cond-mat0405223/bec.tex
1: %\documentclass[aps,pre,showpacs,amssymb,manuscript]{revtex4}
2: \documentclass[aps,pre,showpacs,amssymb]{revtex4}
3: \usepackage{amsfonts}
4: \usepackage{graphicx}
5: \usepackage{bm}
6: 
7: \def\ve{\varepsilon}
8: 
9: \newcommand{\reff}[1]{(\ref{#1})}
10: 
11: \def\NN{{\mathbb{N}}}
12: \def\RR{{\mathbb{R}}}
13: \def\CC{{\mathbb{C}}}
14: \def\ZZ{{\mathbb{Z}}}
15: \def\EE{{\mathbb{E}}}
16: \def\PP{{\mathbb{P}}}
17: \def\SS{{\mathbb{S}}}
18: 
19: 
20: \begin{document}
21: 
22: \title{Collapse  of a Bose-Einstein condensate induced by fluctuations
23: of the laser intensity}
24: 
25: \author{J. Garnier}
26: \affiliation{
27: Laboratoire de Statistique et Probabilit{\'e}s,
28: Universit{\'e} Paul Sabatier,
29: 118 Route de Narbonne, 31062 Toulouse Cedex 4, France,\\
30: Tel. (33) 5 61 55 62 20, 
31: Fax. (33) 5 61 55 60 89, 
32: Email: garnier@cict.fr
33: }
34: 
35: \author{F. Kh. Abdullaev}
36: 
37: \affiliation{Physical-Technical Institute of the Uzbek Academy of Sciences,
38:          G. Mavlyanov str. 2-b, 700084, Tashkent, Uzbekistan}
39: 
40: \author{B. B. Baizakov}
41: 
42: \affiliation{Physical-Technical Institute of the Uzbek Academy of Sciences,
43:          G. Mavlyanov str. 2-b, 700084, Tashkent, Uzbekistan\\
44: and \\
45: Dipartimento di Fisica "E. R. Caianiello" 
46: and Istituto Nazionale di Fisica della Materia (INFM), 
47: Universit{\'a} di Salerno, I-84081 Baronissi (SA), Italy}
48: 
49: \date{\today}
50: 
51: \begin{abstract}
52: The dynamics of a metastable attractive Bose-Einstein condensate trapped by a 
53: system of laser beams is analyzed in the presence of small
54: fluctuations of the laser intensity. It is shown that the condensate
55: will eventually collapse.
56: The expected collapse time is inversely proportional
57: to the integrated covariance of the time autocorrelation function
58: of the laser intensity and it 
59: decays logarithmically with the number of atoms. 
60: Numerical simulations of the stochastic 
61: 3D Gross-Pitaevskii equation confirms analytical predictions for small and moderate 
62: values of mean field interaction.
63: 
64: \pacs{03.75.Kk, 42.65.-k, 42.50.Ar}
65: \end{abstract}
66: 
67: \maketitle
68: \section{Introduction}
69: The experimental realization of Bose-Einstein condensation (BEC)
70: in dilute atomic gases \cite{davis95,anderson96,bradley97} founded a rapidly
71: progressing new field of research \cite{dalfovo}. The physical
72: properties of BECs, which to date comprise eight elements Rb, Na,
73: Li, H, He, K, Cs, Yb and their isotopes, are predominantly
74: determined by interatomic forces. Some of the atomic species
75: ($^7$Li, $^{85}$Rb, $^{133}$Cs) possess a negative $s$-wave
76: scattering length in the ground state and display attractive
77: interactions. 
78: The attractive interaction between the atoms causes the collapse
79: of the BEC so that a stable BEC was not believed to exist \cite{nozieres}.
80: However, when an external spatial confinement is imposed for instance by
81: a system of laser beams, a  trapping potential shows up
82: which can counterbalance the attractive interaction and allows
83: the formation of a metastable BEC.
84: When the number of atoms increases, the attractive interaction
85: becomes stronger and the energy barrier that prevents the 3D BEC 
86: from collapsing becomes weaker.
87: To a given trapping potential there corresponds a critical number of atoms
88: above which the energy barrier vanishes.
89: The case of a quadratic potential has been studied,
90: the critical number of atoms has been computed by a variational
91: approach and by extensive numerical simulations of the Gross-Pitaevskii (GP)
92: equation, and the results have been checked experimentally
93: \cite{dalfovo,roberts01,savage03}.
94: 
95: One of the most important aspects of BECs in the regime of
96: attractive interactions is that they are unstable against
97: collapse. The collapse shows up as a rapid and strong shrinking
98: of the condensate at some critical number of atoms, and is
99: accompanied by significant atomic losses due to many-body
100: processes \cite{donley}. The collapse is initiated when the
101: balance of forces governing the size and shape of the condensate
102: is altered either by internal or external factors.
103: With respect to spatial and energetic stability the magnetic
104: traps appear to be better controllable compared to optical traps
105: \cite{grimm}. On the other hand, due to increasing interest in
106: far-off resonant laser traps for Bose-condensation of atoms which
107: are insensitive to magnetic fields \cite{takasu}, the
108: investigation of different aspects of BEC dynamics in optical
109: traps is becoming a very relevant subject.
110: Of particular interest
111: is the effect of temporal fluctuations
112: of the laser intensity which in turn involve temporal
113: fluctuations of the parabolic trapping potential \cite{savard}.
114: In the present paper we shall consider the BEC dynamics under random
115: fluctuations of the strength of the parabolic trap potential 
116: and we shall show that small fluctuations can lead to the eventual 
117: collapse of the 3D BEC due to a cumulative effect of
118: stochastic perturbations.
119: The random fluctuations have all harmonics in their spectrum,
120: and some of them participate in the parametric resonance
121: leading to collapse.
122: This stochastic parametric resonance in the BEC
123: width oscillations has a rough equivalent particle picture: 
124: the Kramers' exit problem which is concerned with
125: noise activated escape from a potential well \cite{hanggi}.
126: 
127: Quantum tunneling (QT) is considered as playing a key role in the
128: condensate collapse when the number of atoms is close to the critical
129: number \cite{leggett}.
130: We shall see that the BEC instability driven by random
131: fluctuations of the strength of the parabolic trap potential
132: is all the more
133: dramatic as the number of atoms is closer to the critical number.
134: Our consideration thus shows that even weak noise can play a
135: competitive role in this limit with QT and should be taken into account.
136: The effect of optical trap noise was previously considered in the
137: context of stochastic heating of trapped atoms \cite{savard,gehm}.
138: In a far-off resonant optical trap created by a system of red
139: detuned lasers the variable trapping potential can be represented
140: as $V(t,r)=-\alpha |E(t,r)|^2/4$, where $\alpha$ is the atomic
141: polarizability and $E(t,r)$ is the electric field amplitude. The
142: dynamics of trapped atoms can be described by the corresponding
143: Hamiltonian $H=p^2/(2m) + (1/2) m  \omega_0^2 (1+\eta(t))
144: r^2$, where $\omega_0^2 = k_0^2/m$ is the mean square trap
145: oscillation frequency, and $k_0$ is proportional to the
146: time-averaged laser intensity $I_0 \sim |E|^2$. The time
147: dependent spring constant is determined by fractional fluctuations
148: of the laser intensity $\eta(t)=(I(t)-I_0)/I_0$ \cite{savard} .
149: The influence of the fluctuations of the trap potential on
150: the dynamics of 1D GP type equation has been considered in \cite{ABP}
151: and the trap and nonlinearity fluctuations in two dimensional BEC in
152: \cite{ABK,ABG}.
153: 
154: The paper is organized as follows.
155: In Section \ref{sec:model} we give a description of the model and
156: apply a variational approach.
157: In Section \ref{sec:action} we derive the effective dynamics
158:  of the action-angle variables of the system driven by random perturbations.
159: Section \ref{sec:smallnl} (resp. \ref{sec:crit}) are devoted to the asymptotic
160: analysis of the system for small (resp. near-critical) number of atoms.
161: Finally we check the variational approach and our asymptotic analysis
162: in Section \ref{sec:num} by performing direct numerical
163: simulations of the GP equation.
164: 
165: \section{The model and the variational approach}
166: \label{sec:model}%
167: We consider the mean-field GP equation for the single-particle
168: wave function \cite{gross}
169: \begin{equation}
170:   \label{eq:schro0}
171:   i \hbar \psi_t = 
172: -\frac{\hbar^2}{2m} \Delta \psi + V(t,{\bf r}) \psi + g |\psi|^2 \psi .
173: \end{equation}
174: The nonlinear coefficient is
175: $g= 4 \pi \hbar^2
176: a_s/m$
177: where $a_s$ and $m$ are respectively the atomic scattering length and
178: mass.
179: The number of atoms is $N=\int |\psi|^2 dx$.
180: $V$ is the external trapping potential imposed by a system
181: of laser beams. We consider a harmonic model, but we take into
182: account temporal fluctuations of the laser intensity which in turn 
183: induces temporal fluctuations of the quadratic potential
184: \begin{equation}
185:   \label{eq:potential}
186:   V(t,{\bf r}) = \frac{m \omega_0^2}{2} |{\bf r}|^2 [ 1 +  \eta (t) ] .
187: \end{equation}
188: For the optical trap $\omega^{2}= \alpha I/(2 m l_{0}^{2})$, 
189: where $l_{0}$
190: is the size of the laser beam, $I$ is the intensity, $\alpha$ is a
191: constant proportional to the 
192: laser frequency detuning. The random function $\eta(t)$ describes the 
193: laser intensity fluctuations
194: $\eta(t) =(I(t)-I_{0})/I_{0}$.
195: The stationary random process $\eta$ has zero-mean and standard
196: deviation $\sigma_\eta$.
197: We shall see in the following that the
198: standard deviation is not sufficient to predict the collapse of the
199: BEC, but the coherence time and more generally the power spectral
200: density of $\eta$ will play a role.
201: 
202: We now cast Eq.~(\ref{eq:schro0}) in a dimensionless form by introducing
203: the variables
204: $t'=\omega_0 t$, ${\bf r}' = {\bf r} / r_0$,
205: $r_0^{-1} = \sqrt{m \omega_0/ \hbar}$,
206: and $u = \sqrt{4\pi |a_s| r_0^2} \psi$.
207: This yields the following partial differential equation (PDE)
208: \begin{equation}
209:   \label{eq:schro1}
210:    i u_{t'} = - \frac{1}{2}
211: \Delta' u + \frac{1}{2} |{\bf r}'|^2 [1+ \eta'(t')] 
212: u + \sigma_s |u|^2 u ,
213: \end{equation}
214: where $\sigma_s = {\rm sgn}(a_s) = \pm 1$ and $\eta'(t') = \eta(t' /
215: \omega_0)$.
216: From now on we drop the primes. 
217: The next step consists in applying the variational approach.
218: This approximation was first introduced by Anderson \cite{anderson83}
219: and developed in nonlinear optics \cite{malomed}.
220: A similar technique was elaborated for the BEC dynamics 
221: based on the GP equation \cite{perez97}.
222: The variational ansatz for the wave function of the BEC is chosen as
223: the Gaussian \cite{dalfovo}
224: \begin{equation}
225:   \label{eq:ansatz0}
226:   u(t,{\bf r}) = A(t) \exp \left( - \frac{|{\bf r}|^2}{2 a(t)^2} +
227:     \frac{i b(t) |{\bf r}|^2}{2} + i \theta(t) \right) .
228: \end{equation}
229: $a(0) r_0$ is the initial BEC rms width in physical variables
230: $$
231: a(0) = \frac{\sqrt{2} }{ \sqrt{3} \sqrt{N} r_0} \left( \int  |{\bf r}|^2| 
232: \psi(t=0,{\bf r})|^2 d^3 {\bf r}
233: \right)^{1/2}.
234: $$
235: The number of atoms is 
236: $$
237: N= \frac{\sqrt{\pi} r_0}{4 |a_s|}
238:  A(0)^2 a(0)^3 = \frac{\sqrt{\pi} r_0}{4 |a_s|} A(t)^2 a(t)^3.
239: $$
240: Following the standard procedure \cite{malomed}, 
241: we substitute the ansatz into the
242: Lagrangian density generating Eq.(\ref{eq:schro1}) and calculate the
243: effective Lagrangian density in terms of $A$, $a$, $b$, $\theta$
244: and their time-derivatives.
245: The evolution equations for the parameters of the ansatz are then
246: derived from the effective Lagrangian by using the corresponding
247: Euler-Lagrange equations.
248: In particular this approach yields a closed-form 
249: ordinary differential equation (ODE) for 
250: the BEC width $a$
251: \begin{equation}
252: \label{eq:ode1}
253: a_{tt} + a (1 + \eta(t) ) =
254: \frac{1}{a^3} +\frac{\sigma_s P}{a^4} ,
255: \end{equation}
256: where $P = \sqrt{2/\pi} N |a_s| /r_0$.
257: We study in this paper the attractive case ($a_s<0$, $\sigma_s=-1$).
258: The evolution equation finally reads
259: \begin{equation}
260: \label{ode1}
261: a_{tt} + a (1 +  \eta(t) ) =
262: \frac{1}{a^3} - \frac{P}{a^4} .
263: \end{equation}
264: 
265: 
266: \section{Action-angle variables}
267: \label{sec:action}
268: 
269: \subsection{Unperturbed dynamics}
270: The energy $E$ of the unperturbed BEC is given by:
271: \begin{equation}
272: E(t) = \frac{1}{2} a_t^2(t) + {U} (a(t)),
273: \hspace{0.1in}
274: {U}(a) = \frac{1}{2} \left( a^2  + \frac{1}{a^2} 
275: \right) - \frac{P}{3 a^3} .
276: \end{equation}
277: In absence of random fluctuations $\eta \equiv 0$ 
278: the energy $E$ is an integral of motion.
279: The BEC width obeys a simple dynamics with Hamiltonian
280: structure
281: \begin{equation}
282: \label{mvt0}
283: H(p,q) = \frac{1}{2} p^2 + {U}(q)
284: \end{equation}
285: with $q=a$ and $p=a_t$.
286: A straightforward analysis \cite{ruprecht,dalfovo} 
287: shows that if $P<P_c = 4/5^{5/4}
288: \simeq 0.535$, then the potential $U$ possesses a local minimum
289: that we shall denote by $a_0$ (see Fig.~\ref{fig1}).
290: The corresponding ground state has energy $E_0=U(a_0)$.
291: Below $a_0$ there is the local maximum $a_1$ with energy $E_1=U(a_1)$,
292: and below $a_1$ the potential decays to $-\infty$.
293: Above $a_0$ the potential increases to $+\infty$.
294: It crosses the energy level $E_1$ at $a_2$.
295: 
296: 
297: \begin{figure}
298: \begin{center}
299: \begin{tabular}{c}
300: \includegraphics[width=7.cm]{fig1.eps}
301: \end{tabular}
302: \end{center}
303: \caption{Potential $U(a)$ for $P=0.2$.
304: The important points $(a_1<a_0<a_2)$ are also represented.
305: \label{fig1}
306: }
307: \end{figure}
308: 
309: 
310: If the initial conditions $(a(0),a_t(0))$
311: correspond to an energy above $E_1$, or below $E_1$
312: but $a(0)<a_1$, then the condensate width goes to zero in finite
313: time which means that the BEC collapses.
314: On the contrary, if the initial conditions $(a(0),a_t(0))$
315: correspond to an energy between $E_0$ and $E_1$,
316: and $a(0)>a_1$, then the orbits of the motion are closed, 
317: corresponding to periodic oscillations.
318: In order to explicit the 
319: periodic structure of the variables $a$ and $a_t$,
320: we introduce the action-angle variables.
321: The orbits are determined by the  energy imposed by the initial
322: conditions:
323: $$
324: E =  \frac{1}{2} a_t^2(0) + {U} (a(0)) .
325: $$
326: For $E \in (E_0,E_1)$,
327: we introduce $e_1(E)<e_2(E)$ the extremities
328: of the orbit of $a$ for the energy $E$:
329: $$
330: U(e_1(E)) =  U(e_2(E)) = E .
331: $$
332: The action $I$ is defined as a function of the energy $E$ by
333: \begin{equation}
334: \label{def:action}
335: {\cal I} (E)= \frac{1}{2\pi} \oint p dq 
336: %= \frac{1}{2\pi} \oint \sqrt{2E-2{U} (b)} db
337: =
338: \frac{1}{\pi} \int_{e_1(E)}^{e_2(E)} \sqrt{2E-2{U} (b)} db .
339: \end{equation}
340: The motion described by (\ref{mvt0})
341: is periodic, with period
342: \begin{equation}
343: \label{def:calt}
344: {\cal T}(E) 
345: = \oint \frac{dq}{p} 
346: %=\oint \frac{db}{\sqrt{2E - 2{U}(b)}} 
347: =
348:  2
349: \int_{e_1(E)}^{e_2(E)} \frac{db}{\sqrt{2E - 2{U}(b)}}  
350: \end{equation}
351: or else ${\cal T}(E) = 2 \pi \frac{d{\cal I}}{dE}(E)$.
352: The angle $\phi$ is defined as a function of $I$ and $a$ by
353: $$
354: \phi (E,a) = -\int^{a} \frac{\partial p}{\partial I} dq 
355: = -\frac{2\pi}{{\cal T}(E)} \int^{a} \frac{db}{\sqrt{2E-2{U}(b)}} . 
356: $$
357: The transformation $(E,a) \rightarrow (I,\phi)$
358: can be inverted to give the functions
359: ${\cal E}(I)$ and ${\cal A}(I,\phi)$.
360: The BEC width oscillates between the minimum value $e_1(E)$
361: and the maximum value $e_2(E)$.
362: The energy $E$ as well as the action $I$ are constant and fixed by the
363: initial conditions, so the
364: evolution of the BEC width is governed by
365: \begin{eqnarray*}
366:   && a(t) = {\cal A}( {\cal I}(E) , \phi(t)) ,\\
367:   && \phi(t) = \phi(0) - \frac{2 \pi}{{\cal T}(E)} t .
368: \end{eqnarray*}
369: 
370: 
371: 
372: \subsection{Perturbed dynamics}
373: \label{sec:pert1}%
374: From now on we assume $\eta \not\equiv 0$ and we denote
375: by $\sigma_\eta$ the standard deviation of $\eta$.
376: We investigate the stability of the BEC when the unperturbed
377: motion is oscillatory.
378: In particular we aim at studying the collapse time $T_c$
379: defined as the first time $t$ such that $a(t)=0$.
380: While the energy of the BEC is below $E_1$, 
381: the orbit is closed. As soon as the energy reaches the energy level
382: $E_1$,
383: the BEC collapses in a time of order $1$ (w.r.t. $\sigma_\eta$).
384: We shall show that the hitting time for the energy level
385: $E_1$ is of order $\sigma_\eta^{-2}$, so the collapse time $T_c$
386: is imposed by the hitting time $T_h$ defined as the first time
387: $t$ such that $ E(t)=E_1$ or equivalently $I(t)=I_1 := {\cal I}(E_1)$.
388: 
389: In  presence of perturbations, 
390: the motion of $a$ is not purely oscillatory,
391: because the energy and the action are slowly varying in time.
392: We adopt the action-angle formalism, because it allows us to 
393: separate the fast scale of the locally periodic motion and the slow
394: scale of the evolution of the action.
395: Thus, after rescaling $\tau = \sigma_\eta^2 t$ 
396: the action-angle variables satisfy the differential equations
397: \begin{equation}
398: \label{form}
399: \left\{
400: \begin{array}{l}
401: \displaystyle
402: \frac{ d I}{d\tau} = 
403: \frac{1}{\sigma_\eta} \eta(\frac{\tau}{\sigma_\eta^2})
404: h_\phi(I ,\phi) ,\\
405: \displaystyle
406: \frac{ d \phi}{d \tau} = - \frac{1}{\sigma_\eta^2}\omega (I) 
407: - \frac{1}{\sigma_\eta} \eta(\frac{\tau}{\sigma_\eta^2}) 
408: h_I(I,\phi) ,
409: \end{array}
410: \right.
411: \end{equation}
412: where $h(I,\phi)= \frac{1}{2} {\cal A}^2(I,\phi)$
413: and $\omega(I) = \frac{2 \pi }{{\cal T} ( {\cal E} (I))}$
414: are smooth functions and $h$ is
415: periodic with respect to $\phi$ with period $2\pi$.
416: The normalization $\tau = \sigma_\eta^2 t$ has been 
417: chosen so that the random process $\eta$ appears
418: with the scales of a white noise in 
419: the differential equations (\ref{form}).
420: Applying a standard diffusion-approximation theorem 
421: \cite{psv}, we get that $(I(t) )_{t \geq 0}$
422: behaves like a diffusion Markov process
423: with the infinitesimal generator
424: $$
425: {\cal L}_{{I}}= \frac{1}{2} A( {I}) \frac{\partial^2}{\partial  {I}^2}
426: + B ( {I})  \frac{\partial}{\partial  {I}}
427: $$
428: where
429: \begin{eqnarray*}
430:   && A( {I} ) = \frac{1}{\pi} \int_0^{2\pi}
431: \int_0^{\infty}
432: h_\phi( {I} ,\phi ) h_\phi( {I} ,\phi - \omega({I}) t) 
433: \EE[\eta(0) \eta(t) ] dt d\phi ,
434:   \\
435:   && B( {I} ) =  \frac{1}{\pi} \int_0^{2\pi} \int_0^{\infty}
436: h_\phi( {I} ,\phi ) h_{\phi I}( {I} ,\phi - \omega( {I}) t) 
437: \EE[\eta(0) \eta(t) ] dt d\phi .
438: \end{eqnarray*}
439: This means in particular that the probability density function of
440: ${I}(t)$ satisfies the Fokker-Planck equation $\partial_t p = {\cal L}_I^* p$,
441: $p(t=0,{I})  = \delta({I}-I_0)$, where $I_0$
442: is the initial action at time $0$ and ${\cal L}_I^*$
443: is the adjoint operator of ${\cal L}_I$, i.e.
444: ${\cal L}_I^* p=  (1/2) \partial_I^2 \left[ A( {I}) p \right]
445: - \partial_I \left[ B ( {I}) p \right]$.
446: Moreover, 
447: standard results of stochastic analysis allow us to compute
448: recursively the moments of $T_h$ \cite{feller}.
449: Denoting  $I_1 = {\cal I}(E_1)$,
450: the first moment ${\cal \mu}^{(1)} (I)= \EE_I [T_h]$
451: (the mean value of $T_h$ starting from action $I$ at time $0$)
452: satisfies
453: \begin{equation}
454: \label{feller1}
455: {\cal L}_I \mu^{(1)} = -1, \ \ \ \ 
456: \mu^{(1)}(I_1)=0.
457: \end{equation}
458: The $n$-th moment ${\cal \mu}^{(n)} (I)= \EE_I [T_h^n]$
459: satisfies
460: \begin{equation}
461: \label{feller2}
462: {\cal L}_I \mu^{(n)} = -n \mu^{(n-1)}, \ \ \ \ 
463: \mu^{(n)}(I_1)=0.
464: \end{equation}
465: In the following sections we shall apply and discuss these general
466: results in two different frameworks: small and critical nonlinearity.
467: 
468: 
469: \section{Small nonlinearity}
470: \label{sec:smallnl}%
471: \subsection{Expansions of the action-angle variables for small nonlinearity}
472: In this section we assume that $P \ll 1$ which will allow us
473: to derive simple expressions for the physically relevant quantities.
474: The points $a_j$ and $E_j$ can be expanded
475: for small nonlinearity $P$ as
476: \begin{eqnarray*}
477: &&
478: a_0 = 1 
479: + O(P), \ \ \ \
480: a_1 = P +O(P^2) , \ \ \ \
481: a_2= \frac{1}{\sqrt{3} P} +O(1) ,\\
482: && 
483: E_0 = 1 
484: +O(P), \ \ \ \
485: E_1= \frac{1}{6 P^2} + O(\frac{1}{P}) .
486: \end{eqnarray*}
487: Note that, as $P$ becomes small, the potential barrier grows like
488: $P^{-2}$, which shows that the trap looks like a deep 
489: quadratic external potential.
490: The functions $h(I,\phi)$ and $\omega(I)$ can also be expanded
491: for any $\phi $ and $I \leq I_1 = {\cal I}(E_1) = 1/(12 P^2) + 
492: O( 1/P)$:
493: \begin{eqnarray*}
494:   &&h(I,\phi) = \frac{1}{2} + I + \sqrt{I+I^2} \cos(\phi) +O(P) ,\\
495:   &&\omega(I) = 2 +O(P).
496: \end{eqnarray*}
497: Accordingly the unperturbed dynamics of the BEC width for small
498: $P$ is approximately
499: \begin{equation}
500:   \label{dyna1}
501:   a(t) =   \sqrt{ 1+2I_0+2\sqrt{I_0+I_0^2} \cos(2t)}.
502: \end{equation}
503: Figure \ref{fig2} shows that this approximation is indeed 
504: very good for the orbit $a(t)$
505: whatever the initial conditions lying in a closed orbit
506: with energy $< E_1$.
507: 
508: 
509: \begin{figure}
510: \begin{center}
511: \begin{tabular}{cc}
512: {\bf (a)}
513: \includegraphics[width=7.cm]{fig2a.eps}
514: &
515: {\bf (b)}
516: \includegraphics[width=7.cm]{fig2b.eps}
517: %{\bf (c)}
518: %\includegraphics[width=7.cm]{fig2c.eps}
519: %&
520: %{\bf (d)}
521: %\includegraphics[width=7.cm]{fig2d.eps}
522: \end{tabular}
523: \end{center}
524: \caption{Unperturbed dynamics of the BEC width.
525: We assume $P=0.1$, $a_t(0)=0$, $a(0)=2$ (picture a),
526: $a(0)=5.7$ (picture b).
527: The second case corresponds to an energy very close to $E_1$.
528: The results from the resolution of the ODE are compared
529: with the asymptotic formula (\ref{dyna1}).
530: \label{fig2}
531: }
532: \end{figure}
533: 
534: 
535: \subsection{Effective equations in presence of perturbations}
536: In case of small nonlinearity $P \ll 1$, the 
537: above expansions allow us to derive simple effective equations
538: for the BEC action in presence of perturbations.
539: Applying the general results obtained in Section \ref{sec:pert1},
540: we get that the action $I(t)$ behaves like a diffusion process 
541: with the infinitesimal generator
542: $$
543: {\cal L}_{{I}}= \frac{1}{2} \alpha_c \frac{\partial}{\partial {{I}}}
544: \left[
545: ({{I}}+{{I}}^2) \frac{\partial}{\partial {{I}}}
546: \right]
547: $$
548: where
549: $$
550: \alpha_c = \int_0^\infty \cos(2t) \EE[ \eta(0) \eta(t) ]dt.
551: $$
552: The expression of ${\cal L}_{{I}}$ holds true only for 
553: $ {I}< I_1$.
554: We can compute the growths of the first moments of 
555: the action starting from the ground state $I=0$
556: while $ e^{\alpha_c t} \ll P^{-2}$:
557: \begin{eqnarray}
558: \label{momi1}
559:   && \EE_0[{I}(t)] = \frac{1}{2} e^{\alpha_c t} - \frac{1}{2} ,\\
560: \label{momi2}
561:   && \EE_0[{I}(t)^2 ] = \frac{1}{6} e^{3\alpha_c t} -
562:  \frac{1}{2} e^{\alpha_c t}  + \frac{1}{3} .
563: \end{eqnarray}
564: An empirical way to estimate the mean disintegration time
565: is to look for the time $t_1$ such that $\EE_0[I(t_1)]=I_1$,
566: where $I_1=1/(12 P^2)$.
567: From Eq.~(\ref{momi1})  we get $t_1 = (1/\alpha_c) 
568: \ln[1+1/(6P^2)]$.
569: This argument is rough because the expectations are ill-placed.
570: The exact results provided by the rigorous stochastic analysis
571: confirm that this prediction is not correct.
572: Integrating Eqs.~(\ref{feller1}-\ref{feller2}) we get 
573: that the expectation of the disintegration time
574: starting from the ground state $I=0$ is
575: \begin{eqnarray}
576: \label{exptau}
577: \EE_0[ T_h ] &=&
578: \frac{2}{\alpha_c} \ln (1 + \frac{1}{12 P^2} )\\ 
579: \nonumber
580: &\stackrel{P \ll 1}{\simeq}& \frac{2}{\alpha_c}
581: \left( - 2 \ln(P)- \ln(12)  \right) ,
582: \end{eqnarray}
583: while its variance is
584: \begin{eqnarray}
585: \label{vartau}
586: {\rm Var}_0(T_h) &=& %\EE_0[ T_h^2] - \EE_0[ T_h]^2 =
587: \frac{8}{\alpha_c^2} \left[ \ln (1 + \frac{1}{12 P^2} ) 
588: + {\rm dilog}(1 + \frac{1}{12 P^2} ) + \frac{1}{2} 
589: \ln(1  + \frac{1}{12 P^2} )^2 \right]\\
590: \nonumber
591: &\stackrel{P \ll 1}{\simeq} 
592: & \frac{8}{\alpha_c^2}
593: \left[  -2 \ln(P) - \ln(12) -\frac{\pi^2}{6}    \right],
594: \end{eqnarray}
595: where the dilogarithm function is the tabulated function defined as follows:
596: $$
597: {\rm dilog}(x) = \int_1^x \frac{\ln(y)}{1-y} dy .
598: $$
599: Equations (\ref{exptau}-\ref{vartau}) are the most important results
600: of this paper. They show that the collapse time varies as 
601: $\sim \ln (P^{-2})$, while the energy barrier is $\sim P^{-2}$.
602: In physical variables, the expected collapse time is
603: $$
604: \EE_0[ T_c ] = \frac{2}{\omega_0 \alpha} \ln\left(1 +
605: \frac{\hbar \pi}{24 m \omega_0 a_s^2 N^2} \right), \ \ \ \ 
606: \alpha= \omega_0 \int_0^\infty \cos(2\omega_0 t) \EE[\eta(0) \eta
607: (t)] dt.
608: $$
609: Taking the experimental data $\omega_{0} = 10$kHz, $N \simeq 5\cdot
610: 10^{3}$, $a_{s}= -5$nm, and $\alpha = 10^{-4}-10^{-5}$,
611: we obtain the expected collapse time $\approx (1-10)$ seconds.
612: 
613: \subsection{Numerical simulations}
614: \label{subsec:num1}%
615: We compare the theoretical predictions 
616: with numerical simulations of the ODE (\ref{ode1}).
617: We use a fourth-order Runge-Kutta method for the 
618: resolution of the ODE.
619: The random fluctuations
620: are modeled by a stepwise constant random process:
621: $$
622: \eta(t) = \sigma \sum_{j} X_j {\bf 1}_{[ j t_c, (j+1)t_c)}(t) ,
623: $$
624: where the $X_j$ are independent and identically distributed 
625: random variables with uniform distribution  over $(- {1}/2
626: , {1}/2 )$ and $t_c$ is the coherence time of the laser.
627: The coefficient $\alpha_c$ is then given by
628: $$
629: \alpha_c = \sigma^2 \frac{1-\cos(2 t_c)}{48 t_c}  ;
630: $$
631: The first series of simulations were performed with the parameters
632: $\sigma=0.3$ and $t_c=0.5$.
633: We investigate different configurations corresponding to different
634: values of the parameter $P$
635: starting from $a(0)=1$, $a_t(0)=0$ which is very close to the ground
636: state.
637: We have carried $1000$ simulations for each configuration.
638: The theoretical values for the expected value 
639: and standard deviation according to formulas 
640: (\ref{exptau}-\ref{vartau}) are reported in Table \ref{table1}
641: and compared with the values obtained from averaging of the results
642: of the numerical simulations.
643: 
644: 
645: \begin{table}
646: \caption{Comparisons between the averages and rms of the 
647: collapse time obtained from numerical simulations and from theoretical
648: formulas. Here $\sigma=0.3$ and $t_c=0.5$.
649: \label{table1}}
650: \begin{tabular}{|c||c|c|c||c|c|c|}
651: \hline
652:  & \multicolumn{3}{c||}{  $\left< \tau \right>$ }
653:  &\multicolumn{3}{c|}{  ${\rm rms}(\tau)$ } \\
654: \cline{2-7}
655: \raisebox{1.5ex}[0cm][0cm]{$P$}
656: & num & theor & error & num & theor &  error\\
657: \hline
658: \hline
659: $0.05$ &  $4112$ &  $4103$ & $0.2\%$  & $2241$ &  $2335$ &
660: $4\%$\\
661: \hline
662: $0.1$ &   $2585$ &  $2591$ &$0.2\%$  & $1718$ & $1601$ &
663: $7\%$\\
664: \hline
665: $0.2$ &  $1257$ &  $1306$ & $3.5\%$  & $833$ & $865$ &
666: $4\%$ \\
667: \hline
668: $0.3$ &  $586$ &  $760$ & $23\%$  & $407$ & $518$ &
669: $21\%$ \\
670: \hline
671: $0.4$ &  $205$ &  $486$ & $58\%$  & $165$ & $336$ &
672: $51\%$\\
673: \hline
674: \end{tabular}
675: \end{table}
676: 
677: 
678: 
679: Note that the statistical formulas are theoretically valid in the
680: asymptotic framework $P \ll 1$.
681: The numerical simulations show that they are actually valid
682: for $P \leq 0.2$.
683: More exactly, the comparisons between the theoretical predictions
684: and the numerical simulations
685: shows excellent agreement for the mean values,
686: and very good agreement also for the standard deviations.
687: We also plot in Fig.~\ref{fig4} the histograms of the collapse times
688: for two series of simulations.
689: 
690: 
691: 
692: \begin{figure}
693: \begin{center}
694: \begin{tabular}{cc}
695: {\bf (a)}
696: \includegraphics[width=7.55cm]{fig3a.eps}
697: &
698: {\bf (b)}
699: \includegraphics[width=7.cm]{fig3b.eps}
700: \end{tabular}
701: \end{center}
702: \caption{Histograms of the collapse time 
703: obtained from series of $1000$ simulations.
704: Picture a: $P=0.1$. Picture b: $P=0.05$.
705: \label{fig4}
706: }
707: \end{figure}
708: 
709: \begin{table}
710: \caption{Comparisons between the averages and rms of the 
711: collapse time obtained from numerical simulations and from theoretical
712: formulas. Here $\sigma=2$ and $t_c=0.5$.
713: \label{table1b}}
714: \begin{tabular}{|c||c|c|c||c|c|c|}
715: \hline
716:  & \multicolumn{3}{c||}{  $\left< \tau \right>$ }
717:  &\multicolumn{3}{c|}{  ${\rm rms}(\tau)$ } \\
718: \cline{2-7}
719: \raisebox{1.5ex}[0cm][0cm]{$P$}
720: & num & theor & error & num & theor &  error\\
721: \hline
722: \hline
723: $0.05$ &  $98.6$ &  $92.3$ & $6.4\%$  & $55.5$ &  $52.5$ &
724: $5.4\%$\\
725: \hline
726: $0.1$ &   $63.7$ &  $58$ &$8.5\%$  & $39.1$ & $36.0$ &
727: $7.9\%$\\
728: \hline
729: $0.2$ &  $31.9$ &  $29.4$ & $7.8\%$  & $21.2$ & $19.5$ &
730: $8.2\%$ \\
731: \hline
732: $0.3$ &  $16.1$ &  $17.1$ & $6.5\%$  & $11.4$ & $11.7$ &
733: $2.6\%$ \\
734: \hline
735: $0.4$ &  $6.6$ &  $10.9$ & $65\%$  & $4.9$ & $7.6$ &
736: $55\%$\\
737: \hline
738: \end{tabular}
739: \end{table}
740: 
741: 
742: Finally, in Table \ref{table1b},
743: we report results with a high level of fluctuations
744: (namely $\sigma =2$).
745: The theoretical predictions are still in agreement with
746: the numerical
747: simulations for $P \leq 0.3$
748: with an accuracy of $10 \%$ although the considered
749: configurations are at the boundary of the validity
750: of the asymptotic theory.
751: 
752: \section{Critical nonlinearity}
753: \label{sec:crit}
754: \subsection{Expansions of the action-angle variables for critical
755:   nonlinearity}
756: In this section we address the case where the nonlinear
757: parameter $P$ is close to the critical value $P_c= 4/5^{5/4}$.
758: We do so by setting 
759: $P =P_c - \delta$ and assuming $\delta \ll 1$.
760: Once again, all quantities can be expanded in powers of
761: $\delta$. After some algebra, we get
762: \begin{eqnarray*}
763: &&  a_j = a_g + 2^{-1/2} 5^{-1/8} \tilde{a}_j \delta^{1/2} + O(\delta)
764: \mbox{ with } 
765: a_g = 5^{-1/4}, \ \ \
766: \tilde{a}_0=1, \ \ \
767: \tilde{a}_1=-1, \ \ \
768: \tilde{a}_2=2, \\
769: && E_j = E_g + 2^{1/2} 3^{-1} 
770: 5^{7/8} \tilde{E}_j \delta^{3/2}  + O(\delta^2)
771: \mbox{ with } 
772: E_g = 3^{-1} 5^{1/2} + 3^{-1} 5^{3/4} \delta ,\ \  \
773: \tilde{E}_0=-1, \ \ \
774: \tilde{E}_1=1.
775: %, \ \ \ \tilde{E}_2=1. 
776: \end{eqnarray*}
777: More generally,
778: if $a \in [a_1, a_2]$, then it can be parameterized as
779: $a = a_g + 2^{-1/2} 5^{-1/8} \delta^{1/2} \tilde{a}$
780: and the potential at $a$ can be expanded as
781: $$
782: U(a) = E_g + 2^{1/2} 3^{-1} 
783: 5^{7/8} \delta^{3/2} \tilde{U}( \tilde{a})   + O(\delta^2) ,
784: $$
785: where
786: $$
787: \tilde{U}(\tilde{a})  = \frac{1}{2}( \tilde{a}^3 - 3 \tilde{a} ) .
788: $$
789: Note that locally (i.e. around $a_g$) the potential presents a 
790: local minimum at $a_0$ (see Figure \ref{fig5}a),
791: but the shape of the potential well is very different from the
792: one observed in the framework $P \ll 1$ (compare with
793: Fig.~\ref{fig1}).
794: The width of the well $a_2-a_1$ is of the order $\sqrt{\delta}$ and 
795: its depth $E_1-E_0$ is of order $\delta^{3/2}$.
796: The local shape of the potential is given by the cubic function
797: $\tilde{U}$.
798: 
799: \begin{figure}
800: \begin{center}
801: \begin{tabular}{cc}
802: {\bf (a)}
803: \includegraphics[width=7.cm]{fig4a.eps}
804: &
805: {\bf (b)}
806: \includegraphics[width=7.cm]{fig4b.eps}
807: \end{tabular}
808: \end{center}
809: \caption{Picture a: Potential $U(a)$ for $P=P_c-0.01 \simeq 0.525$
810: ($\delta=0.01$).
811: Picture b:
812: Unperturbed dynamics of the BEC width.
813: We assume $a_t(0)=0$, $a(0)=0.66$, $\delta=0.01$.
814: The results from the resolution of the ODE are compared
815: with the asymptotic formula (\ref{dyna2}).
816: \label{fig5}
817: }
818: \end{figure}
819: 
820: We now consider the action-angle variables.
821: If $E \in [E_1,E_2)$, then it can be
822: parameterized as
823: $E =  E_g + 2^{1/2} 3^{-1} 
824: 5^{7/8} \tilde{E} \delta^{3/2}  $
825: with
826: $\tilde{E} \in [\tilde{E}_0, \tilde{E}_1) $.
827: There exist three solutions $e_3(\tilde{E}) \leq \tilde{a}_1 \leq
828: e_1(\tilde{E}) \leq e_2(\tilde{E})  \leq \tilde{a}_2$ of the cubic
829: equation
830: $\tilde{U}(\tilde{a}) = \tilde{E}$.
831: $e_1(\tilde{E})$ and $e_2(\tilde{E})$ determine the extremities
832: of the orbit of the normalized width
833: $\tilde{a}$ for the normalized energy $\tilde{E}$
834: in case of unperturbed dynamics.
835: The cubic equation can be solved:
836: $$
837: e_j(\tilde{E}) = 2 \cos \left( \frac{{\rm arccos}(\tilde{E})+2\pi(j-2)}{3}
838: \right) .
839: $$
840: In particular, 
841: if $E=E_0$ (i.e. $\tilde{E}=\tilde{E}_0$), then 
842: $e_1(\tilde{E}_0) = e_2(\tilde{E}_0) =1$ (and $e_3(\tilde{E}_0) =-2$),
843: which corresponds to the ground state $a(t) \equiv a_0$,
844: or $\tilde{a}(t) \equiv 1$.
845: 
846: The period ${\cal T}(E)$ of the closed orbit 
847: at energy level $E$, as defined by (\ref{def:calt}),
848: can be expanded as well.
849: Introducing 
850: $$
851: \tilde{\cal T}(\tilde{E}) = 
852: \delta^{1/4} 2^{-1/4} 
853: 5^{9/16}{\cal T}( E_g + 2^{1/2}
854: 3^{-1} 5^{7/8} \tilde{E} \delta^{3/2} ) ,
855: $$ 
856: we get that 
857: $\tilde{\cal T}$ is at leading order with respect to $\delta$
858: a $O(1)$-function that can be expressed in terms of tabulated functions
859: $$
860: \tilde{\cal T} (\tilde{E}) = \frac{2 \sqrt{3}}{\sqrt{ e_2(\tilde{E}) -
861:      e_3(\tilde{E})}} K \left( 
862: \rho(\tilde{E}  ) 
863: \right),
864: $$
865: where 
866: $$
867: \rho(\tilde{E}) = \frac{  e_2(\tilde{E}) -  e_1(\tilde{E})}{ e_2(\tilde{E}) -
868:   e_3(\tilde{E})}
869: $$
870: and $K$ is the complete elliptic integral \cite[p. 590]{abra}.
871: We then define a normalized action $\tilde{\cal I}(\tilde{E})$ 
872: for $\tilde{E} \in [\tilde{E}_0,\tilde{E}_1]=[-1,1]$ by
873: $$
874: \tilde{\cal I} (\tilde{E} ) = 
875: \frac{1}{2\pi} \int_{-1}^{\tilde{E}} \tilde{\cal T} (s)ds .
876: $$
877: The function 
878: $\tilde{\cal I} :
879: [\tilde{E}_0 , \tilde{E}_1] \rightarrow [0 , \tilde{I}_1] $
880: is invertible.
881: Its inverse is denoted by
882: $\tilde{\cal E} :
883: [0 , \tilde{I}_1]\rightarrow
884: [\tilde{E}_0 , \tilde{E}_1] $
885: where $\tilde{I}_1 = 18/(5 \pi)$.
886: It is plotted in Fig.~\ref{fig:crit1}a.
887: We can see that $\tilde{\cal E}$ is roughly linear.
888: Similarly we can define the angle ${\cal \phi}( \tilde{E} ,
889: \tilde{a})$
890: and its inverse $\tilde{\cal A} ( \tilde{I} , \phi)$.
891: The function $\tilde{\cal A} :
892: [0 , \tilde{I}_1] \times [0,2\pi) \rightarrow
893: [\tilde{a}_1 , \tilde{a}_2] $
894: can be expressed in terms of Jacobian elliptic functions
895: \begin{equation}
896: \tilde{\cal A} ( \tilde{I} , \phi)=
897: {e}_1 ( \tilde{\cal E}(\tilde{I}) )
898: +
899: \left[ {e}_2 ( \tilde{\cal E}(\tilde{I}) )
900: -
901: {e}_1 ( \tilde{\cal E}(\tilde{I}) ) 
902: \right]
903: {\rm sn}^2 \left( \frac{K( \rho (\tilde{\cal E}(\tilde{I}))}{\pi} \phi,
904: \rho (\tilde{\cal E}(\tilde{I})) \right) ,
905: \end{equation}
906: where sn is the Jacobian sinus 
907: \cite[p. 589]{abra}.
908: In absence of perturbation the action is preserved and
909: the closed orbit of $\tilde{a}(t)$ for a normalized action 
910: $\tilde{I} \in [0, \tilde{I}_1)$
911: is given by
912: \begin{equation}
913: \label{dyna2}
914: \tilde{a}(t)  =
915: \tilde{\cal A} \left(\tilde{I}, \phi(t) \right) \ \mbox{ with } \
916: \phi(t) =- \delta^{1/4} 2^{-1/4} 3^{-1/2} 5^{9/16}
917: \frac{2\pi}{\tilde{\cal T}(\tilde{\cal E}(\tilde{I}))} .
918: \end{equation}
919: The true orbit is $a(t) = a_g + 2^{-1/2} 5^{-1/8} \delta^{1/2} 
920: \tilde{a}(t)$.
921: Figure \ref{fig5}b shows that this approximation (derived
922: in the asymptotic framework $\delta \ll 1$) is indeed 
923: reasonably good.
924: 
925: 
926: \subsection{Effective equations in presence of perturbations}
927: Following the strategy presented in Section \ref{sec:pert1},
928: we introduce the normalized action-angle variables
929: so that 
930: $\tilde{E}(t) = \tilde{\cal E}(\tilde{I}(t)) $ and
931: $\tilde{a}(t) = \tilde{\cal A}(\tilde{I}(t),\phi(t)) $.
932: While the energy of the BEC is below $E_1$, 
933: the orbit is closed. As soon as the energy reaches the energy level
934: $E_1$,
935: the BEC collapses in a time of order $1$ (w.r.t. $\sigma_\eta$).
936: We shall show that the hitting time for the energy level
937: $E_1$ is of order $\sigma_\eta^{-2}$, so the collapse time $T_c$
938: is imposed by the hitting time
939: $T_h$ defined as the first time $t$ such that $\tilde{I}(t)= \tilde{I}_1 $.
940: Here we rescale $\tau = \sigma_\eta^2 \delta^{-3/2} t$.
941: This normalization is chosen so that the random process $\eta$ appears
942: with the scales of a white noise in 
943: the differential equations
944: $$
945: \left\{
946: \begin{array}{l}
947: \displaystyle
948: \frac{ d \tilde{I}}{d\tau} = 
949: \frac{1}{\varepsilon} \eta(\frac{\tau}{\varepsilon^2})
950: \tilde{h}_\phi(\tilde{I},\phi) ,\\
951: \displaystyle
952: \frac{ d \phi}{d\tau} = - \frac{\delta^{1/4}} 
953: {\varepsilon^2} 2^{-1/4} 5^{9/16}  
954: \tilde{\omega} (I) 
955: - \frac{1}{\varepsilon} \eta(\frac{\tau}{\varepsilon^2}) 
956: \tilde{h}_{\tilde{I}} (\tilde{I},\phi) ,
957: \end{array}
958: \right.
959: $$
960: where $\varepsilon=\sigma_\eta \delta^{-3/4}$,
961: $\tilde{h}(\tilde{I},\phi)=2^{-5/4} 3\ 5^{-11/16} 
962: \tilde{\cal A}(\tilde{I},\phi)$,
963: and
964: $\tilde{\omega}( \tilde{I} ) = \frac{2 \pi }{\tilde{\cal T} 
965: ( \tilde{\cal E} (\tilde{I}))}$.
966: Note once again that $\tilde{h}$ and $\tilde{\omega}$
967: are smooth functions, and $\tilde{h}$  is periodic with respect to
968: $\phi$ with period $2  \pi$.
969: By applying a diffusion approximation theorem \cite{psv},
970: we get that $(\tilde{I}(t) )_{t \geq 0}$
971: behaves like a diffusion Markov process with 
972: the infinitesimal generator
973: $$
974: {\cal L}_{\tilde{I}}= \alpha \delta^{-3/2}
975: \frac{\partial}{\partial  \tilde{I}}
976: \tilde{A}( \tilde{I}) \frac{\partial}{\partial  \tilde{I}} 
977: $$
978: where
979: \begin{eqnarray*}
980:  \tilde{A}( \tilde{I} ) 
981: &=& 2^{-1/2} 3^2 5^{-11/8} 
982: \left[ \tilde{e}_2 ( \tilde{\cal E}(\tilde{I}) )
983: -
984: \tilde{e}_1 ( \tilde{\cal E}(\tilde{I}) ) 
985: \right]^2 \frac{K( \rho (\tilde{\cal E}(\tilde{I}))}{\pi^2}
986: \int_0^{K( \rho (\tilde{\cal E}(\tilde{I}))}
987: {\rm cn}^2
988: {\rm dn}^2
989: {\rm sn}^2 
990: \left(s,
991: \rho (\tilde{\cal E}(\tilde{I})) \right) ds ,
992:   \\
993: \alpha &=& \int_0^{\infty} \EE[\eta(0) \eta(t) ] dt  ,
994: \end{eqnarray*}
995: and dn and cn are two tabulated elliptic functions
996: \cite[p. 589]{abra}.
997: The conditions ensuring the diffusion-approximation are
998: $\delta \ll 1$,  $\sigma_\eta^2 \ll \delta^{3/2}$.
999: The diffusion coefficient $\tilde{A}(\tilde{I})$ is plotted in
1000: Fig.~\ref{fig:crit1}b. 
1001: 
1002: \begin{figure}
1003: \begin{center}
1004: \begin{tabular}{cc}
1005: {\bf (a)}
1006: \includegraphics[width=7.cm]{fig5a.eps}
1007: &
1008: {\bf (b)}
1009: \includegraphics[width=7.cm]{fig5b.eps}
1010: \end{tabular}
1011: \end{center}
1012: \caption{Functions $\tilde{I} \mapsto \tilde{\cal E}(\tilde{I})$
1013: (picture a) and $\tilde{I} \mapsto \tilde{A}(\tilde{I})$ (picture b).
1014: \label{fig:crit1}
1015: }
1016: \end{figure}
1017: 
1018: 
1019: Using the results reported in Section \ref{sec:pert1}
1020: we get the following recursive relation ($n \geq 1$)
1021: for the moments of the hitting time $T_h$
1022: \begin{eqnarray}
1023: \label{sysmom}
1024:  && \EE_{\tilde{I}} [T_h^n] = \frac{n \delta^{3/2} }{\alpha}
1025:   \int_{\tilde{I}}^{\tilde{I}_1} 
1026: \frac{\int_0^x \EE_y[T_h^{n-1}] dy}{\tilde{A}(x)}  dx ,
1027: \end{eqnarray}
1028: where $\tilde{I}_1 =18/(5 \pi)$.
1029: In dimensional variables, the result reads as follows. 
1030: Starting from the ground state $a_0$, 
1031: the expected value of the collapse time is
1032: \begin{equation}
1033: \label{exptau2}
1034: \EE_0 [ T_c ] = \frac{
1035: \left(P_c - P \right)^{3/2} }
1036: {\omega_0^2 \alpha} C_1 , 
1037: \end{equation}
1038: where $C_1$ is the constant
1039: $C_1 = \int_{0}^{\tilde{I}_1}
1040: \frac{x}{\tilde{A}(x)} dx$.
1041: By a numerical integration using {\sc Matlab} we have found $C_1 \simeq 8.5$.
1042: More generally, we have
1043: \begin{equation}
1044: \label{vartau2}
1045: \EE_0 [ T_c^n ] = \frac{\left(P_c - P
1046: \right)^{3n/2} }{\omega_0^{2n} \alpha^n} C_n,
1047: \end{equation} 
1048: where $C_n$ are constants obtained recursively from Eq.~(\ref{sysmom}).
1049: By a numerical integration we have found $C_2 \simeq 110$.
1050: 
1051: 
1052: \subsection{Numerical simulations}
1053: We compare the theoretical predictions 
1054: with numerical simulations of the ODE (\ref{ode1}).
1055: We use the same model as in Section \ref{subsec:num1}
1056: with the parameters $\sigma=0.025$ and $t_c=0.5$.
1057: We report in Table \ref{table2}
1058: the theoretical values for the expected value 
1059: and standard deviation according to formulas 
1060: (\ref{exptau2}-\ref{vartau2}) as well as 
1061: the values obtained from averaging of the results
1062: of the numerical simulations.
1063: The statistical formulas are theoretically valid in the
1064: asymptotic framework $\delta (=P_c-P) \ll 1$.
1065: The numerical simulations show that they are actually valid
1066: for $\delta \leq 0.03$.
1067: 
1068: \begin{table}
1069: \caption{Comparisons between the averages and rms of the 
1070: collapse time obtained from numerical simulations and from theoretical
1071: formulas.
1072: \label{table2}}
1073: \begin{tabular}{|c|c||c|c|c||c|c|c|}
1074: \hline
1075:  &
1076:  & \multicolumn{3}{c||}{  $\left< \tau \right>$ }
1077:  &\multicolumn{3}{c|}{  ${\rm rms}( \tau )$ } \\
1078: \cline{3-8}
1079: \raisebox{1.5ex}[0cm][0cm]{$P$}
1080: &
1081: \raisebox{1.5ex}[0cm][0cm]{$\delta$}
1082: & num & theor & error & num & theor &  error\\
1083: \hline
1084: \hline
1085: $0.525$ & $0.01$ & $651$  &  $653$ & $0.3\%$   & $447$ &  $472$ &
1086: $5 \%$\\
1087: \hline
1088: $0.515$ & $0.02$ &  $1754$ &  $1846$ & $5\%$  & $1240$ &  $1334$ &
1089: $7.5\%$\\
1090: \hline
1091: $0.505$ & $0.03$ &  $3175$ &  $3392$ & $7\%$  & $2217$ & $2451$ &
1092: $10.5\%$\\
1093: \hline
1094: $0.495$ & $0.04$ &  $4673$ &  $5222$ & $11.5\%$  & $3107$ & $3775$ &
1095: $21.5\%$ \\
1096: \hline
1097: \end{tabular}
1098: \end{table}
1099: 
1100: \section{Validation of the variational approach}
1101: \label{sec:num}
1102: The analysis carried out in this paper is based on the variational
1103: approach using a Gaussian ansatz.
1104: The Gaussian ansatz for the study of static and dynamic properties of
1105: trapped gases has been widely used (see for instance
1106: \cite{perez97,shi97,stoof97,parola98,abdullaev01}).
1107: The variational principle is shown in these papers
1108: to be a simple Lagrangian-based method that
1109: gives
1110: reasonable accurate ordinary differential equations approximations
1111: to the true dynamics for the solution of the GP equation.
1112: This method merely assumes Gaussian pulse shapes containing a fixed number 
1113: of free parameters and the Lagrangian form of the partial differential
1114: equation is used to obtain the parameter evolution equations.
1115: However it is a questionable approach
1116: because it is based on the a priori assumption that the solution of
1117: the PDE has a form which remains very close to the chosen ansatz.
1118: Accordingly it has to be checked carefully by full
1119: numerical simulations of the PDE.
1120: 
1121: 
1122: Numerical simulations of the stochastic GP equation with spherically
1123: symmetric trap is performed by Crank-Nicholson scheme. The
1124: absorbing boundary condition is employed to imitate the infinite
1125: domain size. This technique allows to prevent re-entering of
1126: linear waves emitted by the condensate under perturbation into the
1127: integration domain.
1128: We have first checked the variational approach
1129: for the unperturbed system.
1130: We have done so by inserting the Gaussian waveform with the 
1131: amplitude and width corresponding to a stationary point
1132: (as predicted by the variational approach)
1133: as an initial condition into the PDE (\ref{eq:schro1}).
1134: We have let the solution evolve in time and we have plotted 
1135: the results in Fig.~\ref{fignum1}a.
1136: As can be seen the Gaussian ansatz is a good approximation 
1137: when $P$ is not close to the critical value $P_c$.
1138: Actually we have found numerically that the critical value for
1139: the existence of the BEC is not $P_c= 0.535$,
1140: as predicted by variational approximation,
1141: but $P_c = 0.459$.
1142: For $P$ very close to the real value of $P_c$,
1143: the Gaussian ansatz  substantially
1144: deviates from the exact solution of the 3D GP
1145: equation, as shown in Fig.~\ref{fignum1}b.
1146: 
1147: In a second step we have performed numerical simulations
1148: of the GP equation (\ref{eq:schro1}) driven by a random Gaussian white
1149: noise $\eta$ with zero-mean and autocorrelation
1150: function $\EE[ \eta(t) \eta(t') ] = \sigma^2 \delta(t-t')$.
1151: We do so by choosing randomly and independently the value of $\eta$ 
1152: at each time step.
1153: The mean
1154: collapse time is calculated as an average over 100 realizations
1155: of random paths along which the width of the condensate evolves
1156: from the value corresponding to the minimum of the effective
1157: potential $a_0$ until the value corresponding to its local
1158: maximum $a_1$ (see Fig.~\ref{fig1}). The initial wave-form is selected as a
1159: Gaussian with parameters predicted by the variational
1160: approximation corresponding to the stationary state of the
1161: condensate. 
1162: Fig.~\ref{fignum2}a represents the collapse time for 
1163: different values of the parameter $P$
1164: which are not too close to the critical value $P_c$.
1165: Comparison with the results from numerical 
1166: simulations of the ODE (\ref{eq:ode1})
1167: shows a very good agreement.
1168: This demonstrates that the variational approach provides accurate
1169: predictions for the behavior of the BEC.
1170: for small non-linearity, and that the asymptotic analysis carried out
1171: in Section \ref{sec:smallnl} holds true for the randomly driven GP
1172: equation.
1173: 
1174: \begin{figure}
1175: \begin{center}
1176: \begin{tabular}{cc}
1177: {\bf (a)}
1178: \includegraphics[width=7.cm]{fig6a.eps}
1179: &
1180: {\bf (b)}
1181: \includegraphics[width=5.7cm]{fig6b.eps}
1182: \end{tabular}
1183: \end{center}
1184: \caption{
1185: \label{fignum1}
1186: Picture a:
1187: Width of the BEC for an initial Gaussian waveform with parameters
1188: corresponding to a stationary point of the potential $U(a)$.
1189: The oscillations are insignificant for small values of $P$, and
1190: become important when $P$ approaches the critical value $P_c =0.459$.
1191: At overcritical $P$ the waveform rapidly shrinks ($a \rightarrow 0$),
1192: i.e. the BEC undergoes collapse.
1193: Picture b: Exact solution of the 3D GP equation (solid line) compared with 
1194: the Gaussian approximation with the same number of atoms and $P=0.44$.}
1195: \end{figure}
1196: 
1197: \begin{figure}
1198: \begin{center}
1199: \begin{tabular}{cc}
1200: {\bf (a)}
1201: \includegraphics[width=7.cm]{fig7a.eps}
1202: &
1203: {\bf (b)}
1204: \includegraphics[width=7.cm]{fig7b.eps}
1205: \end{tabular}
1206: \end{center}
1207: \caption{
1208: \label{fignum2}
1209: Mean collapse time calculated from
1210: stochastic PDE simulations (solid squares) and compared
1211: with the corresponding stochastic ODE simulations (open circles).
1212: Each mean is computed by averaging over a series of $100$ simulations.
1213: Picture a: 
1214: Mean collapse time as a function of $P$ for a white noise strength
1215: $\sigma=0.3$.
1216: Picture b:
1217: Mean collapse time as a function of $\sigma$ for a nonlinear parameter
1218: $P=0.44$ close to the critical value $P_c=0.459$.
1219: }
1220: \end{figure}
1221: 
1222: Finally, we have performed numerical simulations of the GP equation
1223: (\ref{eq:schro1}) driven by a white noise $\eta$
1224: with a nonlinear parameter $P$ very close to the critical
1225: value $P_c=0.459$.
1226: For near-critical values of the parameter $P$ the
1227: Gaussian waveform was found to be not enough accurate. In this
1228: case we employed the exact solution of the GP equation to initiate random
1229: simulations. The exact solution (ground state) of the GP equation is
1230: found by imaginary time-evolution method as described in
1231: \cite{chiofalo}.
1232: It is plotted in Fig.~\ref{fignum1}b. 
1233: The results are plotted in Fig.~\ref{fignum2}b.
1234: We can see that collapse in the perturbed PDE
1235: occurs much earlier than in the ODE model.
1236: This shows that the BEC in full GP equation is unstable against collapse at
1237: near critical nonlinear parameter.
1238: A small perturbation can drive the BEC to collapse through
1239: fluctuations that are not captured by the variational approach.
1240: Accordingly, we can state that the variational approach provides poor
1241: predictions for the behavior of the BEC
1242: for critical non-linearity.
1243: Several reasons can explain the departure:
1244: 1) the Gaussian ansatz is not correct (see Fig.~\ref{fignum1}b).
1245: 2) the study of the ODE model shows that the important
1246: parameter in the near-critical case is not
1247: the value of $P$, but the value of the difference between
1248: $P$ and $P_c$.
1249: But the ODE does not capture the correct value of $P_c$,
1250: so the error committed in the evaluation of the 
1251: difference $P-P_c$ becomes very large when $P$ becomes close to $P_c$.
1252: 3) radiation effects become very important, in the sense that the 
1253: waveform is strongly affected, even when the simulations 
1254: are performed starting from 
1255: the exact numerical waveform plotted in Fig.~\ref{fignum1}b,
1256: so that we feel that it is useless to try to find a more suitable ansatz.
1257: In this respect, one should add that this result is not surprising 
1258: because it is well known in nonlinear optics
1259: that the time-dependent variational approach fails to describe
1260: the regime near the collapse \cite{kuznetsov,berge}.
1261: Finally, it is necessary to mention that
1262: the behavior of the gas close to collapse can be affected
1263: by mechanisms that are not included in the GP equation,
1264: such as inelastic two and three-body collisions 
1265: \cite{edwards96,fedichev96}. 
1266: 
1267: 
1268: \section{Conclusion}
1269: We have considered in this paper a condensate  trapped
1270: by an external potential generated by a system 
1271: of laser beams
1272: in the case of a negative scattering length.
1273: We have studied the stability of the metastable BEC
1274: against small
1275: fluctuations of the laser intensity.
1276: We have shown that collapse of the BEC occurs whatever
1277: the amplitude of the fluctuations after a time which is inversely
1278: proportional to the integrated covariance of the 
1279: autocorrelation function of the fluctuations of the laser intensity.
1280: The statistical distribution of the collapse time has been computed.
1281: The dependence of the mean collapse time with respect to the number atoms $N$
1282: has been thoroughly analyzed.
1283: We have shown that, for $N$ below the critical number of atoms
1284: $N_c$, the mean collapse time decreases logarithmically with 
1285: increasing $N$.
1286: As a byproduct of the analysis we have shown that the variational approach
1287: is very efficient for the analysis of the BEC for a number of atoms $N$
1288: which is not too close to $N_c$, but we have seen that it completely fails
1289: for $N$ close to $N_c$.
1290: 
1291: 
1292: \section{Acknowledgments}
1293: F. Kh. A. and B. B. B. are grateful to the Fund of fundamental
1294: researches of the Uzbekistan
1295: Academy of Sciences for the partial financial support.
1296: B.B.B. also thanks the Physics Department of the
1297: University of Salerno, Italy, for a research grant.
1298: 
1299: 
1300: \begin{thebibliography}{99}
1301: 
1302: \bibitem{davis95}
1303: K. B. Davis, M. O. Mewes, M. R. Andrews, N. J. van Druten,
1304: D. S. Durfee, D. M. Kurn, and W. Ketterle,
1305: Phys. Rev. Lett. {\bf 75}, 3969 (1995).
1306: 
1307: \bibitem{anderson96}
1308: M. H. Anderson, J. R. Ensher, M. R. Mathews, 
1309: C. E. Wieman, and E. A. Cornell,
1310: Science {\bf 269}, 198 (1995).
1311: 
1312: \bibitem{bradley97}
1313: C. C. Bradley, C. A. Sackett,
1314: and R. G. Hulet, Phys. Rev. Lett. {\bf 78}, 985 (1997).
1315: 
1316: \bibitem{dalfovo}
1317: F. Dalfovo, S. Giorgini, L. P. Pitaevskii, and S. Stringary,
1318: Rev. Mod. Phys. {\bf 71}, 463 (1999). %-512
1319: 
1320: \bibitem{nozieres}
1321: P. Nozi{\`e}res and D. Pines,
1322: {\it The theory of quantum liquids}
1323: (Addison Wesley, Reading, 1990).
1324: 
1325: \bibitem{roberts01}
1326: J. L. Roberts, N. R. Claussen, S. L. Cornish, E. A. Donley,
1327: E. A. Cornell, and C. E. Wieman,
1328: Phys. Rev. Lett. {\bf 86}, 4211 (2001).
1329: 
1330: \bibitem{savage03}
1331: C. M. Savage, N. P. Robins, and J. J. Hope,
1332: Phys. Rev. A {\bf 67}, 014304 (2003).
1333: 
1334: \bibitem{donley}
1335: E. A. Donley, N. R. Claussen, S. L. Cornish, J. L. Roberts,
1336: E.A. Cornell, and C.E. Wieman,
1337: Nature, {\bf 412}, 295, (2001).
1338: 
1339: \bibitem{grimm}
1340: R. Grimm, M. Weidem{\"u}ller and Y. B. Ovchinnikov, Adv. At. Mol.
1341: Opt. Phys. {\bf 42}, 95 (2000).
1342: 
1343: \bibitem{takasu}
1344: Y. Takasu, K. Maki, K. Komori, T. Takano, K. Honda, M. Kumakura,
1345:  T. Yabuzaki, and Y. Takahashi,
1346: Phys. Rev. Lett. {\bf 91}, 040404 (2003).
1347: 
1348: \bibitem{savard}
1349: T. A. Savard, K. M. O'Hara, and J. E. Thomas, Phys. Rev. A {\bf
1350: 56} R1095, (1997).
1351: 
1352: \bibitem{hanggi}
1353: P. H{\"a}nggi, P. Talkner, and M. Borkovec,
1354: Rev. Mod. Phys. {\bf 62}, 251 (1990).
1355: 
1356: \bibitem{leggett}
1357: M. Ueda and A. J. Leggett,
1358: Phys. Rev. Lett. {\bf 80}, 1576 (1998).
1359: 
1360: \bibitem{gehm}
1361: M. E. Gehm, K. M. O'Hara, T. A. Savard, and J. E. Thomas, Phys.
1362: Rev. A {\bf 58}, 3914 (1998).
1363: 
1364: \bibitem{ABP}
1365: F. Kh. Abdullaev, J. C. Bronski, and G. Papanicolaou, Physica D {\bf
1366:   135}, 369 (2000).
1367: 
1368: \bibitem{ABK}
1369: F. Kh. Abdullaev, B. B. Baizakov, and V. V. Konotop, 
1370: In {\it Nonlinearity and Disorder:
1371: Theory and Applications}, NATO Science Series, Vol. 45, 
1372: F. Kh. Abdullaev, O. Bang and M. P. Soerensen (Eds),
1373: (Kluwer, Dodrecht, 2001).
1374: 
1375: \bibitem{ABG}
1376: F. Kh. Abdullaev, J. C. Bronski, and R. Galimzyanov, arXive:cond-mat/0205464.
1377: 
1378: \bibitem{gross}
1379: E. P. Gross, Nuovo Cimento {\bf 20}, 454 (1961);
1380: J. Math. Phys. {\bf 4}, 195 (1963);
1381: L. P. Pitaevskii, Zh. Eksp. Teor. Fiz. {\bf 40}, 646 (1961) 
1382: [Sov. Phys. JETP {\bf 13}, 451 (1961)].
1383: 
1384: \bibitem{anderson83}
1385: D. Anderson, Phys. Rev. A {\bf 27}, 3135 (1983).
1386: 
1387: \bibitem{malomed}
1388: B.A. Malomed, Prog. Opt. {\bf 43}, 69 (2002).
1389: 
1390: \bibitem{perez97}
1391: V.M. P{\'e}rez-Garc{\'\i}a, H. Michinel, J.I. Cirac, M. Lewenstein, and
1392: P. Zoller, Phys. Rev. A {\bf 56}, 1424 (1997).
1393: 
1394: \bibitem{ruprecht}
1395: P. A. Ruprecht, M. J. Holland, K. Burnett, and M. Edwards,
1396: Phys. Rev. A {\bf 51}, 4704 (1995). 
1397: 
1398: \bibitem{psv}
1399: G. Papanicolaou and W. Kohler,
1400: Comm. Pure Appl. Math. {\bf 27}, 641 (1974). %-668
1401: 
1402: \bibitem{feller}
1403: W. Feller,
1404: {\it An introduction to probability theory and its applications}
1405: (Wiley, New York, 1971).
1406: 
1407: \bibitem{abra}
1408: M. Abramowitz and I. A. Stegun,
1409: {\it Handbook of mathematical functions}
1410: (Dover, New York, 1972).
1411: 
1412: \bibitem{shi97}
1413: H. Shi and W.-M. Zheng, Phys. Rev. A {\bf 55}, 2930 (1997).
1414: 
1415: \bibitem {stoof97}
1416: H. T. C. Stoof,  J. Stat. Phys. {\bf 87}, 1353 (1997).
1417: 
1418: \bibitem{parola98}
1419: A..Parola, L. Salasnich, and L. Reatto, 
1420: Phys. Rev. A {\bf 57}, R3180 (1998). 
1421: 
1422: \bibitem{abdullaev01}
1423: F. Kh. Abdullaev, A. Gammal, Lauro Tomio, and T. Frederico,
1424: Phys. Rev. A {\bf 63}, 043604 (2001).
1425: 
1426: \bibitem{chiofalo}
1427: M. L. Chiofalo, S. Succi, and M. P. Tosi, Phys. Rev. E {\bf 62},
1428: 7438 (2000).
1429: 
1430: \bibitem{kuznetsov}
1431: E. A. Kuznetsov, A. M. Rubenchik, and V. E. Zakharov,
1432: Phys. Rep. {\bf 142}, 103 (1986).
1433: 
1434: \bibitem{berge}
1435: L. Berg{\'e}, 
1436: Phys. Rep. {\bf 303}, 260 (1998).
1437: 
1438: \bibitem{edwards96}
1439: M. Edwards, P. A. Ruprecht, K. Burnett, R. J. Dodd, and C. W. Clark,
1440: Phys. Rev. Lett. {\bf 77}, 1671 (1996). 
1441: 
1442: \bibitem{fedichev96}
1443: P. O. Fedichev, M. W. Reynolds, and G. V. Shlyapnikov, 
1444: Phys. Rev. Lett. {\bf 77}, 2921 (1996). 
1445: 
1446: 
1447: %\bibitem{baym}
1448: %G. Baym and C. J. Pethick,
1449: %Phys. Rev. Lett. {\bf 76} (1996) 6.%-9
1450: 
1451: %%F. Kh. Abdullaev and R. A. Kraenkel 
1452: %%Phys. Rev. A {\bf 62} (2000) 023613.
1453: 
1454: %\bibitem{ripoll}
1455: %J.J.G. Ripoll and V.M. Perez-Garcia, Phys. Rev. A 59, 2220 (1999);
1456: %J.J. Garcia-Ripoll, V.M. Perez-Garcia, and P. Torres,
1457: %Phys. Rev. Lett. 83, 1715 (1999).
1458: 
1459: %\bibitem{gaididei}
1460: %Y.B. Gaididei, K.O. Rasmussen, and P.L. Christiansen, Phys. Rev. E 52,
1461: %2951 (1995);
1462: %Yu.B. Gaididei, J. Schjodt-Eriksen, and P.L. Christiansen, 60,
1463: %4877 (1999).
1464: 
1465: \end{thebibliography}
1466: 
1467: \end{document}
1468: 
1469:  
1470: 
1471: 
1472: