1: \documentclass[aps,twocolumn,pre]{revtex4}
2: \usepackage{graphicx}
3: \usepackage{amsfonts}
4: \usepackage{amssymb}
5: \usepackage{amsmath}
6:
7: \setcounter{MaxMatrixCols}{10}
8:
9: \def\wfig{4.25cm}
10: \def\hfig{3.0cm}
11: \begin{document}
12:
13: \title{\bf Symmetry Breaking in Symmetric and Asymmetric Double-Well Potentials}
14:
15: \author{
16: G.\ Theocharis$^{1}$,
17: P.G.\ Kevrekidis$^{2}$,
18: D.J.\ Frantzeskakis$^{1}$, and
19: P.\ Schmelcher$^{3,4}$
20: }
21: \affiliation{
22: $^{1}$ Department of Physics, University of Athens, Panepistimiopolis,Zografos, Athens 157 84, Greece \\
23: $^{2}$ Department of Mathematics and Statistics,University of Massachusetts, Amherst MA 01003-4515, USA \\
24: $^{3}$ Theoretische Chemie, Physikalisch-Chemisches Institut, INF 229, Universit\"at Heidelberg, 69120 Heidelberg, Germany \\
25: $^{4}$ Physikalisches Institut, Philosophenweg 12, Universit\"at Heidelberg, 69120 Heidelberg, Germany
26: }
27:
28: \begin{abstract}
29:
30: Motivated by recent experimental studies of matter-waves and optical beams
31: in double well potentials, we study the solutions of
32: the nonlinear Schr\"{o}dinger equation in such a context.
33: %examine analytically and numerically
34: %the nonlinear waves that emerge in such potentials.
35: Using a Galerkin-type approach,
36: we obtain a detailed handle on the nonlinear solution branches of the problem,
37: starting from the corresponding linear ones and predict the relevant bifurcations
38: of solutions for both attractive and repulsive nonlinearities. The results
39: %These analytical predictions are found to be in excellent agreement with the full numerical results and serve to
40: illustrate the nontrivial differences that arise between the
41: steady states/bifurcations emerging in symmetric and asymmetric double wells.
42:
43: \end{abstract}
44:
45: \maketitle
46:
47: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
48:
49: %\section{Introduction}
50:
51: {\it Introduction}.
52: %
53: It is well known that the nonlinear Schr\"{o}dinger (NLS) equation is a
54: fundamental
55: model describing the evolution of a complex field envelope in nonlinear dispersive media \cite{NLS}.
56: As such, it plays a key role in many different contexts, ranging from nonlinear and atom optics to
57: plasma physics, fluid dynamics, and even biophysical models \cite{trubatch}.
58: The interest in the NLS equation has dramatically increased during the last
59: few years, as it also describes the mean-field dynamics of Bose-Einstein condensates (BECs) \cite{gbec}.
60: In this context, the NLS is also known as the
61: Gross-Pitaevskii (GP) equation, and
62: typically incorporates external potentials that are used
63: for the BEC confinement. Such potentials may be, e.g., harmonic (usually implemented
64: by external magnetic fields) or periodic (implemented by the
65: interference of laser beams),
66: so-called optical lattices \cite{reviewsbec}.
67: Importantly, NLS models with
68: %such harmonic or periodic
69: similar external potentials appear also in the
70: context of optics, where they respectively describe the evolution of an optical beam in a
71: graded-index waveguide or in periodic waveguide arrays \cite{kivshar,reviewsopt}.
72:
73: Another type of external potential, which has mainly been studied
74: theoretically in the BEC context \cite{smerzi,kiv2,mahmud,bam,Bergeman_2mode,infeld}
75: is the double well potential.
76: Moreover, it has been demonstrated experimentally that a BEC
77: either tunnels and performs Josephson oscillations between the wells, or is
78: subject to macroscopic quantum self-trapping \cite{markus1}.
79: On the other hand, in the context of optics, a double well potential can be created
80: by a two-hump self-guided laser beam in Kerr media \cite{HaeltermannPRL02}.
81: A different alternative was offered in \cite{zhigang}, wherein the first stages
82: of the evolution of an optical beam, initially focused between the wells of a photorefractive
83: crystal, were monitored.
84:
85: One of the particularly interesting features of either matter-waves or optical beams in
86: double well potentials is the {\it spontaneous symmetry breaking}, i.e., the localization
87: of the respective order parameter in one of the wells of the potential.
88: Symmetry breaking solutions of the NLS model have first been predicted
89: in the context of molecular states \cite{ms} and, apart from the physical contexts
90: of BECs \cite{smerzi,kiv2,mahmud,bam,Bergeman_2mode,infeld} (see also \cite{jackson})
91: and optics \cite{HaeltermannPRL02,zhigang} mentioned above, they
92: have also been studied from a mathematical point of view in Refs.
93: \cite{mathsy,mathsy1}.
94:
95: These works underscore the relevance and timeliness of a better
96: understanding of the dynamics of nonlinear waves in double well potentials.
97: In view of that, in the present work we offer a systematic methodology,
98: based on a two-mode expansion, of how to tackle problems in double wells,
99: as regards their stationary states and the bifurcations (and ensuing instabilities)
100: that arise in them. This way, considering both cases of attractive and repulsive nonlinearities,
101: we illustrate the ways in which a symmetric double well potential is different
102: from an asymmetric one. In particular, we demonstrate that,
103: contrary to the case of symmetric potentials where symmetry
104: breaking follows a pitchfork bifurcation,
105: in asymmetric double wells the bifurcation is of the saddle-node type.
106:
107: The paper is structured as follows: in Section II,
108: we present the model and set the analytical framework.
109: In Section III, we illustrate the value of the method by
110: highlighting the significant differences of symmetric and asymmetric double wells. Finally,
111: in Section IV, we summarize our findings and discuss future directions.
112:
113: {\it Model and Analytical Approach}.
114: %
115: In a quasi-1D setting, the evolution of the mean-field wavefunction of a BEC
116: \cite{reviewsbec} (or the envelope of an optical beam \cite{kivshar})
117: is described by the following normalized NLS (GP) equation,
118: %
119: \begin{eqnarray}
120: i u_t=-\frac{1}{2} u_{xx} + s |u|^2 u + V(x) u - \mu u.
121: \label{beq1}
122: \end{eqnarray}
123: %
124: In the BEC (optics) context, $\mu$ denotes the chemical potential (propagation constant) and
125: $s=\pm 1$ for attractive or repulsive interatomic interactions (focusing or defocusing Kerr nonlinearity)
126: respectively; below, for simplicity, we will adopt the terms attractive and repulsive nonlinearity
127: for $s=\pm 1$ respectively. Finally, in Eq. (\ref{beq1}), $V(x)$ is the double well potential,
128: which is assumed to be composed by a parabolic trap
129: (of strength $\Omega$) and a ${\rm sech}^2$-shaped barrier (of strength $V_0$, width $w$ and location $x_0$);
130: in particular, $V(x)$ is of the form:
131: %
132: \begin{eqnarray}
133: V(x)=\frac{1}{2} \Omega^2 x^2 + V_0 {\rm sech}^2 \left(\frac{x-x_{0}}{w}\right),
134: \label{beq2}
135: \end{eqnarray}
136: %
137: with the choice $x_{0}=0$ ($x_{0} \neq 0$) corresponding to a symmetric (asymmetric) double well.
138: Note that such a double well can be implemented in BEC experiments
139: upon, e.g., combining a magnetic trap with a sharply focused, blue-detuned laser beam \cite{expd}.
140: Similar double wells can also be implemented e.g., in optical systems.
141:
142: The spectrum of the underlying linear Schr{\"o}dinger equation ($s=0$)
143: consists
144: of a ground state, $u_{0}(x)$, and excited states, $u_{l}(x)$ ($l \ge 1$). In the
145: nonlinear problem, using a Galerkin-type approach, we expand $u(x,t)$ as,
146: %
147: \begin{eqnarray}
148: u(x,t)=c_0(t) u_0(x) + c_1(t) u_1(x)+\cdots,
149: \label{beq3}
150: \end{eqnarray}
151: %
152: and truncate the expansion, keeping solely the first two modes;
153: here $c_{0,1}(t)$ are unknown time-dependent complex prefactors. It is worth noticing that
154: such an approximation (involving the truncation of higher order modes and
155: the spatio-temporal factorization of the wavefunction),
156: is expected to be quite useful for a weakly nonlinear analysis. In fact, as will be seen below,
157: we will be able to identify the nonlinear states that stem from the linear ones, as well as their
158: bifurcations.
159:
160: Substituting Eq. (\ref{beq3}) into Eq. (\ref{beq1}),
161: and projecting the result to the corresponding eigenmodes, we
162: obtain the following ordinary differential equations (ODEs):
163: %
164: \begin{eqnarray}
165: i \dot{c}_0 &=& (\omega_0-\mu) c_0 - s A_0 |c_0|^2 c_0 - s B (2 |c_1|^2 c_0
166: + c_1^2 \bar{c}_0),
167: \nonumber
168: \\
169: &-& s \Gamma_1 |c_1|^2 c_1 - s \Gamma_0 (2 |c_0|^2 c_1 + c_0^2 \bar{c}_1)
170: \label{beq4}
171: \\
172: i \dot{c}_1 &=& (\omega_1-\mu) c_1 - s A_1 |c_1|^2 c_1 - s B (2 |c_0|^2 c_1
173: + c_0^2 \bar{c}_1),
174: \nonumber
175: \\
176: &-& s \Gamma_0 |c_0|^2 c_0 - s \Gamma_1 (2 |c_1|^2 c_0 + c_1^2 \bar{c}_0).
177: \label{beq5}
178: \end{eqnarray}
179: %
180: In Eqs. (\ref{beq4})-(\ref{beq5}), dots denote time derivatives,
181: overbars denote complex conjugates, $\omega_{0,1}$ are the eigenvalues
182: corresponding to the eigenstates $u_{0,1}$, while $A_0=\int u_0^4 dx$, $A_1=\int u_1^4 dx$, $B=\int u_0^2 u_1^2 dx$,
183: $\Gamma_0=\int u_0 u_1^3 dx$ and $\Gamma_1=\int u_1 u_0^3 dx$ are constants.
184: Recall that $u_{0}$ and $u_1$ are real (due to the Hermitian nature of
185: the underlying linear Schr{\"o}dinger problem) and are also orthonormal.
186: Notice also that in the symmetric case ($x_{0}=0$), due to the parity of the eigenfunctions, $\Gamma_0=\Gamma_1=0$.
187:
188: We now use amplitude-phase (action-angle) variables, $c_j=\rho_j e^{i \phi_j}$, $j=0,1$
189: ($\rho_j$ and $\phi_j$ are assumed to be real), to
190: derive from the ODEs (\ref{beq4})-(\ref{beq5}) a set of four equations.
191: Introducing the function $\varphi \equiv \phi_1-\phi_0$, we find that the equations for
192: $\rho_0$ and $\phi_0$ are,
193: %
194: \begin{eqnarray}
195: \dot{\rho}_0 &=& s [(\Gamma_0 \rho_0^2
196: + \Gamma_1 \rho_1^2) \sin(\varphi) - \rho_1^2 \rho_0 \sin(2 \varphi)],
197: \label{beq6}
198: \\
199: \dot{\phi}_0 &=& (\omega_0-\mu) + s A_0 \rho_0^2 + 2 s B \rho_1^2
200: + s B \rho_1^2 \cos(2 \varphi)
201: \nonumber
202: \\
203: &+& s \left(\frac{\Gamma_1 \rho_1^3}{\rho_0}
204: + 3 \rho_0 \rho_1 \Gamma_0 \right) \cos(\varphi),
205: \label{beq7}
206: \end{eqnarray}
207: %
208: while the equations for $\rho_1, \phi_1$ are found by
209: interchanging indices $1$ and $0$ in the above equations.
210: Next, taking into regard the conservation of the total norm, we obtain the equation
211: $\rho_0^2+ \rho_1^2=N$, where $N=\int |u|^2 dx$ is the integral of motion of Eq. (\ref{beq1})
212: (the number of particles in BECs, or the power in optics). Finally, subtracting
213: Eq. (\ref{beq7}) for $\dot{\phi}_0$, and the corresponding one for $\dot{\phi}_1$, we obtain:
214: %
215: \begin{eqnarray}
216: \dot{\varphi} &=& -\Delta \omega + s (A_0 \rho_0^2 - A_1 \rho_1^2)
217: \nonumber
218: \\
219: &-& s B (2 + \cos(2 \varphi)) (\rho_0^2-\rho_1^2)-s \frac{\cos(\varphi)}{\rho_0 \rho_1}
220: \nonumber
221: \\
222: &\times& \left[\Gamma_0 \rho_0^2
223: (\rho_0^2-3 \rho_1^2) + \Gamma_1 \rho_1^2 (3 \rho_0^2 - \rho_1^2)\right].
224: \label{beq9}
225: \end{eqnarray}
226: %
227: Equations (\ref{beq6}), (\ref{beq9}) is a dynamical system, which, in principle,
228: can be thoroughly
229: investigated using phase-space analysis (such an approach has
230: been presented in \cite{smerzi,kiv2} for similar systems that were derived using
231: different expansion of the field $u$). Here, we will focus on the
232: fixed points of
233: the system [corresponding to the nonlinear
234: eigenstates of Eq. (\ref{beq1})], and analyze
235: their stability and bifurcations.
236:
237: {\it Results.}
238: %and Comparison with Numerics}.
239: Below we will analyze all possible cases ($s=\pm 1$, $x_0 = 0$, $x_0 \ne 0$)
240: %(attractive and repulsive nonlinearity, for symmetric and asymmetric potentials)
241: for the double well of Eq. (\ref{beq2}) with $V_0=1$,
242: $\Omega=0.1$, and $w=0.5$ (the results do not change qualitatively using different values).
243:
244: First we consider the case of attractive nonlinearity,
245: i.e., $s=-1$, and a symmetric double well potential with $x_{0}=0$
246: (implying that $\Gamma_0=\Gamma_1=0$). In this case,
247: the parameters involved in Eqs. (\ref{beq6}) and (\ref{beq9}) are found to be
248: $A_0 = 0.09078$, $A_1=0.09502$, $B=0.08964$, $\omega_0=0.13282$ and $\omega_1=0.15571$.
249: Then, it is readily observed that the possible real solutions of Eq. (\ref{beq6})
250: are $\rho_0=0$ and $\rho_1=0$, as well as $\varphi=0 \hspace{1.5mm} ({\rm mod} \hspace{1.5mm} \pi)$.
251: The former two are continuations of the linear solutions in the nonlinear regime.
252: However, the latter one is a non-trivial combination of
253: the two modes for $\varphi=\pi$ that results in an {\it asymmetric} pair of mirror-symmetric
254: solutions \cite{jackson,zhigang}, emerging through a {\it pitchfork} bifurcation.
255: From Eq. (\ref{beq9}), we obtain that this new branch of solutions
256: bifurcates from the symmetric branch $(\rho_0,\rho_1)=(\sqrt{N},0)$ for
257: %
258: \begin{eqnarray}
259: N>N_c=\frac{\Delta \omega}{3 B-A_0},
260: \label{beq10}
261: \end{eqnarray}
262: %
263: and for $\mu<\mu_c=\omega_0-A_0 N=0.12115$.
264: These analytical predictions are in {\it excellent} agreement with
265: the numerical results $\mu_c = 0.122 (\pm 0.001)$. It is also easy to see that the
266: anti-symmetric branch $(\rho_0,\rho_1)=(0,\sqrt{N})$ does not give
267: rise to such a bifurcation. The different branches of the full
268: numerical solutions (including the bifurcating ones) have been
269: obtained through numerical fixed point algorithms solving the steady
270: state version of Eq. (\ref{beq1}), and using continuation of the
271: solutions over the parameter $\mu$. The results are shown in the top left panel of Fig.1, where
272: the norm of the solutions $N=\int |u|^2 dx$ is shown as a function of the chemical potential
273: (or propagation constant in optics) $\mu$. In addition, as expected from the nature of the bifurcation,
274: the linear stability analysis has been used to illustrate the following: The emerging new asymmetric
275: (i.e., ``symmetry breaking'') branch of solutions is stable, while the original symmetric branch
276: is unstable beyond the bifurcation point due to a real eigenvalue $\lambda_r$ (see bottom left panel of Fig. 1).
277:
278: \begin{figure}[t]
279: %[h]
280: \includegraphics[width=4.2cm,height=5.2cm]{gt1}
281: \includegraphics[width=4.2cm,height=5.2cm]{gt2}
282: %\includegraphics[width=7cm]{fig1.eps}
283: \caption{(Color online) The top panels show the norm of the solutions of Eq.(\ref{beq1})
284: for attractive nonlinearity ($s=-1$) as a function of $\mu$ for symmetric
285: (left panel, $x_{0}=0$) and asymmetric double well (right panel, $x_{0}=0.5$).
286: The potential parameters are $\Omega=0.1$, $V_0=1$ and $w=0.5$.
287: The
288: %blue
289: solid lines denote the symmetric solution, the
290: %red
291: dashed-dotted lines denote the antisymmetric one, while the
292: %green
293: dashed lines
294: %(and cyan line in right panel)
295: denote the asymmetric solutions that are generated from the bifurcation
296: at $\mu_c \approx 0.122$ (pitchfork) and $\mu_c \approx 0.009$ (saddle-node) respectively.
297: %As a result of that bifurcations the symmetric solutions destabilize as shown in the second row
298: The bottom panels show the maximal real eigenvalue associated with the linear stability of the symmetric branches.
299: %around them. Moreover, due to the asymmetry of the double well, the two asymmetric branches
300: %(dashed green and dashed cyan lines) are no longer equivalent and the
301: %type of the bifurcation change from pitchfork to a saddle-node.
302: }
303: \label{fig1}
304: \end{figure}
305:
306: Next, in the same case ($s=-1$), we consider
307: a double well potenial with a {\it weak asymmetry} ($x_{0}=0.5$).
308: In this case, the constants involved in Eqs. (\ref{beq6})-(\ref{beq9}) are found to be
309: $A_0 = 0.14903$, $A_1=0.15618$, $B=0.02958$, $\omega_0=0.1249$ and $\omega_1=0.16535$, while
310: $\Gamma_0=0.0407$ and $\Gamma_1=-0.04077$. Note that even such a weakly
311: asymmetric case,
312: renders the right well ``shallower'', in the following sense: the density, or power $N$
313: (regarding the ground state of the linear problem), in the right well is smaller than the one in the left well.
314: Thus, in the nonlinear problem, the respective branches that bear the larger part
315: of the density in the right or in the left well (i.e., the ones having, roughly speaking,
316: the shape of a single pulse in each of the wells) are no longer equivalent.
317: This results in a {\it significant} difference between the asymmetric and the symmetric case discussed above,
318: namely there is no longer a pitchfork bifurcation, but instead, there is a {\it saddle-node} bifurcation.
319: This result is shown in the top right panel of Fig. 1, where $N$ is shown as a function of $\mu$.
320: It is readily observed that, due to the nature of the saddle-node bifurcation, two branches
321: (one of which is stable and the other one is unstable)
322: ``collide'' at some critical value of $\mu=\mu_c$ (see below) and disappear. These branches are
323: the more ``symmetric'' one, that has support in both wells (see solid line in top left panel of Fig. 2), and the one
324: pertaining to the state having the form of a single pulse in the shallower well (see dashed line in the
325: rightmost top panel of Fig. 2). The instability of the former branch is depicted in the bottom right panel of Fig. 1,
326: where the maximal real eigenvalue is shown as a function of $\mu$.
327: On the other hand, there exists also another single-pulse branch
328: supported over the deeper well (see top third panel of Fig. 2), which persists all the way to the linear limit.
329: %This branch describes the state in which the pulse is in the
330: %deeper well (see the bottom dashed line in top right panel of Fig. 1, as well as the dashed line
331: %in the third top panel of Fig. 2), and persists all the way to the linear limit.
332: Furthermore, the dash-dotted anti-symmetric branch of the top right of Fig. 1 is shown
333: in the second top panel of Fig. 2.
334:
335: The novel feature described above, namely the asymmetric breakdown of the pitchfork bifurcation into
336: a saddle-node one,
337: %and an ``isolated'' branch
338: is a {\it particular feature} of asymmetric double well potentials
339: that, to the best of our knowledge, has not been previously appreciated.
340: %in this context.
341: Notice that Eq. (\ref{beq9})
342: predicts that the saddle-node bifurcation occurs at $\mu_c=0.08748$,
343: while the numerical result is $\mu_c=0.09 \pm 0.001$; apparently the two results are again in excellent agreement.
344:
345: \begin{figure}[t]
346: %[tbp]
347: \includegraphics[width=4.22cm,height=5.2cm]{dw_fig2b}
348: \includegraphics[width=4.22cm,height=5.1cm]{dw_fig2a}
349: \caption{(Color online) The steady state solutions of Eq.(\ref{beq1}), (see also Fig. \ref{fig1})
350: for the focusing, asymmetric case (top panels) and their linear stability (bottom panels) for $\mu=0.05$.
351: The black-dashed line shows the double well potential.}
352: \label{fig2b}
353: \end{figure}
354:
355: \begin{figure}[t]
356: %[h]
357: \includegraphics[width=4.25cm,height=5.2cm]{gt3.eps}
358: \includegraphics[width=4.25cm,height=5.2cm]{gt4.eps}
359: %\includegraphics[width=7cm]{fig3.eps}
360: \caption{(Color online)
361: Same as in Fig.\ref{fig1} but for the repulsive nonlinearity ($s=+1$).
362: The
363: %red
364: dashed-dotted lines denote the symmetric solution, the
365: %blue
366: solid the antisymmetric, while the
367: %green
368: dashed lines
369: %(and cyan line in right panel)
370: denote the asymmetric solutions generated from the bifurcation at
371: $\mu_c \approx 0.168$ (pitchfork) and $\mu_c \approx 0.207$ (saddle-node)
372: respectively. Contrary to the case $s=-1$, the bifurcations originate from the anti-symmetric branch.}
373: \label{fig3}
374: \end{figure}
375:
376: Let us now consider the repulsive nonlinearity ($s=+1$).
377: In this case, for the symmetric potential ($x_{0}=0$), the pitchfork bifurcation still occurs;
378: however, now it does not originate from the symmetric branch, but rather from the anti-symmetric one with
379: $(\rho_0,\rho_1)=(0,\sqrt{N})$, giving again rise to symmetry breaking.
380: %, i.e., to an asymmetric solution.
381: Analyzing Eq. (\ref{beq9}), we find that this occurs when
382: %
383: \begin{eqnarray}
384: N> N_c \geq \frac{\Delta \omega}{3 B- A_1},
385: \label{beq11}
386: \end{eqnarray}
387: %
388: and for $\mu=\omega_0 + 3 B N=0.16822$, once again in excellent agreement
389: with the numerical result $\mu_c= 0.168 (\pm 0.001)$.
390:
391: On the other hand, in the same case ($s=+1$) but for an asymmetric ($x_{0}=0.5$)
392: double well, the bifurcation still originates from anti-phase (between the wells)
393: solutions, but again (as in the asymmetric case with $s=-1$), the bifurcation
394: is of the saddle-node type.
395: This is theoretically predicted to occur at $\mu_c=0.21342$,
396: once again in very close
397: agreement to the numerical result $\mu_c=0.207 \pm 0.001$.
398: The details of the bifurcation diagrams, are illustrated in Fig. \ref{fig3} (analogously to Fig. 1),
399: while the steady state solutions and their linear stability are shown in Fig. \ref{fig4b} (analogously to Fig. 2).
400:
401: %Finally, in Fig. \ref{fig5}, one can observe the evolution of the unstable solutions for the asymmetric case.
402:
403: \begin{figure}[t]
404: %[tbp]
405: \includegraphics[width=4.25cm,height=5.2cm]{dw_fig4a.eps}
406: \includegraphics[width=4.25cm,height=5.2cm]{dw_fig4b.eps}
407: %\includegraphics[width=7cm]{unstabledwfig4.eps}
408: \caption{(Color online) The steady state solutions of Eq.(\ref{beq1}),
409: (see also Fig. \ref{fig3}) for $s=+1$ in the asymmetric case (top panels) and their linear
410: stability (bottom panels) for $\mu=0.22$. The black-dashed line shows the double well potential.}
411: \label{fig4b}
412: \end{figure}
413:
414: {\it Conclusions.} In conclusion, we have presented a systematic
415: analysis based on a Galerkin, two-mode truncation of the
416: stationary states of symmetric and asymmetric double well potentials.
417: The analysis has been carried out both for repulsive and attractive nonlinearities and,
418: as such, can be relevant to a variety of physical contexts; these include matter-wave physics
419: (most directly), nonlinear optics, as well as other contexts where
420: it is relevant to consider double well potentials in the NLS model proper.
421: We have demonstrated that our analytical approach describes quite accurately, both
422: {\it qualitatively} and {\it quantitatively} the features of the nonlinear solutions;
423: numerical results were shown to be in excellent agreement with the analytical
424: predictions.
425:
426: In the case of a symmetric double well potential, it has been shown
427: that a symmetry-breaking (pitchfork) bifurcation of the ground state occurs
428: for attractive nonlinearities, while it is absent for repulsive nonlinearities.
429: It has also been found that a similar bifurcation of the first excited state occurs
430: in the relevant branches for repulsive nonlinearities, oppositely to the case of attractive ones,
431: where such bifurcation does not happen.
432: Additionally, regarding the above feature, we have illustrated that
433: symmetric potentials are very particular (degenerate) due to their
434: special characteristic of mirror-equivalence of the emerging symmetry-breaking states.
435: We have shown that even weak asymmetries lift
436: this degeneracy and lead to saddle-node bifurcations
437: instead of pitchfork ones that were similarly quantified
438: in both attractive and repulsive nonlinearity contexts.
439:
440: These results underscore the relevance of analyzing steady state features
441: of nonlinear models (in the presence of external potentials) based
442: on the states of the underlying linear equations. It would be
443: particularly interesting to examine the extent to which dynamical
444: features of such models can be captured by similar truncations.
445: Such studies are currently in progress.
446: %and will be reported in
447: %future publications.
448:
449: {\it Acknowledgements.} Constructive discussions with M.K. Oberthaler are kindly acknowledged.
450: This work has been partially supported from ``A.S. Onasis'' Public Benefit Foundation (G.T.) and NSF (P.G.K.).
451:
452: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
453:
454: \begin{thebibliography}{99}
455:
456: \bibitem{NLS} C. Sulem and P.L. Sulem, \newblock {\it The Nonlinear Schr{\"o}dinger Equation},
457: (Springer-Verlag, New York, 1999).
458:
459: \bibitem{trubatch} M.J. Ablowitz, B. Prinari and A.D. Trubatch,
460: {\it Discrete and Continuous Nonlinear Schr{\"o}dinger systems}
461: (Cambridge University Press, Cambridge, 2003).
462:
463: \bibitem{gbec} F. Dalfovo, S. Giorgini, L.P. Pitaevskii, and S. Stringari,
464: %{\it et al.,}
465: Rev. Mod. Phys. {\bf 71}, 463 (1999).
466:
467: \bibitem{reviewsbec} P.G. Kevrekidis and D.J. Frantzeskakis,
468: Mod. Phys. Lett. B {\bf 18}, 173 (2004);
469: V.V. Konotop and V.A. Brazhnyi, Mod. Phys. Lett. B {\bf 18} 627, (2004).
470:
471: \bibitem{kivshar} Yu.S. Kivshar and G.P. Agrawal,
472: \textit{Optical Solitons: From Fibers to Photonic Crystals},
473: Academic Press (San Diego, 2003).
474:
475: \bibitem{reviewsopt} D. N. Christodoulides, F. Lederer and Y. Silberberg,
476: Nature \textbf{424}, 817 (2003); J.W. Fleischer {\it et al.}, Opt. Expr. {\bf 13}, 1780 (2005).
477:
478: \bibitem{smerzi} S. Raghavan {\it et al.}, Phys. Rev. A {\bf 59}, 620 (1999);
479: S. Raghavan {\it et al.}, Phys. Rev. A {\bf 60}, R1787 (1999);
480: A. Smerzi and S. Raghavan, Phys. Rev. A {\bf 61}, 063601 (2000).
481:
482: \bibitem{kiv2} E.A. Ostrovskaya {\it et al.}, Phys. Rev. A \textbf{61}, 031601 (R) (2000).
483:
484: \bibitem{mahmud} K.W. Mahmud, J. N. Kutz and W. P. Reinhardt, Phys. Rev. A \textbf{66}, 063607 (2002).
485:
486: \bibitem{bam} V.S. Shchesnovich, B.A. Malomed, and R.A. Kraenkel, Physica D {\bf 188}, 213 (2004).
487:
488: \bibitem{Bergeman_2mode} D. Ananikian and T. Bergeman, Phys. Rev. A
489: \textbf{73}, 013604 (2006).
490:
491: \bibitem{infeld} P. Zi\'{n} {\it et al.}, Phys. Rev. A
492: \textbf{73}, 022105 (2006).
493:
494: \bibitem{markus1} M. Albiez {\it et al.}, Phys. Rev. Lett. {\bf 95},
495: 010402 (2005).
496:
497: \bibitem{HaeltermannPRL02} C. Cambournac \textit{et al.}, Phys. Rev. Lett.
498: \textbf{89}, 083901 (2002).
499:
500: \bibitem{zhigang} P.G. Kevrekidis {\it et al.},
501: Phys. Lett. A {\bf 340}, 275 (2005).
502:
503: \bibitem{ms} E.B. Davies, Commun. Math Phys. {\bf 64}, 191 (1979)
504:
505: \bibitem{jackson} R. K. Jackson and M. I. Weinstein, J. Stat. Phys. \textbf{116}, 881 (2004).
506:
507: \bibitem{mathsy} W. H. Aschenbacher {\it et al.},
508: %J. Fr\"{o}hlich, G. M. Graf, K. Schnee and M. Troyer,
509: J. Math. Phys. \textbf{43}, 3879 (2002).
510:
511: \bibitem{mathsy1} T. Kapitula and P.G. Kevrekidis, Nonlinearity {\bf 18}, 2491 (2005).
512:
513: \bibitem{expd}
514: C.\ Raman {\it et al.},
515: %M.\ K{\"{o}}hl, D. S.\ Durfee, C. E.\ Kuklewicz,
516: %Z.\ Hadzibabic and W.\ Ketterle,
517: Phys.\ Rev.\ Lett.\ \textbf{83}, 2502
518: (1999); R.\ Onofrio {\it et al.},
519: %C.\ Raman, J.M.\ Vogels, J. R.\ Abo-Shaeer, A.P.\
520: %Chikkatur and W.\ Ketterle,
521: Phys.\ Rev.\ Lett.\ \textbf{85}, 2228 (2000).
522:
523:
524: \end{thebibliography}
525:
526: \end{document}
527:
528: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
529:
530:
531: EXPANSIONS
532: %%%%%%%%%%
533:
534: In the case $x_{0}=0$, the spectrum of the underlying linear Schr{\"o}dinger equation ($s=0$)
535: typically consists of a symmetric ground state, $u_{0}(x)$, and
536: an anti-symmetric first excited state $u_{1}(x)$ . In the case
537: $x_{0} \neq 0$,
538:
539: (for the symmetric
540: double well potential). Our approach, similarly to the spirit of
541: \cite{smerzi,mathsy1} consists of
542: %%%%%%%%%%%%%%
543: In Eq. (\ref{beq3}), $u_{0,1}(x)$ are respectively the ground
544: and first excited state of the underlying linear problem and
545: $c_{0,1}(t)$ are their time-dependent prefactors.
546: Notice, however, that contrary to the spirit of \cite{smerzi}, we will
547: {\it not} expand to the basis of eigenfunctions localized in
548: each well.
549:
550: It should be noticed that the expansion (\ref{beq3}) in the solutions of the respective linear problem
551: is similar to the one used in \cite{zhigang,jackson,mathsy1}. On the other hand, it is different from the one used in
552: \cite{smerzi} (where the expansion was to the basis of eigenfunctions localized in each well)
553: and the ones in \cite{kiv2} (where ``coupled mode theory'')
554:
555: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
556:
557: TERMINOLOGY
558:
559: Note that for
560: $x_{0}=0$ the states $u_{0}(x)$ and $u_{1}(x)$ are symmetric and anti-symmetric respectively;
561: below, we will use the same terminology also for $x_{0} \neq 0$.
562:
563: \begin{figure}[t]
564: %[tbp]
565: \includegraphics[width=4.2cm,height=4.2cm]{unstabledwfig2.eps}
566: \includegraphics[width=4.2cm,height=4.2cm]{unstabledwfig4.eps}
567: \caption{(Color online) Spatio-temporal countour plot of the density of the unstable
568: solutions for the attractive (left panel)
569: and repulsive case (right panel) both for asymmetric potentials.}
570: \label{fig5}
571: \end{figure}
572:
573: %\bibitem{markusref1} B.D. Josephson, Phys. Lett. {\bf 1}, 251 (1962).
574: