1: \documentclass[twocolumn,showpacs,preprintnumbers,aps]{revtex4}
2: \usepackage{amssymb}
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: \usepackage{amsmath}
5: \usepackage{graphicx}
6:
7: %TCIDATA{OutputFilter=Latex.dll}
8: %TCIDATA{LastRevised=Monday, August 15, 2005 14:07:27}
9: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
10: %TCIDATA{Language=American English}
11:
12: \input{tcilatex}
13:
14: \begin{document}
15:
16: \title{Instability of coherent states \ of a real scalar field}
17: \author{Vladimir A. Koutvitsky}
18: \affiliation{Pushkov Institute of Terrestrial Magnetism, Ionosphere and Radiowave
19: Propagation of the Russian Academy of Sciences (IZMIRAN), Troitsk, Moscow
20: Region, 142190, Russia}
21: \author{Eugene M. Maslov}
22: \affiliation{Pushkov Institute of Terrestrial Magnetism, Ionosphere and Radiowave
23: Propagation of the Russian Academy of Sciences (IZMIRAN), Troitsk, Moscow
24: Region, 142190, Russia}
25:
26: \begin{abstract}
27: We investigate stability of both localized time-periodic coherent states
28: (pulsons) and uniformly distributed coherent states (oscillating condensate)
29: of a real scalar field satisfying the Klein-Gordon equation with a
30: logarithmic nonlinearity. The linear analysis of time-dependent parts of
31: perturbations leads to the Hill equation with a singular coefficient. To
32: evaluate the characteristic exponent we extend the Lindemann-Stieltjes
33: method, usually applied to the Mathieu and Lam\'{e} equations, to the case
34: that the periodic coefficient in the general Hill equation is an unbounded
35: function of time. As a result, we derive the formula for the characteristic
36: exponent and calculate the stability-instability chart. Then we analyze the
37: spatial structure of the perturbations. Using these results we show that the
38: pulsons of any amplitudes, remaining well-localized objects, lose their
39: coherence with time. This means that, strictly speaking, all pulsons of the
40: model considered are unstable. Nevertheless, for the nodeless pulsons the
41: rate of the coherence breaking in narrow ranges of amplitudes is found to be
42: very small, so that such pulsons can be long-lived. Further, we use the
43: obtaned stability-instability chart to examine the Affleck-Dine type
44: condensate. We conclude the oscillating condensate can decay into an
45: ensemble of the nodeless pulsons.
46: \end{abstract}
47:
48: \pacs{03.65.Pm, 05.45.Yv, 11.10.Lm, 11.27.+d}
49: \maketitle
50:
51: \section{INTRODUCTION}
52:
53: Nonlinear localized field configurations, solitons, are currently considered
54: as models of various physical objects, from elementary particles and collective
55: excitations in condensed matter to giant
56: lumps of dark matter in the form of soliton stars and galactic halos \cite%
57: {Bishop, Lee, Mielke}. Stability properties of solitons were investigated by many authors,
58: and a number of important results has been obtained (see, e.g., \cite%
59: {Rybakov} and references therein). In particular, Hobart \cite{Hobart} and
60: Derrick \cite{Derrick} have proved that static multidimensional scalar
61: solitons are energetically unstable, and, hence, these objects cannot last
62: in a real world for a long time. One way to avoid this theorem is to invoke
63: time dependence. Along this line main efforts were focused on the stability
64: analysis of the stationary states, i.e., coherent states of a complex scalar
65: field oscillating harmonically in time. It turned out, however, that for a
66: wide class of relativistic models these states can be only conditionally
67: stable, i.e., stable with respect to a certain type of perturbations (e.g.,
68: conserving the scalar charge) \cite{Rybakov}. As to the time-periodic states
69: of a more general form, both complex and real, there are presently no strong
70: analytical results on their stability.
71:
72: In this paper we examine stability of time-periodic configurations of the
73: form%
74: \begin{equation}
75: \phi =\phi _{0}(t,\mathbf{r})=a(t)u(\mathbf{r}) \label{eq2}
76: \end{equation}%
77: satisfying the nonlinear Klein-Gordon equation
78: \begin{equation}
79: \phi _{tt}-\Delta \phi +U^{\prime }(\phi )=0. \label{eq1}
80: \end{equation}%
81: These are coherent states in the sense the field oscillates synchronously at
82: all spatial points. It is necessary to stress that we consider real
83: solutions, so that the energy density oscillates as well (in contrast to the
84: stationary states for which $a(t)$\ $\varpropto e^{i\omega t}$). Solitons
85: with oscillating energy density are usually called pulsons.
86:
87: It turns out that for real $\phi $ the ansatz (\ref{eq2}) determines
88: uniquely the potential $U(\phi )$ in Eq. (\ref{eq1}). Namely, if neither $%
89: a(t)$ nor $u(\mathbf{r})$ are constants, the only potential admitting such
90: solutions will have the form \cite{Maslov1}%
91: \begin{equation}
92: U(\phi )=\frac{1}{2}\phi ^{2}[m^{2}+\lambda (1-\ln \phi ^{2})], \label{eq3}
93: \end{equation}%
94: where $m^{2}$ and $\lambda $ are arbitrary constants.
95:
96: Originally, the Klein-Gordon equation with a logarithmic potential of this
97: type has been introduced in the quantum field theory by G. Rosen \cite{Rosen}%
98: . Later on Bialynicki-Birula and Mycielski \cite{B-BM1} have rediscovered
99: this equation and also considered its nonrelativistic version, the nonlinear
100: Schr\"{o}dinger equation \cite{B-BM2}.
101:
102: In inflationary cosmology and in modern supersymmetric field theories the
103: logarithmic nonlinearities appear naturally when quantum corrections to
104: effective potentials are allowed for \cite{Linde1, Linde2, Barrow, Enqvist1}%
105: . In this context the expression in the square brackets of Eq. (\ref{eq3})
106: can be treated as a dynamic inflaton mass term $m_{S}^{2}$ that is the bare
107: inflaton mass term $m^{2}$ plus the logarithmic correction. It can be
108: represented in the commonly considered form \cite{Enqvist1, Enqvist4,
109: Multamaki1, Multamaki2, Kasuya, Enqvist2, Enqvist3, Pawl} by the substitution $\ln
110: \phi ^{2}=1+\ln (\Phi /M)^{2}$, $\lambda /m^{2}=-K$, where $\Phi $ is an
111: inflaton scalar field, $M$ is a large mass scale, $K$ is a constant (usually
112: negative and small). Thus our consideration is also relevant to dynamics of
113: the pulson exitations of a real inflaton field oscillating around a vacuum
114: value.
115:
116: Note that the multidimensional pulsons probably exist in other scalar models
117: as well. Thus the long-living oscillating spherically symmetric localized
118: states were numerically found in the sine-Gordon, $\phi ^{4}$, and $\phi
119: ^{3}-\phi ^{4}$ models \cite{Bog-Mak1,Bog,Bog-Mak2,Gleiser1,Copeland} (see %
120: \cite{Belova}\ for a review). Unfortunately, the analytic form of these
121: solutions is so far unknown.
122:
123: The model (\ref{eq1}) and (\ref{eq3}) is unique in the sense it has a whole
124: family of exact pulson solutions of the form (\ref{eq2}), all existing in
125: any number of spatial dimensions \cite{Maslov1}. This is also true for
126: complex version of the model \cite{Marques,Bogolubsky}. The real pulsons we
127: are dealing with are the limiting states of the complex ones, when the
128: scalar charge tends to zero. Other limiting states are Q-balls for which $%
129: a(t)\varpropto e^{i\omega t}$ \cite{Enqvist4, Multamaki1}. It is believed
130: that Q-balls can arise due to fragmentation of the Affleck-Dine condensate %
131: \cite{Kasuya, Enqvist2, Enqvist3, Pawl}. We will see below that the parametric
132: instability of the oscillating condensate leads to the resonant
133: fragmentation that can give rise to the pulson formation at the nonlinear
134: stage. Like Q-balls \cite{Multamaki2}, pulsons interact elastically or
135: inelastically in collisions depending on their relative velocities, phases,
136: and rest masses \cite{Maslov2,Maslov3}. Thus, in model (\ref{eq1}) and (\ref%
137: {eq3}) the light pulsons with given relative velocities interact always
138: elastically, independently of their phases. In contrast, the collisions of
139: heavy pulsons can result in formation of the so-called explosons, localized
140: states with exponentially growing amplitude \cite{Maslov2}. For the
141: intermediate masses the picture depends essentially on the phases of the
142: colliding pulsons and impact velocity determining the duration of the
143: interaction \cite{Maslov3}.
144:
145: The above results suggest that there is a domain of parameters where pulsons
146: are stable, at least in short time interactions. But in what sense? How long
147: a pulson conserves its characteristic features once interaction ends? If
148: pulsons are long-lived objects they will be interesting candidates for the
149: dark matter constituents having time-dependent density. What is known about
150: stability of an isolated pulson at the long time scale? Surprisingly, but
151: very few. In Ref. \cite{Bogolubsky} it was argued in favour of its perfect
152: stability. No deviations from the exact solution (\ref{eq2}) were found
153: after about one thousand oscillations. However, our preliminary numerical
154: experiments \cite{Koutvitsky, Koutvitsky-Maslov} have shown that the pulsons
155: of certain amplitudes, even perturbed by computer round-off errors only,
156: gradually lose their coherency, remaining well-localized oscillating
157: objects. This has motivated the closer examination.
158:
159: In the present paper we clarify how long the pulsons can conserve the
160: coherency depending on their parameters. For this purpose we investigate
161: stability of the spherically symmetric pulson solutions (\ref{eq2}) with
162: respect to small initial perturbations of an arbitrary form.
163:
164: The paper is organized as follows. In Sec. II the main properties of the
165: real pulsons of the model considered are reviewed. Section III is wholly
166: devoted to the linear stability analysis. We arrive at the singular Hill
167: equation and generalize the Lindemann-Stieltjes method to evaluate the
168: characteristic exponent. On this basis we examine stability of the pulsons
169: and discuss fragmentation of the oscillating Affleck-Dine type condensate.
170: In Sec. IV we make some remarks concerning the complex pulsons and summarize
171: the main results.
172:
173: \section{PULSONS AS COHERENT STATES}
174:
175: Assuming $\lambda $ positive, let us first eliminate the constants $m^{2}$
176: and $\lambda $ from consideration by the scaling $t\rightarrow \lambda
177: ^{-1/2}t$, $\mathbf{r}\rightarrow \lambda ^{-1/2}\mathbf{r}$, $\phi
178: \rightarrow \phi \exp \frac{m^{2}}{2\lambda }$. In the new variables the
179: field $\phi $ may be thought of as satisfying Eq. (\ref{eq1}) with the
180: potential
181:
182: \begin{equation}
183: U(\phi )=\frac{1}{2}\phi ^{2}(1-\ln \phi ^{2}). \label{eq4}
184: \end{equation}%
185: It is the potential we will deal with. It has local minimum at $\phi =0$ and
186: two maxima at $\phi =\pm 1$, at the minimum the potential having the
187: singularity: its second derivative tends to infinity as $\phi \rightarrow 0$.
188:
189: The substitution of the ansatz (\ref{eq2}) into Eq. (\ref{eq1}) leads then
190: to two independent equations,
191: \begin{equation}
192: a_{tt}=-\frac{d}{da}\left[ \frac{1}{2}a^{2}(1-\ln a^{2})\right] ,
193: \label{eq5}
194: \end{equation}%
195: \begin{equation}
196: \Delta u=-\frac{d}{du}\left[ \frac{1}{2}u^{2}(\ln u^{2}-1)\right] .
197: \label{eq6}
198: \end{equation}%
199: Note that the potentials in the square brackets of Eqs. (\ref{eq5}) and (\ref%
200: {eq6}) have the same form as the potential (\ref{eq4}) taken with plus and
201: minus signs, respectively. The existence of the oscillating localized
202: solutions (\ref{eq2}) is thus apparent from consideration of motion of a
203: mechanical particle in these potentials.
204:
205: Let us consider in more detail the oscillatory solutions of Eq. (\ref{eq5}).
206: Using the Hamiltonian and denoting $\xi =a/a_{\max }$ ($0<a_{\max }<1$, $%
207: -1\leqslant \xi \leqslant 1$), we obtain%
208: \begin{equation}
209: \xi _{t}^{2}=\omega _{0}^{2}(1-\xi ^{2})+\xi ^{2}\ln \xi ^{2}, \label{eq7}
210: \end{equation}%
211: where
212: \begin{equation}
213: \omega _{0}^{2}=1-\ln a_{\max }^{2}>1. \label{eq8}
214: \end{equation}%
215: In the case of small amplitudes, $a_{\max }^{2}\ll 1$, $\omega _{0}^{2}\gg 1$%
216: , Eq. (\ref{eq7}) gives
217: \begin{equation}
218: \xi (t)\approx \cos \omega _{0}t. \label{eq9}
219: \end{equation}%
220: Thus, we have quasi-harmonic high-frequency oscillations which are however
221: nonlinear since their period,
222: \begin{equation}
223: T\approx \frac{2\pi }{\left| \ln a_{\max }^{2}\right| ^{1/2}}, \label{eq10}
224: \end{equation}%
225: depends on the amplitude \cite{Koutvitsky-Maslov}. In the next approximation
226: from Eq. (\ref{eq7}) we find%
227: \begin{equation}
228: T=\frac{2\pi }{\omega _{0}}\left( 1+\frac{0.307}{\omega _{0}^{2}}+O\left(
229: \frac{1}{\omega _{0}^{4}}\right) \right) . \label{eq11}
230: \end{equation}%
231: In the case of near-critical amplitudes, when $a_{\max }^{2}\rightarrow 1$,
232: the oscillations become almost rectangular and have the period%
233: \begin{equation}
234: T\approx 2\sqrt{2}\ln \frac{1}{1-a_{\max }^{2}}. \label{eq12}
235: \end{equation}%
236: Examples of solutions of Eq. (\ref{eq5}) are shown in Fig. 1.
237: %%%%%%%%%%%%% Fig.1.%%%%%%%%%%%%%%%%%%%
238: \begin{figure}[tbp]
239: \includegraphics[width=8cm]{fig1.eps}
240: \caption{Oscillatory solutions of Eq. (\ref{eq5}) for $a_{\max }=0.1$, $%
241: a_{\max }=0.7$, and $a_{\max }=0.9999$.}
242: \end{figure}
243: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
244:
245: The spatial structure of a pulson is determined by Eq. (\ref{eq6}). In the
246: spherically symmetric case this equation has a discrete spectrum of
247: localized $N$-nodal solutions $u_{N}(r)$ with the first derivatives
248: vanishing at the origin \cite{B-BM3} (see Fig. 2). The simplest of them, the
249: nodeless solution, has a Gaussian-like shape,
250: \begin{equation}
251: u_{0}(r)=e^{(3-r^{2})/2}, \label{eq13}
252: \end{equation}%
253: and is usually called gausson \cite{Rosen,B-BM1,B-BM2}. It is agreed that
254: its effective radius equals $\sqrt{2}$. In the multinodal solutions, as $r$
255: increases, the field undergoes spatial oscillations of the half-wavelength $%
256: L\lesssim 2\sqrt{2}$ and then decays as%
257: \begin{equation}
258: u_{N}(r)\approx C_{N}e^{-(r-\rho _{N})^{2}/2}\quad (r\gg r-\rho _{N},\;N\gg
259: 1), \label{eq13a}
260: \end{equation}%
261: where $C_{N}$ is the value of the last extremum of $u_{N}(r)$ attained at $%
262: r=\rho _{N}$, $C_{N}\rightarrow (-1)^{N}e^{1/2}\;(N\rightarrow \infty )$, $%
263: \rho _{N}\sim NL$. Thus, the pulsons of the model (\ref{eq1}) and (\ref{eq4}%
264: ) are well-localized states of the inhomogeneity length $L$, at all points
265: the field oscillating coherently with the period $T$. (To return to the
266: physical units these scales should be multiplied by $\lambda ^{-1/2}$.) In
267: our dimensionless variables the pulsons are characterized by two parameters
268: only: the amplitude $a_{\max }$ and the number of the nodes $N$ (or $T$ and $%
269: u_{N}(0)$, respectively).
270: %%%%%%%%%%%%% Fig.2.%%%%%%%%%%%%%%%%%%%
271: \begin{figure}[th]
272: \includegraphics[width=8cm]{fig2.eps}
273: \caption{Spectrum of the spherically symmetric $N$-nodal solutions $u_{N}(r)$
274: of Eq. (\ref{eq6}): $u_{0}(0)=e^{3/2}$, $u_{1}(0)=9.726$, $u_{2}(0)=15.084$,
275: $u_{3}(0)=20.526$, \dots\ .}
276: \end{figure}
277: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
278:
279: It should be stressed that, due to nonanalyticity of $U(\phi )$ at $\phi =0$%
280: , the right-hand sides of Eqs. (\ref{eq5}) and (\ref{eq6}) are nonanalytic
281: when $a$ and $u$ become zero. Hence, the solutions $a(t)$ and $u(r)$
282: themselves become nonanalytic at those points $t$ and $r$ where they pass
283: through zero. Thus in the solution (\ref{eq9}) we have dropped the terms
284: which are small (of the order of $\omega _{0}^{-2}$) but nonanalytic when $%
285: \xi (t)=0$. In general case from Eq. (\ref{eq7}) it follows that $\xi (t)$
286: passes through zero at $t=t_{m}$ as%
287: \begin{eqnarray}
288: \underset{t\rightarrow t_{m}}{\xi (t)} &=&\pm \omega _{0}(t-t_{m})\Bigl[1+%
289: \frac{1}{6}(t-t_{m})^{2}\ln (t-t_{m})^{2} \notag \\
290: &&+O\left( (t-t_{m})^{2}\right) \Bigr], \label{eq14}
291: \end{eqnarray}
292: where $\pm $ sign is taken for $\xi _{t}(t_{m})\gtrless 0$. It is seen that $%
293: \xi _{ttt}$ becomes infinite as $t\rightarrow t_{m}$. Similarly, one can
294: show that in the vicinity of the $n$-th node ($n=1,\dots,N$)%
295: \begin{eqnarray}
296: \underset{r\rightarrow r_{n}}{u(r)} &=&u_{r}(r_{n})(r-r_{n})\Bigl[1-\frac{%
297: r-r_{n}}{r_{n}} \notag \\
298: &&-\frac{1}{6}{(r-r_{n})^{2}\ln (r-r_{n})^{2}} \notag \\
299: &&+O\left( (r-r_{n})^{2}\right)\Bigr]. \label{eq15}
300: \end{eqnarray}
301: As we will see below, the nonanalyticity of $U(\phi )$ gives rise to some
302: specific features of the stability analysis.
303:
304: \section{THE LINEAR STABILITY ANALYSIS}
305:
306: Consider a small fluctuation $\eta (t,\mathbf{r})$ around the spherically
307: symmetric pulson (\ref{eq2}), $\phi =\phi _{0}(t,r)+\eta (t,\mathbf{r})$. In
308: the linear approximation the equation for $\eta $ reads%
309: \begin{equation}
310: \eta _{tt}-\Delta \eta -(2+\ln \phi _{0}^{2})\eta =0. \label{eq17}
311: \end{equation}%
312: Seeking a solution in the form $\eta (t,\mathbf{r})\varpropto X(t)\Psi (%
313: \mathbf{r})$ we arrive at the equations%
314: \begin{eqnarray}
315: X_{tt}+(E-2-\ln a^{2})X &=&0, \label{eq19} \\
316: \Delta \Psi +(E+\ln u^{2})\Psi &=&0, \label{eq20}
317: \end{eqnarray}%
318: where $E$ is some constant.
319:
320: The expression in the brackets of Eq. (\ref{eq17}) is $-U^{\prime \prime
321: }(\phi _{0})$. It becomes infinite, as well as the expressions in the
322: brackets of Eqs. (\ref{eq19}) and (\ref{eq20}), at the points $t_{m}$ and $%
323: r_{n}$ where $a(t)$ and $u(r)$ become zero. Thus we need to analyze the
324: second order differential equations with singular coefficients. We begin
325: with Eq. (\ref{eq19}) which has the periodic singular coefficient $\ln
326: a^{2}(t)$ and hence belongs to the class of Hill equations.
327:
328: \subsection{Singular Hill equation and generalized Lindemann-Stieltjes method%
329: }
330:
331: It turns out to be very useful to look at the problem as a whole,
332: considering first the Hill equation of a general form
333: \begin{equation}
334: X_{tt}+h(z(t))X=0. \label{eq21}
335: \end{equation}%
336: We will assume that $h(z)$ is an integral function of $z$, while $z(t)$ is a
337: real-valued periodic (of a period $\tau $) even function of $t$, having, in
338: general, singularities, but such, that $h(z(t))$ remains still integrable.
339:
340: It is well known that the Hill equation describes the physical systems in
341: which the parametric resonance can occur. In the context of our stability
342: analysis we will be interested in real resonant solutions of Eq. (\ref{eq21}%
343: ). In accordance with the Floquet theory (see, e.g., \cite{Magnus}), any one
344: of these solutions can be represented as a linear combination of the
345: fundamental solutions
346: \begin{equation}
347: X_{+}(t)=\varphi (t)e^{\mu t},\quad X_{-}(t)=\varphi (-t)e^{-\mu t},
348: \label{eq22}
349: \end{equation}%
350: where $\varphi (t)$ is a $\tau $-periodic or $\tau $-antiperiodic real
351: function, $\mu >0$ is the characteristic exponent. In the case that $z(t)$
352: is unbounded it is impossible to obtain the solutions and evaluate $\mu $ by
353: expansions in Fourier series, following the standard Hill approach. Another
354: way is to apply the Lindemann-Stieltjes method \cite{Whittaker}. In some
355: cases it allows one to obtain the results in a closed analytical form \cite%
356: {Greene,Kaiser,Maslov4,Finkel}. We first outline this method in the context
357: of the general Hill equation (\ref{eq21}) with an extension to the case that
358: the periodic function $z(t)$ is unbounded. In doing so we follow the paper %
359: \cite{Maslov4} where the method was used to construct the resonant solutions
360: of the Lam\'{e} equation.
361:
362: The main idea is as follows. Let us treate $z$ as a new \textquotedblleft
363: time\textquotedblright\ variable instead of $t$. In each interval of
364: monotonicity of $z(t)$ we define
365: \begin{equation}
366: y(z)=X(t). \label{eq24}
367: \end{equation}%
368: Assume that the periodic function $z(t)$ satisfies the equation%
369: \begin{equation}
370: z_{t}^{2}=g(z), \label{eq23}
371: \end{equation}%
372: where $g(z)$ is an integral function of $z$. Eq. (\ref{eq21}) then becomes
373: \begin{equation}
374: g(z)y^{\prime \prime }+\frac{1}{2}g^{\prime }(z)y^{\prime }+h(z)y=0
375: \label{eq25}
376: \end{equation}
377: (hereinafter the prime denotes $d/dz$).
378:
379: Let us first suppose $z(t)$ is bounded. Equation (\ref{eq23}) then shows
380: that it is differentiable. Zeros of the function $g(z)$ on the complex $z$
381: plane, taken to be isolated, are singular points of Eq. (\ref{eq25}). Since $%
382: z(t)$ is periodic and real-valued, among singular points there are two, $%
383: \zeta _{1}$ and $\zeta _{2}$, lying on the real axis and being minimal and
384: maximal values which $z(t)$ acquires at the end points of the intervals of
385: monotonicity. Also, it follows that $g^{\prime }(\zeta _{1,2})\neq 0$.
386: Physically, this is well understood, since $\zeta _{1}$ and $\zeta _{2}$ can
387: be treated as turning points in periodic motion of a mechanical particle,
388: e.g., of a nonlinear oscillator, under the action of the force $g^{\prime
389: }/2 $. From Eq. (\ref{eq23}) it is clear that the interval $\left[ \zeta
390: _{1},\zeta _{2}\right] $ does not contain other singular points of Eq. (\ref%
391: {eq25}).
392:
393: For example, in the case of the Mathieu equation we have $z(t)=\cos ^{2}t$, $%
394: g(z)=4z(1-z)$, so that Eq. (\ref{eq25}) has the regular singular points $%
395: z=\zeta _{1}=0$, $z=\zeta _{2}=1$, both being the turning points. In
396: addition, the equation has an irregular singularity at infinity. For the Lam%
397: \'{e} equation $z(t)=\func{sn}^{2}(t,\varkappa )$, $g(z)=4z(1-z)(1-\varkappa
398: ^{2}z)$. Equation (\ref{eq25}) then has the regular singular points $z=\zeta
399: _{1}=0$, $z=\zeta _{2}=1$, $z=\varkappa ^{-2}>1$, first two of them being
400: the turning points, and a regular singularity at infinity.
401:
402: In general, it is easy to verify that the turning points $\zeta _{1,2}$ are
403: regular singular points of Eq. (\ref{eq25}), the exponents at each being $0$
404: and $1/2$. This implies that in the vicinity of each turning point $\zeta $
405: there exist two independent solutions of Eq. (\ref{eq25}), $y^{(0)}(z;\zeta
406: ) $ and $y^{(1/2)}(z;\zeta )$, having asymptotics $1+O(z-\zeta )$ and $%
407: (z-\zeta )^{1/2}[1+O(z-\zeta )]$, correspondingly.
408:
409: Now let us consider any one interval of monotonicity of the $\tau $-periodic
410: even function $z(t)$. Denote as $y_{1}(z)$ and $y_{2}(z)$ those two linearly
411: independent solutions of Eq. (\ref{eq25}) one of which coinsides, by Eq. (%
412: \ref{eq24}), with the increasing solution (\ref{eq22}), $X_{+}(t)$, and
413: another with the decreasing one, $X_{-}(t)$, on the interval chosen. Since $%
414: \varphi (t)$ is either $\tau $-periodic or $\tau $-antiperiodic, the product
415: $\varphi (t)\varphi (-t)=X_{+}X_{-}=y_{1}y_{2}=w(z)$ is always $\tau $%
416: -periodic even function defined on the whole $t$ axis. Hence, at the end
417: points of the intervals of monotonicity of $z(t)$, i.e., at $t_{m}=m\tau
418: /2\;(m=0,\pm 1,\dots)$, the derivative $[(X_{+}X_{-})_{t}]_{t=t_{m}}=0$ or,
419: what is the same,
420: \begin{equation}
421: \left( w^{\prime }\sqrt{g}\right) _{z=\zeta _{1,2}}=0. \label{eq26}
422: \end{equation}%
423: In the vicinity of a turning point $\zeta $ the solutions $y_{1}$ and $y_{2}$
424: can be represented as linear combinations of the solutions $y^{(0)}$ and $%
425: y^{(1/2)}$. Consequently, the singularity $(z-\zeta )^{1/2}$ is the only one
426: which the function $w(z)=y_{1}y_{2}$ might have. But its existence is in
427: contradiction with Eq. (\ref{eq26}), because $g^{\prime }(\zeta )\neq 0$
428: and, hence, $g(z)_{z\rightarrow \zeta }\sim g^{\prime }(\zeta )(z-\zeta )$.
429: Therefore, the product $y_{1}y_{2}$ is analytic at $z=\zeta _{1,2}$. Recall
430: now that the interval $\left[ \zeta _{1},\zeta _{2}\right] $ does not
431: contain other singular points of Eq. (\ref{eq25}) and the singular points
432: are assumed to be isolated. We thus conclude that on the complex $z$ plane
433: there exists a vicinity of the interval $\left[ \zeta _{1},\zeta _{2}\right]
434: $, i.e., an open domain $D\supset \left[ \zeta _{1},\zeta _{2}\right] $, in
435: which $y_{1}y_{2}$ is an analytic function of $z$. In addition, it follows
436: that $y_{1}^{2}$ and $y_{2}^{2}$ of necessity have singularities of the type
437: $(z-\zeta )^{1/2}$ and, thus, cannot satisfy Eq. (\ref{eq26}).
438:
439: Now consider the case that one of the turning points or the both are at
440: infinity. This implies that at the corresponding instants $t_{m}$ the
441: functions $z(t)$ and $z_{t}(t)$ become unbounded, the latter changing the
442: sign. Nevertheless, we assume that in the vicinities of $t_{m}$ the function
443: $h(z(t))$ in Eq. (\ref{eq21}) is integrable and $X$ is continuous, whence it
444: follows that $X_{t}$ and, therefore, $w_{t}$ are also continuous. Hence, as
445: before, $(w_{t})_{t=t_{m}}=0$ due to evenness and periodicity, so that we
446: arrive at Eq. (\ref{eq26}) again, where $\zeta _{1}=-\infty $ and/or $\zeta
447: _{2}=+\infty $.
448:
449: It is easy to verify that the bilinear combinations $y_{1}^{2}$, $y_{1}y_{2}$%
450: , and $y_{2}^{2}$ constitute the fundamental system of solutions of the
451: third-order differential equation
452: \begin{equation}
453: g(z)w^{\prime \prime \prime }+\frac{3}{2}g^{\prime }(z)w^{\prime \prime
454: }+\left( \frac{1}{2}g^{\prime \prime }(z)+4h(z)\right) w^{\prime
455: }+2h^{\prime }(z)w=0. \label{eq27}
456: \end{equation}%
457: Equation (\ref{eq26}) is thus a common criterion for selection of the
458: solution%
459: \begin{equation}
460: w=y_{1}y_{2} \label{eq28}
461: \end{equation}%
462: from the set of solutions of Eq. (\ref{eq27}). In the case that $z(t)$ is
463: bounded, Eq. (\ref{eq26}) is the equivalent to the requirement that a
464: solution of Eq. (\ref{eq27}) be analytic in $D$. If $z(t)$ is unbounded, Eq.
465: (\ref{eq26}) will give the boundary conditions at infinity which must be
466: satisfied in solving Eq. (\ref{eq27}). In this case $w(z)$ will be analytic
467: in a vicinity $D$ of one of the intervals $(-\infty ,\zeta _{2}]$, $[\zeta
468: _{1},\infty )$, $(-\infty ,\infty )$.
469:
470: Thus, in a neighbourhood of any one point $\zeta \in D$ we can write the
471: expansions%
472: \begin{equation}
473: \begin{pmatrix}
474: w(z) \\
475: g(z) \\
476: h(z)
477: \end{pmatrix}%
478: =\sum\limits_{n=0}^{\infty }%
479: \begin{pmatrix}
480: w_{n} \\
481: g_{n} \\
482: h_{n}
483: \end{pmatrix}%
484: (z-\zeta )^{n}, \label{eq28a}
485: \end{equation}%
486: Substitution of (\ref{eq28a}) into Eq. (\ref{eq27}) leads to the following
487: set of equations for the coefficients:%
488: \begin{equation}
489: m\sum_{n=1}^{m+2}n(m+n)g_{m-n+2}w_{n}+4\sum_{n=0}^{m}(m+n)h_{m-n}w_{n}=0
490: \label{eq34}
491: \end{equation}%
492: $(m=1,2,\dots)$. Thus for $m=1$ we have%
493: \begin{equation}
494: 6g_{0}w_{3}+3g_{1}w_{2}+(g_{2}+4h_{0})w_{1}+2h_{1}w_{0}=0. \label{eq33}
495: \end{equation}%
496: Assuming $w(\zeta )\neq 0$, we normalize $w(z)$ by $w_{0}=1$. Then, at given
497: $w_{1}$ and $w_{2}$ the remaining coefficients $w_{n}$ are determined from
498: Eqs. (\ref{eq34}). The choice of $w_{1}$ and $w_{2}$ is not arbitrary but
499: determined by Eq. (\ref{eq26}). Thus, setting $\zeta =\zeta _{1}$ and,
500: hence, $g_{0}=0$, we must choose $w_{1}$ in such a way that the series (\ref%
501: {eq28a}) for $w$ (or its continuation) converges at the second turning point
502: $\zeta _{2}$, or satisfies the boundary condition at infinity (\ref{eq26})
503: if $\zeta _{2}=+\infty $. For the Mathieu and Lam\'{e} equations this leads
504: to the function $w(z)$ which is an integral one, for the latter it being a
505: polinomial \cite{Whittaker}. In these cases the domain $D$ is evidently the $%
506: z$ plane with $\left| z\right| <\infty $.
507:
508: Let us suppose the function $w(z)$ (\ref{eq28}) is found. Return now to Eqs.
509: (\ref{eq21})-(\ref{eq25}). Denote as $W$ the Wronskian of the solutions (\ref%
510: {eq22}),%
511: \begin{equation}
512: X_{+}X_{-t}-X_{+t}X_{-}=W=\limfunc{const}. \label{eq35}
513: \end{equation}%
514: Setting
515: \begin{eqnarray}
516: y_{1} &=&X_{+},\quad y_{2}=X_{-}\qquad (z_{t}\geqslant 0), \notag \\
517: y_{1} &=&X_{-},\quad y_{2}=X_{+}\qquad (z_{t}\leqslant 0), \label{Eq36}
518: \end{eqnarray}%
519: we then obtain%
520: \begin{equation}
521: y_{1}y_{2}^{\prime }-y_{1}^{\prime }y_{2}=W/\sqrt{g}, \label{eq37}
522: \end{equation}%
523: where $\sqrt{g}\geqslant 0$ is assumed. The system of equations (\ref{eq28})
524: and (\ref{eq37}) can be easily solved, which gives%
525: \begin{equation}
526: y_{1,2}^{2}=\exp \int \frac{f_{\mp }}{w\sqrt{g}}\, dz, \label{eq38}
527: \end{equation}%
528: where%
529: \begin{equation}
530: f_{\pm }=w^{\prime }\sqrt{g}\pm W. \label{eq39}
531: \end{equation}
532:
533: Now let us insert $y_{1,2}$ (\ref{eq38}) back into Eq. (\ref{eq25}). We
534: obtain%
535: \begin{equation}
536: 2gww^{\prime \prime }+g^{\prime }ww^{\prime }-gw^{\prime 2}+4hw^{2}+W^{2}=0.
537: \label{eq40}
538: \end{equation}%
539: By this formula one can find the constant $W^{2}$ from a knowledge of $w(z)$
540: in a vicinity of any point $z$. Thus, calculating (\ref{eq40}) at any one
541: finite turning point $\zeta $ we obtain (in terms of expansions (\ref{eq28a}%
542: ) with normalization $w_{0}=1$)%
543: \begin{equation}
544: W^{2}=-4h_{0}-g_{1}w_{1}. \label{eq41}
545: \end{equation}%
546: Alternatively, one can take zeros $z_{i}$ of $w(z)$. (The functions $g(z)$
547: and $w(z)$ do not have common zeros because otherwise the Wronskian (\ref%
548: {eq35}) would be zero.) Then we find
549: \begin{equation}
550: W^{2}=g(z_{i})w^{\prime 2}(z_{i}). \label{eq42}
551: \end{equation}%
552: The requirement for positivity of $W^{2}$ determines the values of
553: parameters of Eq. (\ref{eq21}) (resonance zones) for which the resonant
554: solutions exist.
555:
556: Let us construct these solutions. Consider the intervals of monotonicity $%
557: t_{1}\leqslant t\leqslant t_{2}$ $(z_{t}\geqslant 0)$, $t_{2}\leqslant
558: t\leqslant t_{3}$ $(z_{t}\leqslant 0)$, etc. According to (\ref{Eq36}) and (%
559: \ref{eq38}) we can write%
560: \begin{eqnarray}
561: \underset{t_{1}\leqslant t\leqslant t_{2}}{X_{\pm }^{2}(t)} &=&X_{\pm
562: }^{2}(t_{1})\exp \int_{\zeta _{1}}^{z}\frac{f_{\mp }}{w\sqrt{g}}\, dz,
563: \notag \\
564: \underset{t_{2}\leqslant t\leqslant t_{3}}{X_{\pm }^{2}(t)} &=&X_{\pm
565: }^{2}(t_{2})\exp \int_{\zeta _{2}}^{z}\frac{f_{\pm }}{w\sqrt{g}}\, dz,%
566: \mathrm{\ etc.} \label{eq43}
567: \end{eqnarray}%
568: To find the characteristic exponent $\mu $ consider, e.g., the growing
569: solution $X_{+}(t)$. Setting $t=t_{2}$, $z(t_{2})=\zeta _{2}$ in the first
570: equation of (\ref{eq43}) and $t=t_{3}$, $z(t_{3})=\zeta _{1}$ in the second
571: one, we can express $X_{+}^{2}(t_{3})$ through $X_{+}^{2}(t_{1})$. Using Eq.
572: (\ref{eq22}) and taking into account that $t_{3}=t_{1}+\tau $, $\varphi
573: (t+\tau )=\pm \varphi (t)$, we thus obtain%
574: \begin{equation}
575: \mu =-\frac{W}{\tau }\int_{\zeta _{1}}^{\zeta _{2}}\frac{dz}{w\sqrt{g}}.
576: \label{eq44}
577: \end{equation}%
578: Recall that $\tau $ is the period of $z(t)$, the constant $W$ is determined
579: from Eq. (\ref{eq41}) or Eq. (\ref{eq42}), its sign being taken opposite to
580: that of the integral in (\ref{eq44}) to provide for positivity of $\mu $.
581: Since $w(z)$ has zeros, the integrals in Eqs. (\ref{eq43}) and (\ref{eq44})
582: are understood as their principal values. Formula (\ref{eq44}) is a simple
583: generalization of the ones used previously in Refs. \cite%
584: {Greene,Kaiser,Maslov4,Finkel}.
585:
586: \subsection{Evaluation of the characteristic exponent\textit{\ }}
587:
588: Let us return to Eq. (\ref{eq19}). It can be written in the form of Eq. (\ref%
589: {eq21}) if we set%
590: \begin{eqnarray}
591: z(t) &=&-\ln (a/a_{\max })^{2},\quad z(0)=0, \label{eq45} \\
592: h(z) &=&E-3+\omega _{0}^{2}+z. \label{eq48}
593: \end{eqnarray}%
594: Equation (\ref{eq7}) then immediately gives%
595: \begin{equation}
596: g(z)=4[\omega _{0}^{2}(e^{z}-1)-z]. \label{eq47}
597: \end{equation}%
598: Zeros of this function are shown in Fig. 3. Since $\xi (t)=a/a_{\max }$
599: oscillates with the period $T$ in the interval $-1\leqslant \xi \leqslant 1$
600: [see Eqs. (\ref{eq7})-(\ref{eq12})], the function $z(t)$ (\ref{eq45})
601: oscillates with the period $\tau =T/2$ between the turning points $\zeta
602: _{1}=0$ and $\zeta _{2}=+\infty $.
603: %%%%%%%%%%%%% Fig.3.%%%%%%%%%%%%%%%%%%%
604: \begin{figure}[th]
605: \centering\includegraphics[width=6cm]{fig3.eps}
606: \caption{The layout of zeros of $g(z)$ (\ref{eq47}) on the complex $z$
607: plane. The vicinity $D$ (shaded) of the interval $[0,\infty )$ belongs to
608: the domain of analyticity of $w(z)$.}
609: \end{figure}
610: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
611: To calculate the characteristic exponent by the formula (\ref{eq44}) we need
612: to know the function $w(z)$ which is the solution of Eq. (\ref{eq27}) with
613: boundary conditions (\ref{eq26}). Unfortunately, for given $h(z)\ $(\ref%
614: {eq48}) and $g(z)$ (\ref{eq47}) equation (\ref{eq27}) cannot be solved
615: analytically. We solve it numerically for various values of the parameters $E
616: $ and $\omega _{0}^{2}=1-\ln a_{\max }^{2}$ \cite{Koutvitsky-Maslov}. Doing
617: so, we use the conditions (\ref{eq26}) in the following way. As discussed
618: above, the fulfilment of (\ref{eq26}) at a finite turning point means
619: analyticity of $w(z)$ in some vicinity of this point. Therefore, we can use
620: the expansions (\ref{eq28a}) setting there $\zeta =\zeta _{1}=0$, $g_{0}=0$,
621: $g_{1}=4(\omega _{0}^{2}-1)$, $g_{n}=4\omega _{0}^{2}/n!$ $(n=2,3,\dots )$, $%
622: h_{0}=E+\omega _{0}^{2}-3$, $h_{1}=1$. Equation (\ref{eq33}) then gives%
623: \begin{equation}
624: w_{2}=-\frac{1+(2E+3\omega _{0}^{2}-6)w_{1}}{6(\omega _{0}^{2}-1)}.
625: \label{eq49}
626: \end{equation}%
627: We thus solve Eq. (\ref{eq27}) with the following conditions at $z=0$: $%
628: w(0)=1$, $w^{\prime }(0)=w_{1}$, $w^{\prime \prime }(0)=2w_{2}$. Given
629: values of $E$ and $\omega _{0}^{2}$, we choose $w_{1}$ so as to satisfy the
630: condition (\ref{eq26}) at infinity,%
631: \begin{equation}
632: \left( w^{\prime }\sqrt{g}\right) _{z\rightarrow +\infty }\rightarrow 0.
633: \label{eq50}
634: \end{equation}%
635: At the same time, since $\mu $ assumed to be real, the values of $E$, $%
636: \omega _{0}^{2}$, and $w_{1}$ must provide for positivity of $W^{2}$,%
637: \begin{equation}
638: W^{2}=4[3-E-\omega _{0}^{2}-(\omega _{0}^{2}-1)w_{1}]>0 \label{eq51}
639: \end{equation}%
640: [see Eqs. (\ref{eq41}) and (\ref{eq44})]. Conditions (\ref{eq50}) and (\ref%
641: {eq51}) determine the resonance zones in the space of parameters $E$ and $%
642: \omega _{0}^{2}$ (or $a_{\max }^{2}$). Hereinafter the zones will be
643: referred to as $Z_{j}$ and numbered sequentially as $E$ grows (with $a_{\max
644: }^{2}$ fixed) starting with $j=-1$ in the region $E<0$. Figure 4 shows the
645: solutions $w(z)$ for zones $Z_{1}$, $Z_{2}$, and $Z_{3}$ lying in the region
646: $E>2$.
647: %%%%%%%%%%%%% Fig.4.%%%%%%%%%%%%%%%%%%%
648: \begin{figure}[th]
649: \includegraphics[width=8cm]{fig4.eps}
650: \caption{Behavior of $w(z)$ for $E=4$ in the different resonance zones: (a)
651: zone $Z_{1}$, $a_{\max }^{2}=0.5625$, (b) zone $Z_{2}$, $a_{\max }^{2}=0.9025
652: $, (c) zone $Z_{3}$, $a_{\max }^{2}=0.9598$. The values of $a_{\max }^{2}$
653: choosen correspond to the centers of zones where $\protect\mu $ achieves its
654: maxima. One can see that the number of zeros of $w(z)$ is unit above the
655: number of a zone.}
656: \end{figure}
657: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
658: Now, knowing $w(z)$, we can calculate the integral in (\ref{eq44}). Because $%
659: w(z)$ has zeros, we first transform the integrand with the help of Eq. (\ref%
660: {eq40}) extracting the total derivative. Owing to the condition (\ref{eq50}%
661: ), the latter does not contribute to the principal value of the integral,
662: while the remaining terms give
663: \begin{equation}
664: \int_{0}^{\infty }\frac{dz}{w\sqrt{g}}=-\frac{1}{2W^{2}}\int_{0}^{\infty }%
665: \left[ \sqrt{g}\left( w^{\prime }\sqrt{g}\right) ^{\prime }\ln w^{2}+8hw%
666: \right] \frac{dz}{\sqrt{g}}. \label{eq52}
667: \end{equation}%
668: %%%%%%%%%%%%% Fig.5.%%%%%%%%%%%%%%%%%%%
669: \begin{figure}[th]
670: \includegraphics[width=8cm]{fig5.eps}
671: \caption{Resonant solutions of Eq. (\ref{eq19}), $X(t)\sim X_{+}(t)=\protect%
672: \varphi (t)e^{\protect\mu t}$, (solid lines) and the function $z(t)=-\ln
673: (a/a_{\max })^{2}$ (dashed lines): (a) zone $Z_{1}$, (b) zone $Z_{2}$, (c)
674: zone $Z_{3}$. The initial conditions are: $a(0)=a_{\max }$, $a_{t}(0)=0$, $%
675: X_{t}(0)=0$, $X(t)$ is normalized in a proper way. The values of $E$ and $%
676: a_{\max }^{2}$ in each zone are the same as in Fig 4. It is seen that $%
677: \protect\varphi (t)$ is $\protect\tau $-periodic in $Z_{1}$, $\protect\tau $%
678: -antiperiodic in $Z_{2}$, $\protect\tau $-periodic in $Z_{3}$, and so on, in
679: accordance with the solutions (\ref{eq43}) [see Eqs. (\ref{eq39}) and (\ref%
680: {eq42}) and Fig. 4].}
681: \end{figure}
682: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
683: The integrand on the right-hand side of Eq. (\ref{eq52}) is more convinient
684: for numerical integration because its singularities are all integrable. We
685: perform the integration in (\ref{eq52}), calculate $W^{2}$ by the formula (%
686: \ref{eq51}), and find the period $T=2\tau $ by integration of Eq. (\ref{eq7}%
687: ). These procedures are carried out numerically for a set of grid points in
688: every resonance zone. In this way from (\ref{eq44}) we obtain the
689: characteristic exponent $\mu $ as a function of $E$ and $a_{\max }^{2}$.
690:
691: To check this result we derive $\mu (E,a_{\max }^{2})$ directly from
692: analysis of numerical solutions of Eq. (\ref{eq19}). Examples of these
693: solutions for resonance zones $Z_{1}$, $Z_{2}$, and $Z_{3}$ are shown in
694: Fig. 5. The growth of the amplitude with time is clearly seen. The function $%
695: \mu (E,a_{\max }^{2})$ so derived is found to be fully coincident with the
696: one obtained by the formula (\ref{eq44}).
697: %%%%%%%%%%%%% Fig.6.%%%%%%%%%%%%%%%%%%%
698: \begin{figure}[ht]
699: \includegraphics[width=8.5cm]{fig6.eps}
700: \caption{The stability-instability chart: (a) $E>2$, first ten zones are
701: shown, (b) $E\leqslant 2$.}
702: \end{figure}
703: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
704:
705: The resulting stability-instability chart is presented in Fig. 6. Figure
706: 6(a) shows the region $E>2$. There is an infinite series of narrow resonance
707: zones $Z_{1}$, $Z_{2}$, $Z_{3}$, \dots , the first one having the highest
708: magnitude ($\approx 0.08$ at the maximum). All these zones originate from
709: the point $E=2$, $a_{\max }^{2}=1$ at which $\mu =0$ [see Eqs. (\ref{eq44})
710: and (\ref{eq51})]. In the region $E\leqslant 2$ we have two zones, $Z_{0}$
711: and $Z_{-1}$, lying in the ranges $0<E<2$ and $E<0$, correspondingly. Since
712: in these zones the values of $\mu $ proved to be much greater than in $Z_{1}$%
713: , $Z_{2}$, \dots , we depict the surface $\mu (E,a_{\max }^{2})$ for this
714: region separately, in Fig. 6(b).
715:
716: \subsection{Spatial structure of the perturbation}
717:
718: Consider now Eq. (\ref{eq20}). It has the form of the Schr\"{o}dinger
719: equation for a quantum particle of the energy $E$ moving in the\ potential $%
720: -\ln u^{2}$. Since the potential tends to $+\infty $ with growing $r$ [as $%
721: r^{2}$, see Eqs. (\ref{eq13}) and (\ref{eq13a})], the energy spectrum is
722: discrete, $E=E_{n}$, and the corresponding eigenfunctions $\Psi _{n}(\mathbf{%
723: r})$ are all localized. In the case of the nodeless pulson (\ref{eq13}) we
724: have the isotropic harmonic oscillator. Its eigenfunctions are well known
725: (see, e.g., \cite{Flugge}). We write them as follows:%
726: \begin{eqnarray}
727: \Psi _{n}(\mathbf{r}) &=&\sum_{l=0}^{n}[1+(-1)^{n-l}]R_{nl}(r)Y_{l}(\theta
728: ,\varphi ), \label{eq53} \\
729: R_{nl}(r) &=&r^{l}e^{-r^{2}/2}\Phi \left( -\frac{n-l}{2},\ l+\frac{3}{2},\
730: r^{2}\right) , \label{eq54} \\
731: Y_{l}(\theta ,\varphi ) &=&\sum_{m=-l}^{l}c_{l,m}P_{l}^{\left| m\right|
732: }(\cos \theta )e^{im\varphi }. \label{eq55}
733: \end{eqnarray}%
734: Here $\Phi (\alpha ,\gamma ,x)$ is the Kummer function, $P_{\nu }^{\mu }(x)$
735: are the associated Legendre functions, $c_{l,m}$ are constants, $%
736: c_{l,-m}=c_{l,m}^{\ast }$. The energy spectrum is given by%
737: \begin{equation}
738: E=E_{n}=2n\qquad (n=0,1,2,\dots). \label{eq56}
739: \end{equation}%
740: (Our energy levels are shifted with respect to the conventional ones since
741: the minimum of the potential $-\ln u_{0}^{2}$ is $-3$.)
742:
743: In the case of the nodal pulsons the picture becomes more complicated due to
744: the loss of the orbital degeneracy. The corresponding eigenfunctions and
745: eigenvalues can be calculated only numerically. As an example, in Fig. 7 is
746: shown the energy spectrum for perturbations of the one-nodal pulson.
747: %%%%%%%%%%%%% Fig.7.%%%%%%%%%%%%%%%%%%%
748: \begin{figure}[th]
749: \centering\includegraphics[width=8 cm]{fig7.eps}
750: \caption{The energy levels for the perturbations with different orbital
751: numbers $l$. The case of the one-nodal pulson.}
752: \end{figure}
753: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
754:
755: Note that there always exist the eigenvalues $E=0$, $l=0$ and the
756: corresponding eigenfunction $\Psi _{0}(r)\varpropto u(r)$. This fact
757: immediately follows from the comparison of Eqs. (\ref{eq6}) and (\ref{eq20}%
758: ). The corresponding $X_{0}(t)$ in $\eta (t,\mathbf{r})$ is an oscillating
759: function with the amplitude growing linearly with time. It is easy to see,
760: however, that this mode is physically meaningless. Indeed, it will formally
761: appear if we perturb the pulson by a small variation of its amplitude $%
762: a_{\max }$ but not the form $u(r)$. Due to nonlinearity, this results in a
763: pulson with slightly shifted frequency. Then the difference of the perturbed
764: and unperturbed pulsons, i.e., $\eta (t,r)$, will have the form of beats
765: generated by two oscillations with close frequencies and the same profile $%
766: u(r)$. The function $X_{0}(t)$ approximates the initial, linearly growing
767: part of a beat. We exclude this mode from the subsequent consideration,
768: since it belongs to the class of perturbations that conserve a pulson as a
769: whole. Next, for the nodal pulsons only, there is a mode with $E=0$, $l=1$
770: (see Fig. 7). Since this mode cannot grow faster than linearly in time, we
771: also do not take it into account. Further, we should exclude the mode
772: resulting from a small translation of the pulson. The corresponding
773: eigenfunction is proportional to $\mathbf{n}\nabla u$, where $\mathbf{n}$ is
774: a displacement vector. Using Eqs. (\ref{eq6}) and (\ref{eq20}) one can
775: easily show that this mode corresponds to $E=2$, $l=1$. Thus the resulting
776: perturbation is written as%
777: \begin{equation}
778: \eta (t,\mathbf{r})=\sum\limits_{n}X_{n}(t)\Psi _{n}(\mathbf{r}),
779: \label{eq31}
780: \end{equation}%
781: where $X_{n}$ is a solution of Eq. (\ref{eq19}) with $E=E_{n}$, $E_{n}\neq
782: 0,2$. If $E_{n}$ and $a_{\max }^{2}$ are in a resonance zone, $X_{n}(t)$
783: will be represented as a linear combination of the solutions (\ref{eq22})
784: and, hence, will grow with time as $e^{\mu (E_{n},a_{\max }^{2})t}$.
785:
786: \subsection{Instability of the pulsons}
787:
788: The arrangement of the resonance zones on the $(E,a_{\max }^{2})$ plane
789: indicates that for any spectrum $E_{n}$ there always exist the ranges of $%
790: a_{\max }^{2}$ where pulsons are unstable. But do the values of $a_{\max
791: }^{2}$ exist for which the pulsons are stable? To answer this question let
792: us return to the surface $\mu (E,a_{\max }^{2})$ depicted in Fig. 6. Take,
793: at first, the spectrum for the nodeless pulson. We choose the sections $\mu
794: _{n}(a_{\max }^{2})$ of the surface $\mu (E,a_{\max }^{2})$ by $E=2n$ and
795: project them on the $(\mu ,a_{\max }^{2})$ plane. As a result, the pattern
796: shown in Fig. 8(a) emerges.
797: %%%%%%%%%%%%% Fig.8.%%%%%%%%%%%%%%%%%%%
798: \begin{figure}[th]
799: \centering\includegraphics[width=8cm]{fig8.eps}
800: \caption{Superposition of the sections $\protect\mu _{n}(a_{\max }^{2})$ of
801: the surface $\protect\mu (E,a_{\max }^{2})$: (a) the nodeless pulson, $%
802: E_{n}=4$, $6$, $8$, \dots , $50$ $(n=2,3,4,\dots ,25)$, (b) the one-nodal
803: pulson, $E_{n}=-0.7142$, $0.4833$, $1.1222$, $1.8996$, \dots .}
804: \end{figure}
805: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
806: It is clearly seen the tendency to the total filling of the interval $%
807: 0<a_{\max }^{2}<1$ by the resonant peaks as the successively higher energy
808: levels are accounted for. This implies that for any given $a_{\max }^{2}$
809: there always exists an unstable mode with $\mu =\mu _{n}(a_{\max }^{2})$,
810: i.e., strictly speaking, all nodeless pulsons are unstable. On the other
811: hand, the figure shows that there are domains of $a_{\max }^{2}$ where the
812: peaks are very small. These domains are the gaps between the main peaks
813: originated from the low-energy cross sections of the surface $\mu (E,a_{\max
814: }^{2})$ over a few first zones. In the gaps the exponent $\mu $ is small, so
815: that the corresponding pulsons are long-lived. For example, in Ref. \cite%
816: {Koutvitsky-Maslov} we observed numerically the nodeless pulson with $%
817: a_{\max }^{2}=0.49$ that conserved its coherency against the radially
818: symmetric perturbations over the course of several hundreds of periods.
819:
820: Further, the above projective procedure is performed using the spectrum of
821: the one-nodal pulson (Fig. 7). The main contribution here is made by the
822: sections with the energies $E_{n}=-0.7142$ $(l=0)$, $0.4833$ $(l=2)$, $1.1222
823: $ $(l=3)$, and $1.8996$ $(l=4)$ falling into zones $Z_{-1}$ and $Z_{0}$, the
824: projections of the first and the third sections overlapping the other ones.
825: The result is presented in Fig. 8(b). We see that $a_{\max }^{2}$ axis is
826: totally full. Thus, the one-nodal pulson has neither stability nor even
827: quasistability domains. It seems likely that things will get worse, not
828: better, if one goes to the multinodal pulsons. We thus conclude that,
829: strictly speaking, all pulsons of the model considered are unstable. But
830: nodeless pulsons can be quasistable in narrow ranges of amplitudes. It is
831: the long-lived pulsons that can be of astrophysical and cosmological
832: interest. If the dark matter consists of scalar particles, such pulsons will
833: be realistic candidates for the dark matter objects having oscillating
834: density \cite{Seidel}.
835:
836: \subsection{On the instability of the Affleck-Dine type condensate}
837:
838: The obtained stability-instability chart turns out to be appropriate for the
839: stability analysis of the nonlocalized coherent states as well. As an
840: example, we consider a uniformly distributed background $\phi _{0}(t)$, a
841: scalar condensate, oscillating around the minimum of the potential (\ref{eq4}%
842: ) at $\phi =0$. This state can be formally obtained from Eq. (\ref{eq2}) if
843: we set there $u(\mathbf{r})\equiv 1$. We thus assume that $\phi _{0}(t)$
844: obeys Eq. (\ref{eq5}). Taking the perturbed state $\phi =\phi _{0}(t)+\eta
845: (t,\mathbf{r})$, in the linear approximation from Eqs. (\ref{eq1}) and (\ref%
846: {eq4}) we readily obtain%
847: \begin{equation}
848: A_{tt}+(k^{2}-2-\ln \phi _{0}^{2})A=0, \label{eq63}
849: \end{equation}%
850: where $A(t,\mathbf{k})$ is the Fourier amplitude of the perturbation, and $%
851: k=\left| \mathbf{k}\right| $. It is seen that the real and imaginary parts
852: of this equation have the form of Eq. (\ref{eq21}) with $h(z)$ given by Eq. (%
853: \ref{eq48}), $z=-\ln (\phi _{0}/\phi _{0\max })^{2}$, $\omega _{0}^{2}=1-\ln
854: \phi _{0\max }^{2}$, and $E=k^{2}$. Returning to the stability-instability
855: chart (Fig. 6) we note that in the region $E\geqslant 0$ maximal values of $%
856: \mu $ are attained in the zone $Z_{0}$ for which $0<E<2$. Interestingly,
857: this band exactly coincides with the one obtained in Ref. \cite{Enqvist3}
858: for the power-law potential approximating (\ref{eq3}) when $\lambda \ll 1$.
859: In the interior of $Z_{0}$ the exponent $\mu $ depends almost not at all on
860: the amplitude of the condensate oscillations and is a sufficiently smooth
861: function of $k^{2}$ with a maximum at $k^{2}=k_{0}^{2}\approx 1$ where $\mu
862: \approx 0.5$. Therefore, if the initial power spectrum $\left| A(0,\mathbf{k}%
863: )\right| ^{2}$ lies in the region $0\lesssim \left| \mathbf{k}\right|
864: \lesssim \sqrt{2}$ and, in addition, its characteristic width along $\mathbf{%
865: k}$ is small, $\Delta k\ll \sqrt{2}$, then the growth of the perturbation
866: amplitude will not be accompanied by significant changes in the structure of
867: the perturbation. The limiting case of such perturbations is a harmonic
868: wave. Otherwise, if $\Delta k\gtrsim \sqrt{2}$, the shape of the power
869: spectrum will vary with time so that a maximum will appear at $k_{0}\approx 1
870: $. As a result, the effective width of the spectrum will become smaller, $%
871: \Delta k\lesssim 1$. In this case, if the initial spectrum is sufficiently
872: isotropic in $\mathbf{k}$ space, the parametric amplification of the
873: perturbations will result in the emergence of the localized field
874: configurations of the characteristic size $\Delta r\sim 1/\Delta k\gtrsim 1$
875: that agrees with the radius of the gausson (see Sec. II). At this scale the
876: field practically does not undergo spatial oscillations since the
877: corresponding wavelength $2\pi /k_{0}\gtrsim 1$. We thus expect that at the
878: nonlinear stage these configurations will turn into the nodeless pulsons.
879: Their period will be equal to the period of the condensate oscillations
880: since in the zone $Z_{0}$ the parametric amplification proceeds at the basic
881: frequency. Gradually, the energy of the oscillating condensate will go to
882: ensemble of the arising pulsons, this process resulting in the damping of
883: the background oscillations. As to the pulsons themselves, they can be
884: long-lived or short-lived depending on their amplitudes, in accordance with
885: the results of the previous Subsection.
886:
887: Note, that numerical simulations performed for the complex version of the
888: model (\ref{eq1}) and (\ref{eq3}) have shown the fragmentation of both the
889: rotating \cite{Kasuya, Enqvist2} and oscillating \cite{Enqvist3}
890: Affleck-Dine condensate. The localized configurations arising in the
891: condensate have been identified with Q-balls. We belive, however, that the
892: configurations observed in the oscillating condensate are in fact the
893: complex pulsons (see Sec. IV), rather than the usual Q-balls. This
894: possibility was early discussed in Ref. \cite{Enqvist4} where an attempt to
895: simulate the complex pulson has been made.
896:
897: The resonant exitation of the pulsons was also observed in the two-vacuum $%
898: \phi ^{4}-\phi ^{6}$ model within a regularly oscillating background \cite%
899: {Maslov4} and in the $\phi ^{4}$ model within an initially thermalized
900: background \cite{Gleiser2}. Note that in two-vacuum models the pulsons can
901: play the role of nuclei of a new phase. In Ref. \cite{Maslov4} the general
902: suggestion has been made that the parametric resonance can underlie the
903: mechanism responsible for the first-order phase transitions in nonlinear
904: non-dissipative systems. This conjecture turns out to be in agreement with
905: recent results of Ref. \cite{Gleiser3} where the resonant nucleation within
906: the thermalized background have been numerically observed in the $\phi
907: ^{3}-\phi ^{4}$ model. Note, in addition, that the dynamical nucleation can
908: also take place in the nonlinear Schr\"{o}dinger equation \cite{Barashenkov}.
909:
910: \section{CONCLUDING REMARKS}
911: In this paper we have examined only the linear stage of instability at which
912: small deformations of the pulson's shape result in loss of the coherence.
913: There is numerical evidence that in time the growth of the perturbations
914: becomes saturated due to nonlinear effects \cite{Koutvitsky-Maslov}. We thus
915: suggest that in the model considered the pulsons, while unstable, remain
916: well localized objects with no tendency for spreading or collapsing.
917:
918: Further, we dealt with a real scalar field. It would be interesting to
919: perform the similar analysis for a complex scalar field too. It is believed
920: that the existence of the scalar charge can stabilize a field lump. For
921: Q-balls this fact is well established (so-called Q-theorem \cite{Lee,
922: Rybakov, Belova}). In contrast, for the complex pulsons this is an open
923: question. As it was shown in Refs. \cite{Marques,Bogolubsky}, the field
924: equation (\ref{eq1}) with $U^{\prime }=-\phi \ln (\phi \phi ^{\ast })$
925: admits the exact pulson solutions of the form $\phi
926: _{0}(t,r)=a(t)u(r)e^{i\theta (t)}$, where $a(t)$, $u(r)$, and $\theta (t)$
927: are real. The function $u(r)$ satisfies Eq. (\ref{eq6}) as before, while $%
928: a(t)$ oscillates with a period $T$ in accordance with the equation
929: \begin{equation}
930: a_{tt}=-\frac{d}{da}\left[ \frac{1}{2}a^{2}(1-\ln a^{2})+\frac{q^{2}}{2a^{2}}%
931: \right] , \label{eq64}
932: \end{equation}%
933: where $q$ is a real constant, $q^{2}<(2e)^{-1}$, and $\theta _{t}=qa^{-2}$.
934: The constant $q$ is proportional to the charge of the scalar field. In
935: contrast to Eq. (\ref{eq5}), the potential in the square brackets of Eq. (%
936: \ref{eq64}) prevents $a(t)$ from being zero. Without loss of generality one
937: may assume $a(t)$ positive, so that the oscillations occur arround the
938: minimum of the potential at $a=a_{0}$, where $a_{0}$ is the least positive
939: root of the equation $a^{4}\ln a^{2}=-q^{2}$. If $a$ is at rest in this
940: minimum, then $\theta (t)=qa_{0}^{-2}t+\theta (0)$, and we have the standard
941: Q-ball. Physically, Eq. (\ref{eq64}) describes the motion of a mechanical
942: particle with an angular momentum $q$ in the potential $(a^{2}/2)(1-\ln
943: a^{2})$ \cite{Landau}. The condition for its trajectory to be closed is $%
944: \theta (T)-\theta (0)=2\pi m/n$, where $m$ and $n$ are arbitrary integers.
945: In fact, it relates the energy of the particle and its angular momentum
946: whereby such trajectories exist. In our case this means periodicity of the
947: solution $\phi _{0}(t,r)$ with the period $nT$. Obviously, there is an
948: infinity of such solutions. Taking $\phi _{0}(t,r)$ and considering the
949: partial perturbation $\eta \varpropto X(t)\Psi (\mathbf{r})$ one can find
950: that the function $\Psi (\mathbf{r})$, assumed to be real, satisfies Eq. (%
951: \ref{eq20}) as before, while $X(t)$ obeys the equation
952: \begin{equation}
953: X_{tt}+(E-1-\ln a^{2})X=e^{2i\theta }X^{\ast }, \label{eq65}
954: \end{equation}%
955: where $E$ is a real constant. This equation can be represented as a system
956: of four real first-order equations with periodic coefficients of the periods
957: $T$ and $nT$. It is significant that, since $a(t)\neq 0$, these coefficients
958: are bounded in time, so that one can attempt to estimate the characteristic
959: exponent of the system using the standard methods \cite{Yakubovich}.
960:
961: Also, it would be interesting to examine stability of a selfgravitating
962: pulson. Hopefully, gravitation can expand the domains of (quasi)stability,
963: as it is the case for Q-balls \cite{Rybakov}.These are possible subjects of
964: our future work.
965:
966: In the present paper we have investigated stability of both the coherent
967: localized states (pulsons) and nonlocalized states (uniformly oscillating
968: scalar condensate) of the real scalar field. Our main analytical result is
969: the generalization of the Lindemann-Stieltjes method to the case that the
970: periodic coefficient in the Hill equation is unbounded in time. Our main
971: numerical result is the stability-instability chart with the values of
972: characteristic exponent calculated in the resonance zones. Using this chart
973: we have found the gaps in the set of the pulson amplitude values in which
974: the real nodeless pulsons conserve the coherency for an extremely long time.
975: Also, considering the oscillating scalar condensate, we have detemined the
976: wavelength of the most unstable mode. This wavelength turned out to be equal
977: to the characteristic size of the nodeless pulson. We thus suggest the
978: pulsons can be formed due to resonant fragmentation of the scalar
979: condensate. These are our main physical results.
980:
981: \acknowledgments {\ The authors thank I. Bogolubsky, Yu.P. Rybakov, and A.
982: Shagalov for useful discussions. This work was partly supported by the RAS\
983: Presidium Program ``Nonstationary phenomena in astronomy''. }
984:
985: \begin{thebibliography}{99}
986: \bibitem{Bishop} A.R. Bishop, J.R. Krumhansl, and S.E. Trullinger, Physica D \textbf{1}, 1
987: (1980).
988:
989: \bibitem{Lee} T.D. Lee and Y. Pang, Phys. Rep. \textbf{221}, 251 (1992).
990:
991: \bibitem{Mielke} E.W. Mielke and F.E. Schunck, Nucl. Phys. B \textbf{564},
992: 185 (2000).
993:
994: \bibitem{Rybakov} Yu.P. Rybakov and V.I. Sanyuk, \textit{Multidimensional
995: Solitons }(Peoples' Friendship University Press, Moscow, 2001), in Russian.
996:
997: \bibitem{Hobart} R.H. Hobart, Proc. Phys. Soc. \textbf{82}, 201 (1963).
998:
999: \bibitem{Derrick} G.H. Derrick, J. Math. Phys. \textbf{5}, 1252 (1964).
1000:
1001: \bibitem{Maslov1} E.M. Maslov, Phys. Lett. A \textbf{151}, 47 (1990).
1002:
1003: \bibitem{Rosen} G. Rosen, Phys. Rev. \textbf{183}, 1186 (1969).
1004:
1005: \bibitem{B-BM1} I. Bialynicki-Birula and J. Mycielski, Bull. Acad. Pol.
1006: Sci., Ser. Sci. Math. Astronom. Phys. \textbf{23}, 461 (1975).
1007:
1008: \bibitem{B-BM2} I. Bialynicki-Birula and J. Mycielski, Ann. Phys. (N.Y.)
1009: \textbf{100}, 65 (1976).
1010:
1011: \bibitem{Linde1} A.D. Linde, \textit{Particle Physics and Inflationary
1012: Cosmology }(Harwood Academic, Chur, Switzerland, 1990).
1013:
1014: \bibitem{Linde2} A. Linde, Phys. Lett. B \textbf{284}, 215 (1992).
1015:
1016: \bibitem{Barrow} J.D. Barrow and P. Parsons, Phys. Rev. D \textbf{52}, 5576
1017: (1995).
1018:
1019: \bibitem{Enqvist1} K. Enqvist and J. McDonald, Phys. Lett. B \textbf{425},
1020: 309 (1998).
1021:
1022: \bibitem{Enqvist4} K. Enqvist and J. McDonald, Nucl. Phys. \textbf{B570},
1023: 407 (2000).
1024:
1025: \bibitem{Multamaki1} T. Multam\"{a}ki and I. Vilja, Nucl. Phys. \textbf{B574}%
1026: , 130 (2000).
1027:
1028: \bibitem{Multamaki2} T. Multam\"{a}ki and I. Vilja, Phys. Lett. B \textbf{482%
1029: }, 161 (2000).
1030:
1031: \bibitem{Kasuya} S. Kasuya and M. Kawasaki, Phys. Rev. D \textbf{62}, 023512
1032: (2000).
1033:
1034: \bibitem{Enqvist2} K. Enqvist, A. Jokinen, T. Multam\"{a}ki, and I. Vilja,
1035: Phys. Rev. D \textbf{63}, 083501 (2001).
1036:
1037: \bibitem{Enqvist3} K. Enqvist, S. Kasuya, and A. Mazumdar, Phys. Rev. D
1038: \textbf{66}, 043505 (2002).
1039:
1040: \bibitem{Pawl} A. Pawl, Nucl. Phys. B \textbf{679}, 231 (2004).
1041:
1042: \bibitem{Bog-Mak1} I.L. Bogolubsky and V.G. Makhankov, Pis'ma Zh. Eksp.
1043: Teor. Fiz. \textbf{24}, 15 (1976) [JETP Lett. \textbf{24}, 12 (1976)].
1044:
1045: \bibitem{Bog} I.L. Bogolubsky, Pis'ma Zh. Eksp. Teor. Fiz. \textbf{24}, 579
1046: (1976) [JETP Lett. \textbf{24}, 544 (1976)].
1047:
1048: \bibitem{Bog-Mak2} I.L. Bogolubsky and V.G. Makhankov, Pis'ma Zh. Eksp.
1049: Teor. Fiz. \textbf{25}, 120 (1977) [JETP Lett. \textbf{25} 107 (1977)].
1050:
1051: \bibitem{Gleiser1} M. Gleiser, Phys. Rev. D \textbf{49}, 2978 (1994).
1052:
1053: \bibitem{Copeland} E.J. Copeland, M. Gleiser, and H.-R. M\"{u}ller, Phys.
1054: Rev. D \textbf{52}, 1920 (1995).
1055:
1056: \bibitem{Belova} T.I. Belova and A.E. Kudryavtsev, Usp. Fiz. Nauk \textbf{167%
1057: }, 377 (1997) [Physics-Uspekhi, \textbf{40}, 359 (1997)].
1058:
1059: \bibitem{Marques} G.C. Marques and I. Ventura, Preprint of IFUSP P-83
1060: (1976); Rev. Bras. Fis. \textbf{7}, 297 (1977).
1061:
1062: \bibitem{Bogolubsky} I.L. Bogolubsky, Zh. Eksp. Teor. Fiz. \textbf{76}, 422
1063: (1979) [JETP \textbf{49}, 213 (1979)].
1064:
1065: \bibitem{Maslov2} E.M. Maslov and A.G. Shagalov, in \textit{Nonlinear
1066: Evolution Equations and} \textit{Dynamical Systems,} Proc. NEEDS'92, edited
1067: by V. Makhankov, I. Puzynin, and O. Pashaev (World Scientific, Singapore,
1068: 1993), p. 159.
1069:
1070: \bibitem{Maslov3} E.M. Maslov and A.G. Shagalov, Phys. Lett. A \textbf{224},
1071: 277 (1997).
1072:
1073: \bibitem{Koutvitsky} V. Koutvitsky, Contribution to the International
1074: Conference on Theoretical Physics, Paris, 2002, \textit{Book of Abstracts
1075: TH-2002}, edited by D. Iagolnitzer, P. Ribeca, and J. Zinn-Justin, p. 198.
1076:
1077: \bibitem{Koutvitsky-Maslov} V.A. Koutvitsky and E.M. Maslov, Phys. Lett. A
1078: \textbf{336}, 31 (2005).
1079:
1080: \bibitem{B-BM3} I. Bialynicki-Birula and J. Mycielski, Phys. Scr. \textbf{20}%
1081: , 539 (1979).
1082:
1083: \bibitem{Magnus} W. Magnus and S. Winkler, \textit{Hill's Equation} (John
1084: Wiley \& Sons, New York, 1966).
1085:
1086: \bibitem{Whittaker} E.T. Whittaker and G.N. Watson, \textit{A Course of
1087: Modern Analysis} (University Press, Cambridge, 1927).
1088:
1089: \bibitem{Greene} P.B. Greene, L. Kofman, A. Linde, and A.A. Starobinsky,
1090: Phys. Rev. D \textbf{56}, 6175 (1997).
1091:
1092: \bibitem{Kaiser} D.I. Kaiser, Phys. Rev. D \textbf{57}, 702 (1998).
1093:
1094: \bibitem{Maslov4} E.M. Maslov and A.G. Shagalov, Contribution to the
1095: International Conference \textquotedblleft Solitons, Collapses and\textit{\ }%
1096: Turbulence,\textquotedblright\ Chernogolovka, Russia, 1999 [Physica D
1097: \textbf{152-153}, 769 (2001)].
1098:
1099: \bibitem{Finkel} F. Finkel, A. Gonz\'{a}lez-L\'{o}pez, A.L. Maroto, and M.A.
1100: Rodr\'{\i}guez, Phys. Rev. D \textbf{62}, 103515 (2000).
1101:
1102: \bibitem{Flugge} S. Fl\"{u}gge, \textit{Practical Quantum Mechanics, Vol. 1}
1103: (Springer, Berlin, 1971).
1104:
1105: \bibitem{Seidel} E. Seidel and W.-M. Suen, Phys. Rev. Lett. \textbf{66},
1106: 1659 (1991).
1107:
1108: \bibitem{Gleiser2} M. Gleiser and R.C. Howell, Phys. Rev. E \textbf{68},
1109: 065203(R) (2003).
1110:
1111: \bibitem{Gleiser3} M. Gleiser and R.C. Howell, Phys. Rev. Lett. \textbf{94},
1112: 151601 (2005).
1113:
1114: \bibitem{Barashenkov} I.V. Barashenkov and E.Yu. Panova, Physica D \textbf{69%
1115: }, 114 (1993).
1116:
1117: \bibitem{Landau} L.D. Landau and E.M. Lifshitz,\ \textit{Course of
1118: Theoretical Physics, Vol. 1. Mechanics }(Pergamon Press, Oxford, 1976)%
1119: \textit{.}
1120:
1121: \bibitem{Yakubovich} V.A. Yakubovich and V.M. Starzhinskii, \textit{Linear
1122: Differential Equations with Periodic Coefficients }(John Wiley \& Sons, New
1123: York, 1975).
1124: \end{thebibliography}
1125:
1126: \end{document}
1127: