0809.2007/ms.tex
1: \pdfoutput=1
2: %%
3: %% This is file `template-6s.tex',
4: %% generated with the docstrip utility.
5: %%
6: %% The original source files were:
7: %%
8: %% template.raw  (with options: `6s')
9: %% 
10: %% Template for the LaTeX class aipproc.
11: %% 
12: %% (C) 1998,2000,2001 American Institute of Physics and Frank Mittelbach
13: %% All rights reserved
14: %% 
15: %%
16: %% $Id: template.raw,v 1.12 2005/07/06 19:22:14 frank Exp $
17: %%
18: 
19: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
20: %% Please remove the next line of code if you
21: %% are satisfied that your installation is
22: %% complete and working.
23: %%
24: %% It is only there to help you in detecting
25: %% potential problems.
26: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
27: 
28: \input{aipcheck}
29: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
30: %% SELECT THE LAYOUT
31: %%
32: %% The class supports further options.
33: %% See aipguide.pdf for details.
34: %%
35: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
36: 
37: \documentclass[
38:     ,final            % use final for the camera ready runs
39: %%  ,draft            % use draft while you are working on the paper
40:     ,numberedheadings % uncomment this option for numbered sections
41: %%  ,                 % add further options here if necessary
42:   ]
43:   {aipproc}
44: 
45: \layoutstyle{6x9}
46: 
47: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
48: %% FRONTMATTER
49: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
50: 
51: \begin{document}
52: 
53: \title
54: {Nonlinear Dynamics of Bose-Einstein Condensates with Long-Range Interactions}
55: 
56: \classification{03.75.Kk, 34.20.Cf, 02.30.-f, 47.20.Ky}
57: \keywords      {Bose-Einstein condensation, Gross-Pitaevskii equation, nonlinear dynamics effects}
58: 
59: \author{G. Wunner}{
60:   address={Institut f\"ur Theoretische Physik 1, Universit\"at Stuttgart, 70550 Stuttgart, Germany}
61: }
62: \author{H. Cartarius}{
63:   address={Institut f\"ur Theoretische Physik 1, Universit\"at Stuttgart, 70550 Stuttgart, Germany}
64: }
65: \author{T.  Fab\v ci\v c}{
66:   address={Institut f\"ur Theoretische Physik 1, Universit\"at Stuttgart, 70550 Stuttgart, Germany}
67: }
68: \author{P. K\"oberle}{
69:   address={Institut f\"ur Theoretische Physik 1, Universit\"at Stuttgart, 70550 Stuttgart, Germany}
70: }
71: \author{J. Main}{
72:   address={Institut f\"ur Theoretische Physik 1, Universit\"at Stuttgart, 70550 Stuttgart, Germany}
73: }
74: \author{T. Schwidder}{
75:   address={Institut f\"ur Theoretische Physik 1, Universit\"at Stuttgart, 70550 Stuttgart, Germany}
76: }
77: 
78: %
79: %\author{<author2>}{
80: %  address={<common address for author2 and author3>}
81: %}
82: %
83: %\author{<author3>}{
84: %  address={<common address for author2 and author3>}
85: %  ,altaddress={<author1 address>} % additional visiting address
86: %}
87: 
88: 
89: \begin{abstract}
90: The motto of this paper is: Let's face Bose-Einstein condensation
91: through nonlinear dynamics. We do this by choosing variational forms
92: of the condensate wave functions (of given symmetry classes),
93: which convert the Bose-Einstein condensates via the time-dependent 
94: Gross-Pitaevskii equation into Hamiltonian systems that can be studied using the
95: methods of nonlinear dynamics. We consider in particular cold quantum
96: gases where  long-range interactions between the neutral atoms
97: are present, in addition to the conventional short-range contact 
98: interaction, viz. gravity-like interactions, and dipole-dipole interactions.
99: The results obtained serve as a useful guide in the search for 
100: nonlinear dynamics effects in numerically exact quantum calculations 
101: for Bose-Einstein
102: condensates. A main result is the prediction of the existence of
103: stable islands as well as chaotic regions for excited states of
104: dipolar condensates, which could be checked experimentally.
105: \end{abstract}
106: 
107: \maketitle
108: 
109: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
110: %% MAINMATTER
111: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
112: \section{Introduction}
113: 
114:  At sufficiently low temperatures
115: a condensate of weakly interacting bosons can be represented by a single wave function 
116: whose dynamics obeys the Gross-Pitaevskii equation \cite{gross61,pitaevskii61}. The Gross-Pitaevskii equation can be thought of as the Hartree equation for the ground state
117: of $N$ interacting identical bosons, all occupying the same single-particle orbital $\psi$.
118:  Because of its nonlinearity
119: the equation exhibits features not familiar from ordinary Schr\"odinger equations of quantum mechanics.
120: For example, Huepe et al. \cite{hue99,hue03} demonstrated that for Bose-Einstein condensates
121: with attractive contact interaction, described by a negative  $s$-wave scattering length, {\em bifurcations} 
122: of the stationary solutions of the Gross-Pitaevskii equation appear, and  determined both the stable
123: (elliptic) and the unstable (hyperbolic) branches of the solutions. The bifurcation points correspond to critical
124: particle numbers, above which, for given strength of the attractive interaction, collapse of the condensate sets in. In Bose-Einstein condensates of $^7$Li \cite{sacket99,gerton00}
125: and $^{85}$Rb atoms \cite{donley01,roberts01} these collapses were  experimentally observed.
126:  
127: In those condensates the short-range contact interaction is the only interaction to be considered. In Bose-Einstein
128: condensates of {\em dipolar} gases \cite{santos00,baranov02,goral02a,goral02b,giovanazzi03} also a long-range dipole-dipole interaction is present.  
129: Alternatively, following a proposal by O'Dell et al. \cite{dell00,giova01},
130: by using a combination
131: of 6 triads of appropriately tuned laser light condensates can be produced
132: in which an attractive long-range gravity-like $1/r$ interaction is present.
133: These types of condensates offer the 
134: opportunity to study degenerate quantum gases with adjustable long-range {\em and} short-range interactions. While the experimental realization of condensates
135: with gravity-like interaction lies still in the future, the
136: achievement of Bose-Einstein condensation in a gas of chromium atoms  \cite{griesmaier05}, with a large dipole moment, has opened the
137: way to promising experiments on dipolar gases  \cite{stuhler05}, which could show a wealth of novel phenomena \cite{giovanazzi02,santos03,li04,dell04}. In particular,
138: the experimental observation of the collapse of dipolar quantum gases has been reported  \cite{koch08} which occurs when 
139: the contact interaction is reduced, for a given particle number, below some critical value using a Feshbach resonance.
140: 
141: In this experimental situation it is most timely and appropriate to extend the investigations of the nonlinearity effects of the
142: Gross-Pitaevskii equation to quantum gases in which both the contact interaction 
143: and a long-range interaction is active, and this is the topic of the present paper.
144: 
145: \section{Scaling Properties of the Gross-Pitaevskii Equations with   Long-Range Interactions}
146: 
147: \subsection{Gravity-like interaction, isotropic trap}\label{scaling1/r}
148: 
149: For an additional gravity-like long-range interaction $ 
150: V_u(\vec r, \vec r^\prime \,) = - {u}/{|\vec r - \vec r^\prime |}$
151: the time-dependent Gross-Pitaevskii equation for the orbital $\psi$ reads
152: \begin{equation}\label{HFat}
153: \Big[ \hskip -1mm - \Delta + \gamma^2 r^2 + N 8 \pi \frac{a}{a_u} |\psi(\vec r, t)|^2 
154: - 2 N  \hskip -1mm  \int \frac{|\psi({\vec r\,}^\prime,t)|^2}{|{\vec r}- {\vec r\,}^\prime |}
155: d^3 {\vec r\,}^\prime \Big] \; \psi(\vec r, t) = i  \frac{\partial}{\partial t}  \, \psi(\vec r, t) \, , 
156: \end{equation}
157: where we have used \cite{Pap07} the ``Bohr radius'' $a_u = {\hbar^2}/{(m u)}$, the ``Rydberg energy''  
158: $E_u =  {\hbar^2}/{(2m a_u^2)}$, and the Rydberg time $\hbar/E_u$ as natural units
159: of length, energy, and time, respectively, to bring the equation in dimensionless form. 
160: Furthermore, in  (\ref{HFat}) $a$ is the  $s$-wave scattering length, which  characterizes the strength
161: of the contact interaction
162: $V_s = \delta({\vec r}- {\vec r\,}^\prime)\, 4 \pi a \hbar^2/m $,
163: $N$ is the particle number, and $\gamma = \hbar \omega_0/(2E_u)$ is the dimensionless trap frequency.
164: It can be shown \cite{Pap07} that the solutions of  (\ref{HFat}) do not depend on all these
165: three physical quantities but only on the two relevant parameters $\gamma/N^2$ and $N^2 a/a_u$. Thus
166: one has, e.g., for the mean-field energy  $E (N, N^2 a^\ast/a_u, \gamma/N^2)/N^3 = \,E(N=1, a/a_u, \gamma)$, with $ N^2 a^\ast/a_u =   a/a_u$.   
167: 
168: \subsection{Dipolar interaction, axisymmetric trap}\label{scaling}
169: 
170: In Bose condensates of $^{52}$Cr atoms \cite{griesmaier05}, which possess a large magnetic moment of $\mu = 6 \mu_{\rm B}$, the long-range dipole-dipole interaction 
171: \begin{equation}
172: V_{\rm dd}({\vec r}, {\vec r\,'}) =\frac{\mu_0 \mu^2}{4 \pi} \,\frac{1-3\cos^2 \theta'}{|{\vec r}-{\vec r}\,'|^3} \nonumber 
173: \end{equation}
174: must also be considered. Defining the dipole length by $a_{\rm d} = {\mu_0 \mu^2 m}/{(2\pi \hbar^2)}$,
175: and using as unit of energy $E_{\rm d} = \hbar^2/(2m a_{\rm d}^2)$, of frequency $\omega_{\rm d} = 
176: 2 E_{\rm d}/\hbar$ and of time $\hbar/E_{\rm d}$, one obtains the Gross-Pitaevskii equation for dipolar 
177: gases in axisymmetric traps in dimensionless form
178: \begin{eqnarray}\label{gpedd}{\Big[} \hskip -1mm - \Delta + \gamma_{\rho}^2 \rho^2 + \gamma_z^2 z^2
179:   +  N 8 \pi \frac{a}{a_{\rm d}} |\psi({\vec r},t)|^2
180: \hspace*{-2mm} &+& \hspace*{-2mm}
181:   N  \hskip -1mm  \int |\psi({{\vec r}\,}^\prime,t)|^2 
182: % \hspace*{-5mm}&& \hspace*{-5mm} 
183: \frac{(1 - 3 \cos^2
184:   \vartheta^\prime)}{|{{\vec r}}- {{\vec r}\,}^\prime |^3}
185:  d^3 {{\vec r}\,}^\prime
186:  \Big] \; \psi({\vec r},t) \nonumber \\  &=& \hspace*{-2mm}i \frac{\partial}{\partial t} \, \psi({\vec r},t) \, .
187: \end{eqnarray}
188: The physical parameters characterizing the condensates are the particle number $N$, the scattering length
189: $a/a_{\rm d}$ and the trap frequencies $\gamma_\rho$ and $\gamma_z$ perpendicular to and along the
190: direction of alignment of the dipoles (alternatively, the geometric mean 
191: (${\bar{\gamma}} = \gamma_\rho^{2/3} \gamma_z^{1/3}$ and the aspect ratio $\lambda = \gamma_z/\gamma_\rho$ 
192: is used). However, a closer inspection of the scaling properties of (\ref{gpedd}) reveals \cite{Koeb08}
193: that the solutions depend only on three parameters, viz. 
194:  $N^2 {\bar{\gamma}}, ~\lambda, ~a/a_{\rm d}$. For the mean-field energy, e.g., the scaling law
195: reads 
196: $E (N, a/a_{\rm d}, N^2{\bar{\gamma}}^\ast, \lambda)  \,=\,E(N=1, a/a_{\rm d}, 
197: {\bar{\gamma}}, \lambda)\, /N$, with $ N^2{\bar{\gamma}}^\ast = {\bar{\gamma}}$. 
198: 
199: \section{Quantum Results: Solutions of the Stationary Gross-Pitaevskii Equations}
200: 
201: For the $1/r$ interaction (monopolar quantum gases) we have solved  \cite{Pap07} the stationary Gross-Pitaevskii
202: equation both variationally, using an isotropic Gaussian-type orbital  $\psi = A \exp(-k^2 r^2/2)$,
203: and numerically accurate, by outward integration of the nonlinear Schr\"odinger equation.
204: For the dipole-dipole interaction (dipolar quantum gases) we have performed a variational calculation  \cite{Koeb08}
205: using an axisymmetric Gaussian-type orbital  $\psi = A \exp(-k_{\rho}^2 {\rho}^2/2-k_z^2 z^2/2)$.
206: 
207: Fig.~\ref{bifurcationmonopolar1} shows the results for  the chemical potential (eigenvalue of the stationary Gross-Pitaevskii 
208: equation) for the two interactions, plotted as a function of the scattering length. As can be seen, below a critical scattering
209: length no stationary solutions exist, while two stationary solutions are born at the critical scattering length in a tangent 
210: bifurcation. At the bifurcation point the chemical potential, the mean-field energy, and the wave functions of the 
211: two branches of solution are identical. Such behavior is obviously a consequence of the nonlinearity of the underlying
212: Schr\"odinger equation, and is reminiscent of exceptional points \cite{Hei99,Kato66} discussed so far only in the context of  
213: open quantum systems with non-Hermitian Hamiltonians (see Ref. \cite{Car08a} for references). In fact, a closer inspection shows \cite{Car08a} that the bifurcation
214: points can be identified as exceptional points: traversing circles around them in the 
215: complex-extended parameter plane, the
216: eigenvalues are permuted, which is a clear signature of exceptional points. 
217: \begin{figure}\label{bifurcationmonopolar1}
218: \caption{Bifurcations of the particle-number scaled chemical potential. Left: $1/r$ interaction, for different trap 
219: frequencies $\gamma$ (in units of $N^2$); solid curves: accurate numerical calculation, dashed curves: variational calculation. Right:
220: dipole-dipole interaction, variational results for the geometric mean trap frequency 
221: $N^2 \bar \gamma = 3.4 \times 10^4$ used in the experiments of Koch et al. \cite{koch08}
222: and different values of the trap aspect ratio $\lambda$.}
223: \includegraphics[width=\textwidth]{fig1}
224: \end{figure}
225: 
226: \section{Nonlinear Dynamics of Bose-Einstein Condensates with Atomic Long-Range Interactions}
227: Starting point of accurate numerical calculations are the time-dependent Gross-Pitaevskii equations 
228: (\ref{HFat}) and  (\ref{gpedd}). For variational calculations one makes use of the
229: fact that these equations follow from the variational principle
230: $||i \phi(t) -H \psi(t)||^2 ={\rm min}$, where the variation is performed with respect to $\phi$, and finally $\phi$ is set equal to $\dot \psi$. Using
231: a complex parametrization of the trial wave function  $\psi(t) = 
232: \chi({\vec \lambda}(t))$, the variation leads to equations of motion for the parameters $ {\vec \lambda}$ (cf. \cite{Fab07})
233: \begin{equation} \label{vareq}\textstyle{
234:  \left\langle \frac{\partial \psi}{\partial {\vec \lambda}}\Big|i \dot \psi - H \psi \right\rangle
235:  = 0  \leftrightarrow  
236: K \dot {\vec \lambda} = -i {\vec h} \hbox{~with~} K = \left\langle \frac{\partial \psi}{ \partial{{\vec \lambda}}}
237:  \Big|\frac{\partial \psi}{\partial{{\vec \lambda}}} \right\rangle, 
238:   {\vec h} = \left\langle \frac{\partial \psi}
239:  { \partial{{\vec \lambda}}}
240:  \Big|H \Big|\psi \right\rangle}\,.
241: \end{equation}
242: 
243: \subsection{Time evolution of condensates with $1/r$ interaction, variational and exact}\label{1/r}
244: For simplicity we consider the case of selftrapping, with no external trap.
245: As one can convince oneself, the results can be easily generalized
246: to the case where an external radially symmetric trap potential is
247: present. 
248: We choose a Gaussian trial wave function $\psi(r,t) = \exp\{ {i} [A(t) r^2 + \gamma(t)]\}$, where $A$ and $\gamma$ are complex functions of time, whose  
249: equations of motion follow from (\ref{vareq}). Decomposing $A$ into real and imaginary parts,  $A = A_r + i A_i $, and replacing them by  two other
250: dynamical quantities \cite{Bro89,Car08b}, $q = \sqrt{3/(4 A_i)} \equiv \sqrt{\langle r^2 \rangle}$, $ p = A_r\sqrt{3/A_i}$, converts those equations into
251: the canonical  equations of motion for $p$ and $q$ that follow from the 
252:  Hamiltonian
253: \begin{equation}\label{varham}
254:  E = H(q,p) = T+V = p^2+\frac{9}{4q^2}
255:  +\frac{3 \sqrt{3}a}{2\sqrt{\pi}q^3} -\frac{\sqrt{3}}{\sqrt{\pi}q}. \nonumber
256: \end{equation}
257: In this way the Gross-Pitaevskii equation is mapped onto the Hamiltonian of a one-dimensional
258: classical autonomous system with a nonlinear potential $V(q)$. 
259: The potential has no extremum for $a < a_{\rm cr} = -3\pi/8 \approx  -1.18$, possesses
260: a saddle point for $a = a_{\rm cr}$, and 
261: a maximum and a minimum for $a > a_{\rm cr}$. The critical scattering length
262: corresponds to the bifurcation point in the variational calculation.
263: For different values of $a$ (in units of $a_u$) phase portraits of trajectories moving according to the Hamiltonian (\ref{varham}) are shown in Fig.~\ref{phaseportraits}.
264: \begin{figure}\label{phaseportraits}
265: \caption{Phase portraits of the dynamics of the complex width function $A(t)$ associated with the 
266: Hamiltonian (\ref{varham}) for attractive $1/r$ interaction for different values of the
267: scattering length $a$, measured in units of $a_u$.
268: Left: $a = -1 > a_{\rm cr}$: two stationary states appear as fixed points; 
269: middle:  $a = -1.18 =  a_{\rm cr}$: coalescence of the fixed points; right:
270: $ a = -1.3 < a_{\rm cr}$: no stationary solutions exist.}
271: \includegraphics[width=\textwidth]{fig2}
272: \end{figure}
273: 
274: The linear stability analysis of both the variational and the exact quantum 
275: stationary solutions proves \cite{Car08b} that the state corresponding to the elliptical
276: fixed-point indeed is dynamically stable (small perturbations of the
277: state show oscillating behavior), while the stationary state 
278: corresponding to the hyperbolic fixed-point is dynamically unstable 
279: (exponential growth of small perturbations).
280: 
281: This behavior is recovered in exact numerical solutions of the 
282: time-dependent Gross-Pitaevskii equation with $1/r$ interaction, but also 
283: new features emerge. The solution is carried out \cite{Car08b} using
284: the split-operator technique and fast Fourier transforms. To investigate
285: the behavior of condensate wave functions in the vicinity of the
286: exact numerical stable and unstable stationary states, we consider 
287: condensates which are obtained by deforming the stationary states
288: by $\psi(r)  = f\cdot \psi_{\pm}(r\cdot f^{2/3})$, where $f$ is a 
289: numerical stretching factor (this choice of the perturbation does not 
290: affect the norm of the state).
291: 
292: In Fig.~\ref{wavefunctions} we show examples of the exact BEC dynamics
293: in the vicinity of the unstable and stable stationary states. 
294: In Fig.~\ref{wavefunctions} a) we start the time evolution with
295: the numerical solution for the unstable stationary state (in the classical picture this
296: corresponds to the trajectory starting at the hyperbolic fixed point,
297: see left part of Fig.~\ref{phaseportraits}). Because of unavoidable 
298: numerical deviations from the theoretically exact unstable state, the 
299:  wave function determined numerically
300:  is stationary only for some time but then
301: begins to oscillate. Obviously we have started with a state which
302: in the variational picture would be located in the elliptical domain
303: close to the hyperbolic fixed point. Note, however, that the
304: oscillation is not strictly periodic. By contrast, in 
305: Fig.~\ref{wavefunctions} b), where the time evolution starts 
306: with the unstable stationary state stretched by a factor of $f = 1.001$,
307: as time proceeds the wave function contracts towards the origin, 
308: and the condensate collapses. In the variational picture this corresponds
309: to a trajectory initially close to the hyperbolic fixed point but
310: located on the hyperbolic side. Note that in a realistic experimental
311: situtation during the  collapse further mechanisms have to be
312: taken into account, such as inelastic collisions. The inclusion of such
313: mechanisms, however, clearly goes beyond the scope of the present paper.
314: 
315: Figure~\ref{wavefunctions} c) displays a behavior not present in the variational
316: analysis. We start again in the vicinity of the unstable
317: stationary state ($f = 0.99$) and find that the width of the
318: condensate gradually grows and grows. An inspection of the 
319: wave function on a logarithmic scale shows \cite{Car08b} that
320: wave function amplitude builds up at large distances from 
321: the origin, giving rise to this behavior. Finally,
322: Fig.~\ref{wavefunctions} d), e)  show examples for the quantum mechanical
323: time evolution of condensates in the vicinity of the stable 
324: ground state. For a large stretching factor (panel d)) 
325: the condensate is found to oscillate and to expand, while
326: for a small stretching factor (panel e)) we find the quasiperiodic oscillations
327: that we would expect from the variational analysis. This demonstrates
328: that the variational nonlinear dynamics approach is capable of 
329: predicting essential features of the exact quantum mechanical 
330: time behavior of the condensates, but that the quantum mechanical
331: behavior is even richer.
332:    
333:     
334: \begin{figure}\label{wavefunctions}
335: \caption{Time evolution, for attractive $1/r$ interaction, of the root-mean-square widths of the condensate wave functions
336:  in the vicinity
337: of the unstable (panels a), b), c)) and the stable stationary (panels d), e)) 
338: state. a): Scaled scattering length (in units of $a_{\rm d}$) a = -1.0, stretching factor
339: f= 1.00; b): a = -0.85, f = -1.001; c): a = -0.85, f = 0.99; d): a = -0.85,
340: f =1.25,  and e): a = -0.85, f = 1.01. 
341: }
342: \includegraphics[width=1 \textwidth]{fig3}
343: \end{figure}
344: 
345: 
346: 
347: 
348: \subsection{Dynamics of condensates with dipolar interaction, variational}
349: We choose a Gaussian trial wave function adapted to the axisymmetric 
350: trap geometry,  $\psi(\rho,z,t) = e^{i(A_\rho \rho^2+ A_zz^2+\gamma)}$, where
351: the  complex width parameters $A_\rho$, $A_z$, and the complex phase
352: are functions of time. Their dynamical equations follow from the time-dependent
353: variational principle  (\ref{vareq}). Introducing new variables 
354: $q_\rho, q_z, p_\rho, p_z$ via
355: \begin{equation}
356: {\rm Re}\; A_\rho = {p_\rho}/{(4 q_\rho)},~~
357: {\rm Im}\; A_\rho = {1}/{(4 q_\rho^2)},~~
358: {\rm Re}\; A_z = {p_z}/{(4 q_z)},~~
359: {\rm Im}\; A_z = {1}/{(8 q_z^2)}
360: \end{equation}
361: one finds that their dynamical equations
362: are equivalent to the canonical equations of motion belonging to the
363: Hamiltonian 
364: \begin{eqnarray}\label{hamiltoniandd}
365: H &=& T+V =\frac{p_\rho^2}{2}+\frac{p_z^2}{2} \; +\frac{1}{2 q_\rho^2}+  2 \gamma_\rho^2 q_\rho^2+\frac{a/a_{\rm d}} {2 \sqrt{2\pi} q_\rho^2q_z} + \frac{1}{8 q_z^2}+2 \gamma_z^2 q_z^2\nonumber \\
366: &+&\frac{1+{q_\rho^2}/{q_z^2}-{3 q_\rho^2 \arctan \sqrt{{q_\rho^2}/{(2 q_z^2)}-1}}{\Big/}{\left(q_z^2 \sqrt{{2 q_\rho^2}/{q_z^2}-4}\right)}}{6 \sqrt{2 \pi} q_\rho^4 q_z\left({1}/{q_z^2}-{2}/{q_\rho^2}\right)}\,.
367: \end{eqnarray}
368: Thus the variational ansatz has turned the problem into one corresponding to 
369: a two-dimensional nonintegrable Hamiltonian system, which will exhibit all 
370: the features familiar from nonlinear dynamics studies of such systems. From 
371: the shape of the potential, which is shown in 
372: Fig.~\ref{vnumplot2} as a function of the "position'' variables
373: $q_\rho, q_z$, these features can  already  be read  off  qualitatively.
374: At the potential minimum sits the stable 
375: stationary ground state (elliptic fixed point), while at the saddle point 
376: one finds an unstable excited stationary state (hyperbolic fixed point).
377: %Up to the saddle point excitation energy one expects bound motion which 
378: %can be regular or chaotic, above that energy one expects trajectories
379: %that either remain in the potential well or escape over the saddle point.  
380: \begin{figure}\label{vnumplot2}
381: \caption{Potential $V(q_\rho, q_z)$ in the Hamiltonian (\ref{hamiltoniandd}) for
382: dipole-dipole long-range interaction.}
383: \includegraphics[width=0.55\textwidth]{fig4}
384: \end{figure}
385: To quantitatively characterize the dynamics of the variational condensate wave functions 
386:  we follow the trajectories  in the four-dimensional configuration space spanned 
387: by the coordinates of the real and imaginary parts of $A_\rho$ and $A_z$.
388: Since the total mean-field energy is a constant of motion the
389: trajectories are confined to three-dimensional hyperplanes, and their
390: behavior can  most conveniently be visualized by two-dimensional 
391: Poincar{\' e} surfaces of section defined by requiring one of the 
392: coordinates to assume a fixed 
393: %real 
394: value. 
395: 
396: We  consider
397: Poincar{\' e} surfaces of section defined by the condition that 
398: the imaginary part of  $A_z(t)$ is zero. Each time the trajectory 
399:  crosses the plane ${\rm Im}\;{A_z} = 0$, the 
400:  real and  imaginary
401: parts of $A_\rho(t) = A_\rho^r(t) + i A_\rho^i(t)$ are recorded. 
402: In Fig.~\ref{poincare} surfaces of section  are plotted for 
403: five different, increasing, values of
404: the mean-field energy. 
405: The physical parameters of the experiment of Koch
406: et al. \cite{koch08}  are adopted, and the scattering length is fixed to 
407: $a/a_{\rm d}=0.1$, away from its critical
408: value.
409: At these parameters, the variational mean-field energy of the ground state 
410: is $N E_{\rm gs} =4.24 \times 10^5$ (in units of $E_{\rm d}$) and represents the local minimum on the two-dimensional
411: mean-field energy landscape, plotted as a function of the width parameters. The variational energy
412: of the second, unstable, stationary state at these experimental parameters is $N E_{\rm es}=6.24 \times 10^5$, it corresponds to 
413: the saddle point on the mean-field energy surface. 
414: Between these two energy values the motion on the trajectories is 
415: bound, while for energies above the saddle-point energy
416: the motion on the trajectories can become unbound: once the saddle point is traversed by 
417: a trajectory $A_\rho(t)$, $A_z(t)$, the parameters run to infinity, 
418: meaning a shrinking of the quantum state to vanishing width, i~e., a collapse of the condensate takes place.
419: 
420: \begin{figure}
421: \centering\includegraphics[width=1\textwidth]{fig5}
422: \caption{Poincar{\' e} surfaces of section of the condensate wave functions 
423: for dipolar interaction represented by their width 
424: parameters at the scaled trap frequency $N^2 \bar{\gamma} = 3.4\times 10^{4}$, 
425:  aspect ratio $\lambda =6$, 
426: and the scattering length $a/a_{\rm d}=0.1$. The surfaces of section correspond to increasing values
427: of the mean-field energy (in units of $E_{\rm d}$): a) $NE = 4.5 \times 10^5$, b) $NE =  6.00 \times 10^5$, c) and d)  $NE =  6.24 \times 10^5$,
428: e)  $NE =  9 \times 10^5$, f)  $NE =  6.00 \times 10^6$.
429: } 
430: \label{poincare}
431: \end{figure}
432: 
433: 
434: 
435: The energy in Fig.~\ref{poincare} a) lies slightly above the energy of the stationary ground state.
436: The initially stationary state has evolved into a periodic orbit (fixed point in the surface of section), 
437: corresponding to
438: a state of the condensate whose motion  is periodic. The oscillations of the width parameters
439: $A_\rho(t)$ and $A_z(t)$ represent oscillatory stretchings of the condensate along the $\rho$ and $z$
440: directions.  The  stable {\em periodic} 
441: orbit in the surface of section is surrounded by elliptical, quasi-periodic orbits, representing quasi-periodic
442: oscillations of the condensate. As the energy is increased further, in Fig.~\ref{poincare} b), new periodic
443: and quasi-periodic orbits are born, and the motion is still regular.  In Fig.~\ref{poincare} c) we have reached 
444: the saddle-point energy. Now chaotic orbits have appeared in the vicinity of the unstable excited stationary state (hyperbolic fixed point). Figure~\ref{poincare} d) shows an enlargement of this 
445: region in phase space.
446: In contrast to the (quasi-) periodic stretching oscillations 
447: of the condensate within the elliptical islands, the chaotic motion of the parameters describes
448: a condensate which does not yet collapse but whose widths  fluctuate irregularly.
449: 
450: In the surfaces of section shown in Fig.~\ref{poincare} e) and f), with mean-field energies well above the saddle-point energy,   
451: regular islands are still clearly visible.
452: These stable islands are surrounded by chaotic trajectories. 
453: Since ergodic motion along  these trajectories comes 
454: close to every point in the configuration space, the chaotic motion sooner or later leads to a crossing of 
455: the saddle point and then to the collapse of the condensate wave functions.
456: It can be seen that with growing energy  
457: above the saddle point  the sizes of the stable regions shrink. The kinematically 
458: allowed regions surrounding
459: the stable islands are hardly recognizable any more since  high above 
460: the saddle-point energy the 
461: chaotic motion becomes more and more unbound, and thus trajectories cross the Poincar{\' e} surfaces
462: of section only a few times, if ever, before they escape to infinity and  collapse takes place. 
463: 
464: 
465: It must be stressed, however, that stable islands persist even far above the saddle-point energy, 
466: implying the existence of quasi-periodically oscillating  nondecaying modes of dipolar condensate wave functions.  
467: 
468: 
469: 
470: \section{Summary and Conclusions}
471: We have demonstrated that variational forms of the Bose-Einstein condensate wave functions
472: convert the condensates via the Gross-Pitaevskii equation into Hamiltonian systems that can 
473: be studied using the methods of nonlinear dynamics. We have also shown that these results
474: serve as a useful guide in the search for nonlinear dynamics effects in the numerically
475: accurate quantum calculations of Bose-Einstein condensates with long-range interactions.
476: The existence of stable islands as well as chaotic regions for excited states of 
477: dipolar Bose-Einstein condensates is a result that could be checked experimentally. 
478: One way of creating the collectively excited states one might think of  is
479: to prepare the condensate in the ground state, and then to non-adiabatically reduce the trap frequencies.
480: 
481: One might question
482: whether the Gross-Pitaevskii equation is adequate to describe the types
483: of complex dynamics 
484: discussed in this paper in ``real'' condensates. For example, in the chaotic regime 
485: local density maxima might occur for which losses by two-body  or three-body collisions 
486: would have to be taken into account. However, by virtue of the scaling laws 
487: discussed in Sections  \ref{scaling1/r}  and  \ref{scaling}
488: parameter ranges can always be found where the particle densities remain small even in 
489: these regimes, and the Gross-Pitaevskii equation is applicable.
490: 
491: 
492: The advantage of the simple variational ansatz  adopted in this paper is that
493: the analysis of the nonlinear dynamical properties of Bose-Einstein condensates 
494: becomes particularly transparent. Numerical quantum calculations to confirm the variational 
495: findings for dipolar gases and the extensions to structured condensate states 
496: \cite{goral00,dutta07} are under way. 
497: We have already seen in Sec.~\ref{1/r}, by 
498: comparing variational and accurate numerical quantum results for 
499:  Bose-Einstein  condensates with attractive
500: $1/r$ interaction, that  the nonlinear dynamical properties predicted 
501: by the variational calculation were confirmed by the full quantum calculations. 
502: We therefore have good reason to believe that this will also be true 
503: once the full quantum calculations of the dynamics of excited condensate 
504: wave functions of  dipolar gases have become available.
505: 
506: 
507: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
508: %% BACKMATTER
509: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
510: 
511: \begin{theacknowledgments}
512: Part of this work has been supported by Deutsche Forschungsgemeinschaft. H.C. 
513: is grateful for support from the Landesgraduiertenf\"orderung of
514: the Land Baden-W\"urttemberg.
515: \end{theacknowledgments}
516: 
517: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
518: %% The bibliography can be prepared using the BibTeX program or
519: %% manually.
520: %%
521: %% The code below assumes that BibTeX is used.  If the bibliography is
522: %% produced without BibTeX comment out the following lines and see the
523: %% aipguide.pdf for further information.
524: %%
525: %% For your convenience a manually coded example is appended
526: %% after the \end{document}
527: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
528: 
529: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
530: %% You may have to change the BibTeX style below, depending on your
531: %% setup or preferences.
532: %%
533: %%
534: %% For The AIP proceedings layouts use either
535: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
536: 
537: \bibliographystyle{aipproc}   % if natbib is available
538: %\bibliographystyle{aipprocl} % if natbib is missing
539: 
540: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
541: %% You probably want to use your own bibtex database here
542: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
543: %\bibliography{paper,bec_bank_1}
544: \begin{thebibliography}{32}
545: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
546: \providecommand{\enquote}[1]{``#1''}
547: \expandafter\ifx\csname url\endcsname\relax
548:   \def\url#1{\texttt{#1}}\fi
549: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
550: \providecommand{\eprint}[2][]{\url{#2}}
551: 
552: \bibitem[Gross(1961)]{gross61}
553: E.~P. Gross, \emph{Nuovo Cimento} \textbf{20}, 454 (1961).
554: 
555: \bibitem[Pitaevskii(1961)]{pitaevskii61}
556: L.~P. Pitaevskii, \emph{Sov. Phys. JETP} \textbf{13}, 451 (1961).
557: 
558: \bibitem[Huepe et~al.(1999)]{hue99}
559: C.~Huepe, S.~M\'etens, G.~Dewel, P.~Borckmans, and M.~E. Brachet, \emph{Phys.
560:   Rev. Lett.} \textbf{82}, 1616--1619 (1999).
561: 
562: \bibitem[Huepe et~al.(2003)]{hue03}
563: C.~Huepe, L.~S. Tuckerman, S.~M\'etens, and M.~E. Brachet, \emph{Phys. Rev. A}
564:   \textbf{68}, 023609 (2003).
565: 
566: \bibitem[Sacket et~al.(1999)]{sacket99}
567: C.~A. Sacket, J.~M. Gerton, M.~Welling, and R.~G. Hulet, \emph{Phys. Rev.
568:   Lett.} \textbf{82}, 876 (1999).
569: 
570: \bibitem[Gerton et~al.(2000)]{gerton00}
571: J.~M. Gerton, D.~Strekalov, I.~Prodan, and R.~G. Hulet, \emph{Nature}
572:   \textbf{406}, 692 (2000).
573: 
574: \bibitem[Donley et~al.(2001)]{donley01}
575: E.~A. Donley, N.~R. Claussen, S.~L. Cornish, , J.~L. Roberts, E.~A.Cornell, and
576:   C.~E. Wiemann, \emph{Nature} \textbf{412}, 295 (2001).
577: 
578: \bibitem[Roberts et~al.(2001)]{roberts01}
579: J.~L. Roberts, N.~R. Claussen, S.~L. Cornish, E.~A. Donley, E.~A. Cornell, and
580:   C.~E. Wieman, \emph{Phys. Rev. Lett.} \textbf{86}, 4211 (2001).
581: 
582: \bibitem[Santos et~al.(2000)]{santos00}
583: L.~Santos, G.~V. Shlyapnikov, P.~Zoller, and M.~Lewenstein,
584:   \emph{Phys.\,Rev.\,Lett.} \textbf{85}, 1791 (2000).
585: 
586: \bibitem[Baranov et~al.(2002)]{baranov02}
587: M.~Baranov, L.~Dobrek, K.~G{\'o}ral, L.~Santos, and M.~Lewenstein,
588:   \emph{Phys.\,Scr.} \textbf{T102}, 74 (2002).
589: 
590: \bibitem[G{\'o}ral et~al.(2002)]{goral02a}
591: K.~G{\'o}ral, L.~Santos, and M.~Lewenstein, \emph{Phys.\,Rev.\,Lett.}
592:   \textbf{88}, 170406 (2002).
593: 
594: \bibitem[G{\'o}ral and Santos(2002)]{goral02b}
595: K.~G{\'o}ral, and L.~Santos, \emph{Phys.\,Rev.\,A} \textbf{66}, 023613 (2002).
596: 
597: \bibitem[Giovanazzi et~al.(2003)]{giovanazzi03}
598: S.~Giovanazzi, A.~G{\"o}rlitz, and T.~Pfau, \emph{J.\,Opt.\,B.} \textbf{5},
599:   S208 (2003).
600: 
601: \bibitem[O'Dell et~al.(2000)]{dell00}
602: D.~O'Dell, S.~Giovanazzi, G.~Kurizki, and V.~M. Akulin,
603:   \emph{Phys.\,Rev.\,Lett.} \textbf{84}, 5687 (2000).
604: 
605: \bibitem[Giovanazzi et~al.(2001)]{giova01}
606: S.~Giovanazzi, D.~O'Dell, and G.~Kurizki, \emph{Phys.\,Rev.\,A} \textbf{63},
607:   031603(R) (2001).
608: 
609: \bibitem[Griesmaier et~al.(2005)]{griesmaier05}
610: A.~Griesmaier, J.~Werner, S.~Hensler, J.~Stuhler, and T.~Pfau,
611:   \emph{Phys.\,Rev.\,Lett.} \textbf{94}, 160401 (2005).
612: 
613: \bibitem[Stuhler et~al.(2005)]{stuhler05}
614: J.~Stuhler, A.~Griesmaier, T.~Koch, M.~Fattori, T.~Pfau, S.~Giovanazzi,
615:   P.~Pedri, and L.~Santos, \emph{Phys.\,Rev.\,Lett.} \textbf{95}, 150406
616:   (2005).
617: 
618: \bibitem[Giovanazzi et~al.(2002)]{giovanazzi02}
619: S.~Giovanazzi, A.~G{\"o}rlitz, and T.~Pfau, \emph{Phys.\,Rev.\,Lett.}
620:   \textbf{89}, 130401 (2002).
621: 
622: \bibitem[Santos et~al.(2003)]{santos03}
623: L.~Santos, G.~V. Shlyapnikov, and M.~Lewenstein, \emph{Phys.\,Rev.\,Lett.}
624:   \textbf{90}, 250403 (2003).
625: 
626: \bibitem[Yi et~al.(2004)]{li04}
627: S.~Yi, L.~You, and H.~Pu, \emph{Phys. Rev. Lett.} \textbf{93}, 040403 (2004).
628: 
629: \bibitem[O'Dell et~al.(2004)]{dell04}
630: D.~H.~J. O'Dell, S.~Giovanazzi, and C.~Eberlein, \emph{Phys. Rev. Lett.}
631:   \textbf{92}, 250401 (2004).
632: 
633: \bibitem[Koch et~al.(2008)]{koch08}
634: T.~Koch, T.~Lahaye, J.~Metz, B.~Fr{\"o}hlich, A.~Griesmaier, and T.~Pfau,
635:   \emph{Nature Physics} \textbf{4}, 218--222 (2008).
636: 
637: \bibitem[Papadopoulos et~al.(2007)]{Pap07}
638: I.~Papadopoulos, P.~Wagner, G.~Wunner, and J.~Main, \emph{Phys. Rev. A}
639:   \textbf{76} (2007).
640: 
641: \bibitem[K{\"o}berle et~al.(2008)]{Koeb08}
642: P.~K{\"o}berle, H.~Cartarius, T.~Fab{\v c}i{\v c}, J.~Main, and G.~Wunner,
643:   \emph{New J. Phys.}  (2008), submitted, Preprint: arXiv:0802.4055v2.
644: 
645: \bibitem[Heiss(1999)]{Hei99}
646: W.~D. Heiss, \emph{Eur. Phys. J. D} \textbf{7}, 1--4 (1999).
647: 
648: \bibitem[Kato(1966)]{Kato66}
649: T.~Kato, \emph{Perturbation theory for linear operators}, Springer, Berlin,
650:   1966.
651: 
652: \bibitem[Cartarius et~al.(2008{\natexlab{a}})]{Car08a}
653: H.~Cartarius, J.~Main, and G.~Wunner, \emph{Phys. Rev. A} \textbf{77}, 013618
654:   (2008{\natexlab{a}}).
655: 
656: \bibitem[Fab\v{c}i\v{c} et~al.(2007)]{Fab07}
657: T.~Fab\v{c}i\v{c}, J.~Main, and G.~Wunner, \emph{Nonlinear Phenomena in Complex
658:   Systems} \textbf{10}, 86--91 (2007).
659: 
660: \bibitem[Broeckhove et~al.(1989)]{Bro89}
661: J.~Broeckhove, L.~Lathouwers, and P.~V. Leuven, \emph{J. Phys. A} \textbf{22},
662:   4395--4408 (1989).
663: 
664: \bibitem[Cartarius et~al.(2008{\natexlab{b}})]{Car08b}
665: H.~Cartarius, T.~Fab{\v c}i{\v c}, J.~Main, and G.~Wunner, \emph{Phys. Rev. A}
666:   \textbf{78}, 013615 (2008{\natexlab{b}}).
667: 
668: \bibitem[G{\'o}ral et~al.(2000)]{goral00}
669: K.~G{\'o}ral, K.~Rz{\c a}{\.z}ewski, and T.~Pfau, \emph{Phys.\,Rev.\,A}
670:   \textbf{61}, 051601(R) (2000).
671: 
672: \bibitem[Dutta and Meystre(2007)]{dutta07}
673: O.~Dutta, and P.~Meystre, \emph{Phys.\,Rev.\,A} \textbf{75}, 053604 (2007).
674: 
675: \end{thebibliography}
676: 
677: \end{document}
678: