97: gases, realised by several research groups using magnetic traps
98: with various kinds of confining geometries \cite{cw,ak}.
99: These types of experimental
100: apparatus open up the possibility for studying  
101: quantum effects, such as Josephson tunneling and self-trapping
102: \cite{wwcch,albiez}, in a macroscopic setting.
103: 
104: {}From the theoretical point of view, the two-site Bose--Hubbard model 
105: (see eq. (\ref{ham}) below), also known as the 
106: {\it canonical Josephson Hamiltonian} \cite{l01}, has been a useful model in 
107: understanding tunneling phenomena. The simplicity of the model means
108: that it is amenable to detailed mathematical analysis, as we will discuss below.
109: However despite this apparent simplicity, the
110: Hamiltonian captures the essence of competing linear and non-linear
111: interactions, leading to interesting, non-trivial behaviour. 
112: 
113: The Hamiltonian is given by 
114: \bea
115: H&=& \frac {k}{8}  (N_1- N_2)^2 - \frac {\Delta \mu}{2} (N_1 -N_2)
116:  -\frac {\cal {E} _J}{2} (b_1^\dagger b_2 + b_2^\dagger b_1).
117:  \label {ham} \eea
118:  where $b_1^\dagger, b_2^\dagger$ denote the single-particle creation
119:  operators in the two wells and  $N_1 = b_1^\dagger b_1,
120:  N_2 = b_2^\dagger b_2$ are the corresponding
121:  boson number operators. The total boson number $N_1+N_2$
122:  is conserved and set to the fixed value of $N$.
123:  The coupling $k$ provides the strength of the scattering interaction between 
124:  bosons,
125:  $\Delta \mu$ is the external potential and ${\cal E}_J$ is the coupling
126:  for the
127:  tunneling.
128:  The change ${\cal E}_J\rightarrow -{\cal E}_{J}$ corresponds to the
129:  unitary
130:  transformation $b_1 \rightarrow b_1,\,b_2\rightarrow -b_2$, while
131:  $\Delta \mu \rightarrow -\Delta \mu$ corresponds to $b_1
132:  \leftrightarrow b_2$.
133:  Therefore we will restrict our analysis to the case of
134:  ${\cal E}_J,\,\Delta\mu\geq 0$.
135:  For $k>0$, following \cite{l01} it is useful to divide the parameter
136:  space into three regimes; viz. Rabi ($k/{\cal E}_J<< N^{-1}$),
137:  Josephson ($N^{-1}<<k/{\cal E}_J<<N$) and Fock ($N<<k/{\cal E}_J $).
138:  For these three regimes, there is a correspondence between
139:  (\ref{ham}) and the motion of a pendulum \cite{l01}. In the Rabi and
140:  Josephson
141:  regimes this motion is semiclassical, in contrast to the Fock regime.
142:  For both the Fock and Josephson regimes the analogy corresponds to a
143:  pendulum with fixed length, while in the Rabi regime the length varies.
144: 
145: 
146: 
147: 
148: 
149: In the present work, we give an overview of some of the mathematical
150: aspects of (\ref{ham}). We 
151: undertake an analysis of the classical and the quantum
152: dynamics of the system, and discuss how  
153: the system exhibits a threshold coupling, originally identified in \cite{milb},
154: about which the dynamics abruptly changes. 
155: Below this threshold point the dynamics is delocalised, 
156: while above it the dynamics turns out to be localised 
157: (macroscopic self-trapping). This can be seen at both the classical and
158: quantum level. 
159: The result is in qualitative agreement with experimental results \cite{albiez}.
160: We conclude by giving an outline of the algebraic Bethe ansatz solution 
161: of (\ref{ham}).
162:   
163: \section{Classical dynamics}
164: 
165: First we study a classical analogue  
166: of  the model. 
167: Let $N_j,\,\theta_j,\,j=1,\,2$ be
168: quantum variables satisfying the canonical relations
169: $$[\theta_1,\,\theta_2]=[N_1,\,N_2]=0,~~~~~[N_j,\,\theta_k]=i\delta_{jk}I.$$
170: Using the fact that
171: $$\exp(i\theta_j)N_j=(N_j+1)\exp(i\theta_j) $$
172: we make a change of variables from the operators
173: $b_j,\,b_j^\dagger,\,j=1,\,2$ via
174: $$b_j=\exp(i\theta_j)\sqrt{N_j},
175: ~~~b_j^\dagger=\sqrt{N_j}\exp(-i\theta_j) $$
176: such that the Heisenberg canonical commutation relations are preserved.
177: Now define the variables
178: $$z=(N_1-N_2)/N $$
179: $$\phi=N(\phi_1-\phi_2)/2 $$
180: where $z$ represents the fractional occupation difference 
181: (or the {\it imbalance}) and 
182: $\phi$ the phase difference. In the classical limit where 
183: $N$ is large, but still finite,  
184: we may equivalently consider the Hamiltonian 
185: \cite{rsfs,ks}
186: \begin{equation}
187: H(z,\phi)= \frac{{\cal E}_JN}{2} \left(\frac{\lambda}{2}z^2-\beta z 
188: - \sqrt{1-z^2}\cos({2\phi}/{N})\right)
189: \label{hsc}
190: \end{equation}
191: where $$\lambda=\frac{kN}{2{\cal E}_J},~~~~~
192: \beta=\frac{\Delta\mu}{{\cal E}_J} $$
193: and $(z,\,\phi)$ are canonically conjugate variables.
194: We note the Hamiltonian (\ref{hsc}) obeys the symmetries 
195:  \begin{eqnarray}\left.H\left(z,\phi\right)\right|_{\lambda,\beta}
196: &=&-\left.H\left(z,\phi+N\pi/2\right)\right|_{-\lambda,-\beta} \nonumber \\
197:  \left.H\left(z,\phi\right)\right|_{\lambda,\beta}
198: &=&\left.H\left(-z,\phi\right)\right|_{\lambda,-\beta}. \label{symm}
199: \end{eqnarray}
200: 
201: 
202: The classical dynamics is given by Hamilton's equations of motion
203: \begin{eqnarray}
204: \dot{\phi}&=& \frac{\partial H}{\partial z}=\frac{{\cal E}_JN}{2} \left(
205: \lambda z - \beta +\frac{z}{\sqrt{1-z^2}}\cos({2\phi}/{N})\right)\nonumber \\
206: \dot{z}&=&-\frac{\partial H}{\partial \phi}=-{\cal E}_J 
207: \left(\sqrt{1-z^2}\sin({2\phi}/{N})\right).
208: \label{eqmov}
209: \end{eqnarray}
210: Now we study the fixed points of the Hamiltonian
211:   (\ref{hsc}),
212:   determined by the condition
213:   $\dot{z}=\dot{\phi}=0$.
214: This leads to the following classification:
215: 
216:   \begin{itemize}
217:   \item $\phi=0$ and $z$ is a solution of
218: 
219:   \begin{equation}
220:   \lambda z -\beta = - \frac{z}{\sqrt{1-z^2}}
221:   \label{teq}
222:   \end{equation}
223:   which has a unique real solution for $\lambda> 0$.
224: 
225:   \item $\phi={N\pi}/{2}$  and $z$ is a solution of
226:   \begin{equation}
227:   \lambda z - \beta= \frac{z}{\sqrt{1-z^2}}.
228:   \label{teq1}
229:   \end{equation}
230:   This equation has either one, two or three real solutions for $\lambda> 0$. \\
231:   \end{itemize}
232: 
233: From eq. (\ref{teq1}) we can determine that there are fixed point
234: bifurcations for certain choices of the coupling parameters.
235: These bifurcations allow us to divide the coupling parameter space in
236: two regions.  
237: Setting $f(z)=\lambda z-\beta$ and 
238: $g(z)=z({1-z^2})^{-1/2}$, the boundary between the regions occurs when
239: $f(z)$ is the tangent line to $g(z)$ at some value $z_0$. A standard
240: analysis shows this occurs when $\lambda=g'(z_0)=(1-z_0^2)^{-3/2}$.
241: Requiring $f(z_0)=g(z_0)$ then yields the following relationship   
242: \begin{equation}
243: \lambda=(1+\left|\beta\right|^{2/3})^{3/2}
244: \label{boundary}
245: \end{equation}
246: determining the boundary. This is depicted in Fig. \ref{fig3}.
247: 
248: %\begin{equation}
249: %\beta = z\frac{\partial}{\partial z}g(z) - g(z)
250: %\label{beta}
251: %\end{equation}
252: 
253: \vspace{1.0cm}
254: \begin{figure}[ht]
255: \begin{center}
256: \epsfig{file=param.eps,width=12cm,height=5cm,angle=0}
257: \caption{Coupling parameter space diagram identifying the different
258: types of solutions for
259: equation (\ref{teq1}). In region I there is just one  solution for 
260: $z$,  a local maximum. 
261: In region II there are  three solutions for
262: $z$, two local maxima and a saddle point.
263: The boundary separating regions I and II is given by (\ref{boundary}). 
264: }
265: \label{fig3}
266: \end{center}
267: \end{figure}
268: 
269: This leads us to the following classification: 
270: \begin{itemize}
271: \item $ 0 < \lambda<1$:
272: For any value of $\beta$ there is just one real solution, for
273: which the Hamiltonian attains a local maximum. \\
274: 
275: \item  $\lambda >1$: Here transition couplings $\pm\beta_{0}$ appear,
276: which can be seen from Fig. \ref{fig3}.
277: For $\beta \in (-\beta_0 ,\beta_{0})$, the equation has two locally
278: maximal fixed points and one saddle point, while for $\beta>\beta_{0}$ or  $\beta <
279: -\beta_{0}$
280:  the equation has just one real solution, a locally
281:   maximal fixed point. 
282:      \end{itemize}
283:       
284: We remark that in the absence of the external potential
285:        $(\Delta\mu=\beta=0)$ the transition value is given by
286: 	 $\lambda_0=1$. Using the symmetry relation (\ref{symm}) we can deduce that for 
287: the attractive case $\lambda<0$, $\lambda_0=-1$ is the coupling
288: marking a bifurcation between a locally minimal fixed point (for $\lambda>-1$) and
289: two locally minimal fixed points and a saddle point (for $\lambda<-1$). This is a supercritical
290: pitchfork bifurcation of the classical ground state. The results of \cite{hines} predict that the ground-state 
291: entanglement, as measured by the von Neumann entropy, is maximal at this coupling. The numerical results of \cite{pd} confirm this.  
292: 
293: 
294: Next we look at the dynamical evolution. 
295: In that which follows we will consider the 
296: equations (\ref{eqmov}) in the absence of the external field 
297: ($\Delta \mu=0$ or, equivalently, $\beta=0$). 
298: An analysis including the effect of this term can be found in references 
299: \cite{ours,buon}.
300: We integrate (\ref{eqmov}) to find the time evolution for the 
301: imbalance $z$, using the initial condition $z(0)=1,\,\phi(0)=0$. 
302: By plotting $z$ against the time, it is evident that
303: there is a threshold coupling $\lambda_c=2$ separating two different 
304: behaviours in the 
305: classical dynamics, as can be seen in Fig. \ref{cd}: 
306: \begin{itemize}
307: \item[(i)] For $\lambda<2$ the  system oscillates 
308: between $z=-1$ and $z=1$. Here the evolution is delocalised;  
309: \item[(ii)] For $\lambda>2$ the system oscillates between $z=0$ and $z=1$. 
310: Here the evolution is localised. 
311: \end{itemize}
312: The threshold occuring at $\lambda_c=2$ 
313: was first observed in \cite{milb}. 
314: 
315: \vspace{1.0cm}
316: \begin{figure}[ht]
317: \begin{center}
318: \epsfig{file=cladinam.eps,width=15cm,height=8cm,angle=0}
319: \caption{Time evolution for the imbalance $z$. The solid line is for 
320: $\lambda=1.9$, 
321: while the dashed curve is for $\lambda=2.1$. Here we are using $N=100$, ${\cal E}_J=1$ 
322: and the initial conditions $z(0)=1$, $\phi(0)=0$. The threshold coupling
323: occurs at $\lambda_c=2$.}
324: \label{cd}
325: \end{center}
326: \end{figure}
327: 
328: To help visualise the classical dynamics, 
329: it is useful to plot the level curves (constant energy curves)  
330: of the Hamiltonian (\ref{hsc}) in phase space. 
331: Given an initial condition $(z(0), \phi(0))$, 
332: the system follows a trajectory along the level curve $H(z(0),\phi(0))$.
333: In Fig. {\ref{level2} we plot the level curves for different values of 
334: $\lambda$ ($\lambda=1.5$ on the left and $\lambda=2.5$ on the right), 
335: where we take $2\phi/N\in[-\pi,\,\pi]$.
336: We can observe clearly two distinct scenarios:
337: \begin{itemize}
338: \item $ \lambda > 2$: Here we see that for the orbit with initial condition $z_0=1,\,\phi(0)=0$, $\phi$ increases monotonically (running phase mode). The evolution of $z$ is bounded in the interval $[0,1]$, leading to localisation
339: (self-trapping). 
340: \item $ \lambda < 2$: Here we see that for the orbit with initial condition $z(0)=1,\,\phi(0)=0$, the evolution of 
341: $\phi$ is oscillatory and bounded in the interval $(-N\pi/2,\,N\pi/2)$. The evolution of $z$ is not bounded, leading to delocalisation. 
342: \end{itemize}
343: 
344: 
345: 
346: \begin{figure}[ht]
347: \begin{center}
348: \begin{tabular}{cc}
349:             &             \\
350:     (a)& (b)    \\
351: \epsfig{file=below.eps,width=6cm,height=6cm,angle=0}&            
352: \epsfig{file=above.eps,width=6cm,height=6cm,angle=0} \\
353: \end{tabular}
354: \end{center}
355: \caption{Level curves of the Hamiltonian (\ref{hsc}) (a) for $\lambda=1.5$ (below the 
356: threshold point) and (b) for $\lambda=2.5 $ (above the threshold point). 
357: We are using $N=100$ and ${\cal E}_J=1 $.
358: Above the threshold coupling running phase modes occur leading to localised evolution of $z$. Below the threshold
359: coupling the evolution of $z$ is delocalised.} 
360: \label{level2}
361: \end{figure}
362: 
363: 
364: The threshold coupling $\lambda_c=2$ (or  
365: $k/{\cal E}_\j=4/N$, in terms of the original variables)
366: separates two distinct dynamical behaviours.
367: This value for the threshold 
368: between delocalisation and self-trapping also occurs for the quantum 
369: dynamics, as we will show in the next section. 
370: 
371: 
372: 
373: %%%%%%%%%%%%%%%%%%%%%%
374: 
375: 
376: \section{Quantum dynamics}
377: 
378: We will investigate the quantum dynamics of the 
379: Hamiltonian in the absence of the external potential
380: $(\Delta \mu=0$)
381: using the exact diagonalisation method. 
382: It is well known that the  
383: time evolution of any state is determined 
384: by $|\Psi(t) \rangle = U(t)|\phi_0 \rangle$, 
385: where $U$ is the temporal evolution operator given by 
386: $U(t)=\sum_{m=0}^{M}|m \rangle \langle m|\exp(-i E_m t)$,  
387: $|m\rangle$ is an eigenstate with energy $E_m$ and $|\phi_0 \rangle$ 
388: represents the initial state.
389: Using these expressions we can compute the expectation value of the 
390: relative number of particles
391: \begin{equation}
392: \langle(N_1-N_2)(t)\rangle=\langle \Psi (t)|N_1-N_2|\Psi (t)\rangle. 
393: \end{equation} 
394: From Fig. \ref{fig.2} it is seen that the qualitative behaviour in
395: each region does not depend on the
396: number of particles. 
397: We find that in the interval
398: $k/{\cal E}_\j\in [1/N^2,1/N]$ (close to the Rabi regime) 
399: % the system exhibits a collapse and revival pattern, in qualitative agreement 
400: % with experiments. 
401: the collapse and revival time takes the constant value
402: $t_{cr}=4\pi$\footnote{The ratio $k/{\cal E}_\j=1/N^2$ means that
403: we are using $k=1$ and ${\cal E}_\j=N^2$ and similarly for the
404: other cases. }
405: %This convention, which will be used throughout the
406: %paper unless noted otherwise, fixes the time scale.}, although
407: %different $N$ with fixed $k/{\cal E}_\j$ 
408: %display different plateau times for the
409: %oscillations.
410: % i.e. the time for which the expectation value is
411: % essentially constant. 
412: In the interval between $k/{\cal E}_\j=1/N$ and $k/{\cal
413: E}_\j=1$ the system undergoes a transition from oscillations which vary between positive and negative values
414: of $\left<N_1-N_2\right>$ 
415: (delocalised) to one where $\left<N_1-N_2\right>$ is close to $N$ (self-trapping). 
416: 
417: \vspace{1.00cm} 
418: \begin{figure}[ht]
419: \begin{center}
420: \epsfig{file=fockreg.eps,width=15cm,height=8cm,angle=0}
421: \caption{Time evolution of the expectation value for the relative
422: number of particles
423: for different ratios of the coupling
424: $k/{\cal E}_\j$ from the top (Rabi regime) to the bottom (Fock
425: regime):
426: $k/{\cal E}_\j= 1/N^2,1/N,1,N,N^2$ for $N=100, 400$ and the initial
427: state is $|N,0\rangle $.}
428: \label{fig.2}
429: \end{center}
430: \end{figure}
431: 
432: Now we focus in more detail the time evolution of the expectation value of the
433: relative number of particles in the interval $k/{\cal E}_\j\in [1/N,1]$
434: In Fig. \ref{fig.3} we present the case $N=100$: we observe 
435: the evolution of the dynamics from a collapse and revival sequence 
436: for $k/{\cal E}_\j<4/N$, through  the self-trapping
437: transition at $k/{\cal E}_\j=4/N$, and toward small amplitude
438: harmonic oscillations in the imbalance of the localised state when
439: $k/{\cal E}_\j=1$. It is also interesting to observe in the localised
440: phase $k/{\cal E}_\j>4/N$ 
441: the re-emergence of a collapse and revival sequence.
442: % Where there is clear collapse and revival
443: % in the self-trapping phase we find that $t_{cr}$ increases with
444: % increasing $k/{\cal E_\j}$. 
445: Further increases in $k/{\cal E_\j}$
446: lead to a decaying of the collapse and revival sequence toward
447: harmonic oscillations which occur at $k/{\cal E}_\j=1$. 
448: A more detailed investigation, using another initial conditions can be found in ref. \cite{our}.
449: 
450: \vspace{1.0cm}
451: \begin{figure}[ht]
452: \begin{center}
453: \epsfig{file=brjmedio.eps,width=15cm,height=8cm,angle=0}
454: \caption{Time evolution of the expectation value between
455: $k/{\cal E}_\j=1/N$
456: and  $k/{\cal E}_\j= 1$. On the left, from the top to the bottom
457: $k/{\cal E}_\j= 1/N,2/N,3/N,4/N$ and on
458: the right, from the top to the bottom  $k/{\cal E}_\j=5/N,10/N,50/N,1$, where
459:  $N=100$ and  the initial state is $|N,0\rangle $.}
460: \label{fig.3}
461: \end{center}
462: \end{figure}
463: 
464: {}From the above picture it is clear that the threshold coupling
465: $k/{\cal E}_\j=4/N $ predicted by the classical analysis, representing 
466: the boundary between a delocalised evolution ($k/{\cal E}_\j < 4/N $) 
467: and self-trapped evolution ($k/{\cal E}_\j> 4/N $), 
468: also holds for the quantum dynamics.
469: 
470: 
471: \section{Exact Bethe ansatz solution}
472: 
473: In this final section we briefly discuss the exact
474: Bethe ansatz solution of (\ref{ham}). More details can be found in \cite{jrlreview,angela}.
475: We begin with the $SU(2)$-invariant $R$-matrix, depending on the
476: spectral parameter $u$:
477: \begin{equation}
478: R(u) =  \left ( \begin {array} {cccc}
479: 1&0&0&0\\
480: 0&b(u)&c(u)&0\\
481: 0&c(u)&b(u)&0\\
482: 0&0&0&1\\
483: \end {array} \right ),
484: \label{r}
485: \end{equation}
486: with $b(u)=u/(u+\eta)$ and
487: $c(u)=\eta/(u+\eta)$.
488: Above, $\eta$ is an arbitrary parameter, to be chosen later.
489: It is easy to check that $R(u)$ satisfies the Yang--Baxter equation
490: \begin{equation}
491: R _{12} (u-v)  R _{13} (u)  R _{23} (v) =
492: R _{23} (v)  R _{13}(u)  R _{12} (u-v).
493: \label{ybe}
494: \end{equation}
495: Here $R_{jk}(u)$ denotes the matrix  acting non-trivially on the
496: $j$-th and $k$-th spaces and as the identity on the remaining space.
497: 
498: Next we define the Yang--Baxter algebra $T(u)$,
499: \begin{equation}
500: T(u)=\left(\matrix{A(u)&B(u)\cr
501: C(u)&D(u)\cr}\right)
502: \label{mono}
503: \end{equation}
504: subject to the constraint
505: \begin{equation}
506: R_{ab}(u-v) T_a(u) T_b(v)=
507: T_b(v) T_a(u)R_{ab}(u-v).
508: \label{yba}
509: \end{equation}
510: We may choose the following realization for the Yang--Baxter algebra
511: \begin{equation}
512: \pi (T_a(u)) = L_{a1}(u + \omega )L_{a2}(u - \omega),
513: \end{equation}
514: written in terms of 
515: the Lax operators \cite{jrlreview}
516: \begin{equation}
517: L_i(u)=\left(\matrix{u+\eta N_i&b_i\cr
518: b_i^{\dagger}&\eta^{-1}\cr}\right)\;\;\;\;\;\; i=1,2.
519: \end{equation}
520: Since $L(u)$ satisfies the relation
521: \begin{equation}
522: R_{ab}(u-v)L_{ai}(u)L_{bi}(v)=L_{bi}(v)L_{ai}(u)R_{12}(u-v), ~~~i=1,\,2   
523: \end{equation}
524: it is easy to check that the Yang-Baxter algebra (\ref{yba})
525: is also obeyed.
526: 
527: Finally, defining the transfer matrix through
528: \begin{equation}
529: t(u) = \pi ( {\rm tr}_a T_a(u)) = \pi ( A(u)+D(u) )
530: \label{tm}
531: \end{equation}
532: it follows from (\ref{yba}) that the
533: transfer matrix commutes for different values of the spectral
534: parameters;
535: i.e., the model is integrable.
536: Now it is straightforward to check that the
537: Hamiltonian (\ref{ham})
538: is related with the
539: transfer matrix  $t$ (\ref{tm}) through
540: $$
541: H=-\kappa \left (t(u) -\frac{1}{4} (t'(0))^2-
542: u t'(0)-\eta^{-2}
543: +\omega^2 -u^2\right),
544: $$
545: where the following identification has been made for the coupling
546: constants
547: \begin{equation}
548: \frac {k}{4} =  \frac {\kappa \eta^2}{2}, ~~~
549: \frac {\Delta \mu}{2} =  -\kappa \eta \omega, ~~~
550: \frac {\cal {E}_J}{2} =  \kappa . \nonumber
551: \end{equation}
552: 
553: We can apply the algebraic Bethe ansatz method, using the  Fock
554: vacuum as the pseudovacuum, to
555: find the Bethe ansatz equations (BAE)
556: \beq
557: \eta^2 (v^2_i -\omega^2)=
558: \prod ^N_{j \neq i}\frac {v_i -v_j - \eta}{v_i -v_j +\eta}
559: \label{becbae} \eeq
560: and the energies of the Hamiltonian (see for example
561: \cite{jrlreview,angela})
562: \begin{eqnarray}
563: E&=&-\kappa\left(\eta^{-2}\prod_{i=1}^N(1+\frac{\eta}{v_i-u})
564: -\frac{\eta^2N^2}{4} -u\eta N-u^2
565: \right. \nonumber \\
566: &&~~~~~~\left.
567: -\eta^{-2}+\omega^2+(u^2-\omega^2)\prod_{i=1}^N(1-\frac{\eta}{v_i-u})
568: \right).     \label{becnrg}
569: \end{eqnarray}
570: Surprisingly this expression is independent of the spectral parameter $u$, which
571: can be chosen arbitrarily.
572: 
573: Using the Bethe ansatz solution, it is possible to derive form factors and correlation functions. Details are given in 
574: \cite{jrlreview}.
575: ~~\\
576: 
577: % \vspace{2.0cm}
578: \centerline{{\bf Acknowledgements}}
579: ~\\
580: AF and GS would like to thank S. R. Dahmen for discussions and CNPq-Conselho Nacional de Desenvolvimento
581: Cient\'{\i}fico e Tecnol\'ogico for financial support. AT thanks FAPERGS-Funda\c{c}\~ao de 
582: Amparo \`a Pesquisa do Estado do Rio Grande do Sul for financial support. 
583: JL gratefully acknowledges funding from the Australian Research Council and The University 
584: of Queensland through a Foundation Research Excellence Award.
585: % \newpage
586: 
587: 
588: 
589: %%%%%%%%%%%%%%%%%%
590: 
591: 
592: \begin{thebibliography}{99}
593: 
594: \bibitem{bose} S. N. Bose, Z. Phys.  {\bf 26} (1924) 178
595: \bibitem{eins} A. Einstein, Phys. Math. K1  {\bf 22} (1924) 261
596: \bibitem{cw} E. A. Cornell and C. E. Wieman, Rev. Mod. Phys. {\bf 74}
597: (2002) 875
598: \bibitem{ak} J. R. Anglin and W. Ketterle, Nature {\bf 416} (2002) 211
599: \bibitem{wwcch} J. Williams, R. Walser, J. Cooper, E. A. Cornell and M.
600: Holland, Phys. Rev. A {\bf 61} (2000) 0336123 
601: \bibitem{albiez} M. Albiez, R. Gati, J. F\"olling, S. Hunsmann, 
602: M. Cristiani and M. K. Oberthaler, Phys. Rev. Lett. {\bf 95} (2005) 010402
603: \bibitem{l01} A. J. Leggett, Rev. Mod. Phys.  {\bf 73} (2001) 307
604: %\bibitem{anderson} M. H. Anderson, J. R. Ensher, M. R. Mathews, C. E. Wieman and E. A. Cornell, Science  {\bf 269} (1995) 198
605: \bibitem{milb} G. J. Milburn, J. Corney, E. M. Wright and D. F. Walls,
606: Phys. Rev. A  {\bf 55} (1997) 4318
607: \bibitem{rsfs} S. Raghavan, A. Smerzi, S. Fantoni and S.R. Shenoy, Phys. Rev. A {\bf 59} (1999) 620
608: \bibitem{ks} S. Kohler and F. Sols, Phys. Rev. Lett. {\bf 89} (2002)
609: 060403 
610: \bibitem{ours} A. P. Tonel, J. Links, and A. Foerster, J. Phys. A: Math. Gen. 
611: {\bf 38} (2005) 6879
612: \bibitem{buon} P. Buonsant, R. Franzosi, V. Penna, 
613: J. Phys. B: At. Mol. Opt. Phys. {\bf 37} (2004) S229 
614: \bibitem{our} A. P. Tonel, J. Links and A. Foerster, 
615: J. Phys. A: Math. Gen. {\bf 38} (2005) 1235
616: \bibitem{hines} A.P. Hines, R.H. McKenzie and G.J. Milburn, Phys. Rev. A {\bf 71} (2005) 042303
617: \bibitem{pd} F. Pan and J.P. Draayer, Phys. Lett. A {\bf 339} (2005) 403
618: \bibitem{jrlreview} J. Links,  H.-Q. Zhou, R. H. McKenzie and M. D.
619: Gould,
620: J. Phys. A: Math. Gen. {\bf 36} (2003) R63
621: \bibitem{angela} A. Foerster, J. Links,  H.-Q. Zhou,
622: in {\it Classical and
623: quantum nonlinear integrable systems: theory and applications}, edited
624: by A.
625: Kundu (Institute of Physics Publishing, Bristol and Philadelphia, 2003)
626: pp 208--233
627: 
628: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
629: %\bibitem{int3} M. R. Andrews, C. G. Townsend, H. -J. Miesner,
630: %D. S. Durfee, D. M. Kurn and W. Ketterle, Science  {\bf 275} (1997) 637
631: %\bibitem{dctw} E. A. Donley, N. R. Claussen, S. T. Thompson and C. E.
632: %Wieman, Nature {\bf 417} (2003) 529
633: 
634: %\bibitem{esks91} V.Z. Enol'skii, M. Salerno, N.A. Kostov and A.C.
635: %Scott,Phys. Scr. {\bf 43} (1991) 229. 
636: 
637: 
638: \end{thebibliography}
639: 
640: \end{document}
641: