nlin0703051/all.tex
1: \documentclass[multicol,aps,pre,epsf,amsmath,citesort,twocolumn]{revtex4}
2: \usepackage{graphicx,amsfonts,amsmath}
3: 
4: \def \BEA {\begin{eqnarray}}
5: \def \EEA {\end{eqnarray}}
6: \def \BC {\begin{cases}}
7: \def \EC {\end{cases}}
8: \def \ve {\varepsilon}
9: \def \d {\delta}
10: \def \l {\lambda}
11: \def \wt {\tilde{\omega}}
12: \def \Tt {\tilde{T}}
13: \def \tt {\tilde{t}}
14: \def \w {\omega}
15: \def \dw {\Delta\omega}
16: \def \a {\alpha}
17: \def \b {\beta}
18: \def \g {\gamma}
19: \def \at {\tilde{a}}
20: \def \af {\hat{a}}
21: \def \la {\langle}
22: \def \ra {\rangle}
23: \def \D {\Delta}
24: \def \tth {\tilde{\theta}}
25: \def \ffi {\varphi}
26: \def \i {\imath}
27: \def \eqN {\overset{N}{=}}
28: \def \eq1 {\overset{1}{=}}
29: \def \eqpi {\overset{\pi}{=}}
30: \def \p {\partial}
31: \def \kb {\bar{k}}
32: \def \eb {\bar{e}}
33: \def \ub {\bar{u}_2}
34: \def \T {\theta}
35: \def \tinf {\theta_{\infty}}
36: 
37: \begin{document}
38: 
39: \title{Interactions of renormalized waves in thermalized Fermi-Pasta-Ulam chains}
40: %
41: %Four-wave resonant interactions of renormalized waves in thermalized
42: %$\beta$-Fermi-Pasta-Ulam chains}
43: \author{Boris Gershgorin$^1$, Yuri V. Lvov$^1$ and David Cai$^2$, \\ \ \\
44: %
45: $^1$ \textit{Department of Mathematical Sciences, Rensselaer
46: Polytechnic
47: Institute, Troy, NY 12180}\\
48: %
49: $^2$ \textit{Courant Institute of Mathematical Sciences,
50:         New York University, New York, NY 10012}\\
51: }
52: \begin{abstract}
53: The dispersive interacting waves in Fermi-Pasta-Ulam (FPU) chains of
54: particles in \textit{thermal equilibrium} are studied from both
55: statistical and wave resonance perspectives. It is shown that, even
56: in a strongly nonlinear regime, the chain in thermal equilibrium can
57: be effectively described by a system of weakly interacting
58: \textit{renormalized} nonlinear waves that possess (i) the
59: Rayleigh-Jeans distribution and (ii) zero correlations between
60: waves, just as noninteracting free waves would. This renormalization
61: is achieved through a set of canonical transformations. The
62: renormalized linear dispersion of these renormalized waves is
63: obtained and shown to be in excellent agreement with numerical
64: experiments. Moreover, a dynamical interpretation of the
65: renormalization of the dispersion relation is provided via a
66: self-consistency, mean-field argument. It turns out that this
67: renormalization arises mainly from the trivial resonant wave
68: interactions, i.e., interactions with no momentum exchange.
69: Furthermore, using a multiple time-scale, statistical averaging
70: method, we show that the interactions of near-resonant waves give
71: rise to the broadening of the resonance peaks in the frequency
72: spectrum of renormalized modes. The theoretical prediction for the
73: resonance width for the thermalized $\beta$-FPU chain is found to be
74: in very good agreement with its numerically measured value.
75: \end{abstract}
76: \maketitle
77: \section{Introduction}
78: The study of discrete one-dimensional chains of particles with the
79: nearest-neighbor interactions provides insight to the dynamics of
80: various physical and biological systems, such as crystals, wave
81: systems, and biopolymers~\cite{fpu50, Peyrard, Toda}. In the thermal
82: equilibrium state, such nonlinear chains can be described by the
83: canonical Gibbs measure~\cite{LL5} with the Hamiltonian \BEA
84: H=\sum_j
85: \frac{p_j^2}{2}+\frac{(q_j-q_{j+1})^2}{2}+V(q_j-q_{j+1}),\label{intro_Hpq}
86: \EEA where $p_j$ and $q_j$ are the momentum and the displacement
87: from the equilibrium position of the $j$-th particle, respectively,
88: $V(q_j-q_{j+1})$ is the anharmonic part of the potential, and the
89: mass of each particle and the linear spring constant are scaled to
90: unity. In this article, we only consider the potentials of the
91: \textit{restoring} type, i.e., the potentials for which the Gibbs
92: measure exists. In order to study interactions of waves in such
93: systems, one usually introduces the canonical complex normal
94: variables $a_k$ via \BEA
95: a_k=\frac{P_k-\i\w_kQ_k}{\sqrt{2\w_k}},\label{intro_a} \EEA where
96: $P_k$ and $Q_k$ are the Fourier transforms of $p_j$ and $q_j$,
97: respectively, and $\w_k=2\sin(\pi k/N)$ is the linear dispersion
98: relation of the waves represented by $a_k$. In terms of the $a_k$,
99: the Hamiltonian~(\ref{intro_Hpq}) becomes \BEA
100: H=\sum\w_k|a_k|^2+V(a)\label{Ham_a}, \EEA where $V(a)$ is the
101: combination of various products of $a_k$ and $a_k^*$ corresponding
102: to various wave-wave interactions. If the potential in
103: Eq.~(\ref{intro_Hpq}) is harmonic, i.e., $V\equiv 0$, then $a_k$
104: correspond to ideal, \textit{free} waves, which have no energy
105: exchanges among different $k$ modes. In thermal equilibrium, the
106: Boltzmann distribution $ \exp(-\theta^{-1}\sum\w_k|a_k|^2) $ with
107: temperature $\T$, gives rise to the following properties of free
108: waves \BEA
109: \la a_k^*a_l\ra&=&n_k\delta^k_l,\label{intro_asa}\\
110: \la a_ka_l\ra&=&0\label{intro_aa}, \EEA for any $k$ and $l$, where
111: $n_k\equiv\la|a_k|^2\ra=\theta/\w_k$ is the power spectrum. If the
112: anharmonic part of the potential is sufficiently weak, then
113: corresponding waves $a_k$ remain almost free, and
114: Eqs.~(\ref{intro_asa}) and~(\ref{intro_aa}) would be approximately
115: satisfied in the weakly nonlinear regime. However, when the
116: nonlinearity becomes stronger, waves $a_k$ become strongly
117: correlated, and, in general, the correlations between waves
118: [Eq.~(\ref{intro_aa})] no longer vanish. In particular, $\la
119: a_ka_{N-k}\ra\neq 0$, as will be shown below. Naturally, the
120: question arises: can the strongly nonlinear system in thermal
121: equilibrium still be viewed as a system of almost free waves in some
122: statistical sense? In this article, we address this question with an
123: affirmative answer: it turns out that the system~(\ref{intro_Hpq})
124: can be described by a complete set of \textit{renormalized}
125: canonical variables $\at_k$, which still possess the wave properties
126: given by Eqs.~(\ref{intro_asa}) and~(\ref{intro_aa}) with a
127: renormalized linear dispersion. The waves that correspond to these
128: new variables $\at_k$ will be referred to as \textit{renormalized}
129: waves. Since these renormalized waves possess the equilibrium
130: Rayleigh-Jeans distribution~\cite{ZLF} and vanishing correlations
131: between waves, they resemble free, non-interacting waves, and can be
132: viewed as statistical normal modes. Furthermore, it will be
133: demonstrated that the renormalized linear dispersion for these
134: renormalized waves has the form $\wt_k=\eta(k)\w_k$, where $\eta(k)$
135: is the linear frequency renormalization factor, and is
136: \textit{independent} of $k$ as a consequence of the Gibbs measure.
137: 
138: In our method, the construction of the renormalized variables
139: $\at_k$ does not depend on a particular form or strength of the
140: anharmonic potential, as long as it is of the restoring type with
141: only the nearest neighbor interactions, as in Eq.~(\ref{intro_Hpq}).
142: Therefore, our approach is non-perturbative and can be applied to a
143: large class of systems with strong nonlinearity. However, in this
144: article, we will focus on the $\beta$-FPU chain to illustrate the
145: theoretical framework of the renormalized waves. We will verify that
146: $\at_k$ effectively constitute normal modes for the $\beta$-FPU
147: chain in thermal equilibrium by showing that (i) the theoretically
148: obtained renormalized linear dispersion relationship is in excellent
149: agreement with its dynamical manifestation in our numerical
150: simulation, and (ii) the equilibrium distribution of $\at_k$ is
151: still a Rayleigh-Jeans distribution and $\at_k$'s are uncorrelated.
152: Note that similar expressions for the renormalization factor $\eta$
153: have been previously discussed in the framework of an approximate
154: virial theorem~\cite{Alabiso} or effective long wave dynamics via
155: the Zwanzig-Mori projection~\cite{Lepri_renorm}. However, in our
156: theory, the exact formula for the renormalization factor is derived
157: from a \textit{precise} mathematical construction of statistical
158: normal modes, and is valid for all wave modes $k$ --- no longer
159: restricted to long waves.
160: 
161: Next, we address how renormalization arises from the dynamical wave
162: interaction in the $\beta$-FPU chain.
163: %Here, we will invoke the formalism of \textit{wave turbulence}~\cite{WT}.
164: %Wave turbulence theory provides a convenient way of describing resonant spectral energy transfer in a system of weakly interacting waves.
165: %This transfer occurs via the wave interactions on the resonance manifold, which is
166: %one of the main objects of study in wave turbulence.
167: We will show that the $\beta$-FPU chain can be \textit{effectively}
168: described as a four-wave interacting Hamiltonian system of the
169: renormalized resonant waves $\at_k$. We will study the resonance
170: structure of the $\beta$-FPU chain and find that most of the exact
171: resonant interactions are \textit{trivial}, i.e., the interactions
172: with no momentum exchange among different wave modes. In what
173: follows, the renormalization of the linear dispersion will be
174: explained as a collective effect of these trivial resonant
175: interactions of the renormalized waves $\at_k$. We will use a
176: self-consistency argument to find an approximation, $\eta_{sc}$, of
177: the renormalization factor $\eta$. As will be seen below, the
178: self-consistency argument essentially is of a mean-field type, i.e.,
179: the renormalization arises from the scattering of a wave by a
180: mean-background of waves in thermal equilibrium via trivial resonant
181: interactions. We note that our self-consistency, mean-field argument
182: is not limited to the weak nonlinearity. Very good agreement of the
183: renormalization factor $\eta$ and its dynamical approximation
184: $\eta_{sc}$ --- for weakly as well as strongly nonlinear waves ---
185: confirms that the renormalization is, indeed, \textit{a direct
186: consequence} of the trivial resonances.
187: 
188: 
189: We will further study the properties of these renormalized waves by
190: investigating how long these waves are coherent, i.e., what their
191: frequency widths are. Therefore, we consider \textit{near-resonant}
192: interactions of the renormalized waves $\at_k$, i.e., interactions
193: that occur in the vicinity of the resonance manifold, since most of
194: the exact resonant interactions are trivial, i.e., with no momentum
195: exchanges, and they, cannot effectively redistribute energy among
196: the wave modes.
197: 
198: We will demonstrate that near-resonant interactions of the
199: renormalized waves $\at_k$ provide a mechanism for effective energy
200: exchanges among different wave modes. Taking into account the
201: near-resonant interactions, we will study analytically the frequency
202: peak broadening of the renormalized waves $\at_k$ by employing a
203: multiple time-scale, statistical averaging method. Here, we will
204: arrive at a theoretical prediction of the spatiotemporal spectrum
205: $|\af_k(\w)|^2$, where $\af_k(\w)$ is the Fourier transform of the
206: normal variable $\at_k(t)$, and $\w$ is the frequency. The predicted
207: width of frequency peaks is found to be in good agreement with its
208: numerically measured values.
209: 
210: In addition, for a finite $\beta$-FPU chain, we will mention the
211: consequence, to the correlation times of waves, of the momentum
212: exchanges that cross over the first Brillouin zone. This process is
213: known as the umklapp scattering in the setting of phonon
214: scattering~\cite{Umklapp}. Note that, in the previous
215: studies~\cite{Kramer} of the FPU chain from the wave turbulence
216: point of view, the effects arising from the finite  nature of the
217: chain were not taken into account, i.e., only the limiting case of
218: $N\rightarrow\infty$, where $N$ is the system size, was considered.
219: 
220: The article is organized as follows. In
221: Section~\ref{sect_Hformalism}, we discuss a chain of particles with
222: the nearest-neighbor nonlinear interactions. We demonstrate how to
223: describe a strongly nonlinear system as a system of waves that
224: resemble free waves in terms of the power spectrum and vanishing
225: correlations between waves. We show how to construct the
226: corresponding renormalized variables with the renormalized linear
227: dispersion. In Section~\ref{sect_FPU}, we rewrite the $\beta$-FPU
228: chain as an interacting four-wave Hamiltonian system. We study the
229: dynamics of the chain numerically and find excellent agreement
230: between the renormalized dispersion, obtained analytically and
231: numerically. In Section~\ref{sect_Dispersion}, we describe the
232: resonance manifold analytically and illustrate its controlling role
233: in long-time averaged dynamics using numerical simulation. In
234: Section~\ref{sect_Renormalization}, we derive an approximation for
235: the renormalization factor for the linear dispersion using a
236: self-consistency condition. In Section~\ref{sect_Width}, we study
237: the broadening effect of frequency peaks  and predict analytically
238: the form of the spatiotemporal spectrum for the $\beta$-FPU chain.
239: We provide the comparison of our prediction with the numerical
240: experiment. We present the conclusions in
241: Section~\ref{sect_Conclusions}.
242: \section{Renormalized waves}
243: \label{sect_Hformalism} Consider a chain of particles coupled via
244: nonlinear springs. Suppose the total number of particles is $N$ and
245: the momentum and displacement from the equilibrium position of the
246: $j$-th particle are $p_j$ and $q_j$, respectively. If only the
247: nearest-neighbor interactions are present, then the chain can be
248: described by the Hamiltonian \BEA H=H_2+V,\label{Ham_pq} \EEA where
249: the quadratic part of the Hamiltonian takes the form \BEA
250: H_2=\frac{1}{2}\sum_{j=1}^N{p_j^2}+(q_j-q_{j+1})^2,\label{H2} \EEA
251: and the anharmonic potential $V$ is the function of the relative
252: displacement $q_j-q_{j+1}$. Here periodic boundary conditions
253: $q_{N+1}\equiv q_1$ and $p_{N+1}\equiv p_1$ are imposed. Since the
254: total momentum of the system is conserved, it can be set to zero.
255: 
256: In order to study the distribution of energy among the wave modes,
257: we transform the Hamiltonian to the Fourier variables $Q_k$, $P_k$
258: via \BEA \BC
259: Q_k=\displaystyle{\frac{1}{\sqrt{N}}}\sum_{j=0}^{N-1}q_je^{\frac{2\pi \i kj}{N}},\\
260: P_k=\displaystyle{\frac{1}{\sqrt{N}}}\sum_{j=0}^{N-1}p_je^{\frac{2\pi
261: \i kj}{N}}. \EC\label{PQ} \EEA This transformation is
262: canonical~\cite{LL1,Licht} and the Hamiltonian~(\ref{Ham_pq})
263: becomes \BEA
264: H&=&\frac{1}{2}\sum_{k=1}^{N-1}|P_k|^2+\omega_k^2|Q_k|^2+V(Q),\label{Ham_PQ}
265: \EEA where $\w_k=2\sin(\pi k/N)$ is the linear dispersion relation.
266: Note that, throughout the paper, for the simplicity of notation, we
267: denote the periodic wave number space by the set of integers in the
268: range $[0,N-1]$, i.e., we drop the conventional factor, $2\pi/N$.
269: The zeroth mode vanishes due to the fact that the total momentum is
270: zero.
271: 
272: If the system~(\ref{Ham_PQ}) is in thermal equilibrium, then the
273: canonical Gibbs measure, with the corresponding partition function
274: \BEA Z=\int_{-\infty}^{\infty}e^{-H(p,q)/\theta}dpdq,\label{Zpq}
275: \EEA with the temperature $\theta$, can be used to describe the
276: statistical behavior of the system. We consider the systems with the
277: anharmonic potential of the restoring type, i.e., the potential for
278: which the integral in Eq.~(\ref{Zpq}) converges. It can be easily
279: shown that for system~(\ref{Ham_PQ}) the average kinetic energy
280: $K_k$ of each mode is independent of the wave number \BEA \la
281: K_k\ra=\la K_l\ra,\label{Kkind} \EEA where $k$ and $l$ are wave
282: numbers, $K_k\equiv|P_k|^2/2$, and $\la\dots\ra$ denotes averaging
283: over the Gibbs measure. Similarly, the average quadratic potential
284: $U_k$ of each mode is independent of the wave number \BEA \la
285: U_k\ra=\la U_l\ra,\label{Qkind} \EEA where
286: $U_k\equiv\w_k^2|Q_k|^2/2$.
287: 
288: If the nonlinear interactions are weak, then it is convenient to
289: further transform the Hamiltonian~(\ref{Ham_PQ}) to the complex
290: normal variables defined by Eq.~(\ref{intro_a}). This transformation
291: is canonical, i.e., the dynamical equation of motion becomes \BEA
292: \i\dot{a}_k=\frac{\partial H}{\partial a_k^*}.\label{a_canonical}
293: \EEA In terms of these normal variables, the
294: Hamiltonian~(\ref{Ham_PQ}) takes the form~(\ref{Ham_a}). For the
295: system of noninteracting waves, i.e.,
296: $H=\sum_{k=1}^{N-1}\w_k|a_k|^2$, we obtain a standard virial theorem
297: in the form \BEA \la K_k\ra\vline_{V=0}=\la
298: U_k\ra\vline_{V=0}.\label{free} \EEA As a consequence of this virial
299: theorem, we have the properties of free waves, which were already
300: mentioned above [Eqs.~(\ref{intro_asa}) and~(\ref{intro_aa})], i.e.,
301: \BEA
302: \la a_k^*a_l\ra&=&\frac{1}{2\w_k}(\la|P_k|^2\ra+\w_k^2\la|Q_k|^2\ra)\delta^k_l=\frac{\theta}{\w_k}\delta^k_l,\label{asa}\\
303: \la
304: a_ka_l\ra&=&\frac{1}{2\w_k}(\la|P_k|^2\ra-\w_k^2\la|Q_k|^2\ra)\delta^{k+l}_N=0,\label{aa}
305: \EEA for all wave numbers $k$ and $l$. Note that
306: equation~(\ref{asa}) gives the classical Rayleigh-Jeans distribution
307: for the power spectrum of free waves~\cite{ZLF} \BEA
308: n_k=\frac{\T}{\w_k}.\label{PSa} \EEA However, if the nonlinearity is
309: present, the waves $a_k$ and $a_{N-k}$ become correlated, i.e., \BEA
310: \la
311: a_ka_{N-k}\ra=\frac{1}{2\w_k}(\la|P_k|^2\ra-\w_k^2\la|Q_k|^2\ra)\neq
312: 0,\label{akNk} \EEA since the property~(\ref{free}) is no longer
313: valid.
314: 
315: As we mentioned before, a complete set of new renormalized variables
316: $\at_k$ can be constructed, so that the strongly nonlinear system
317: can be viewed as a system of ``free'' waves in the sense of
318: vanishing correlations and the power spectrum, i.e., the new
319: variables $\at_k$ satisfy the properties of free waves given in
320: Eqs.~(\ref{asa}) and~(\ref{aa}). Next, we show how to construct
321: these renormalized variables $\at_k$.
322: 
323: Consider the generalization of the transformation~(\ref{intro_a}),
324: namely, the transformation from the Fourier variables $Q_k$ and
325: $P_k$ to the renormalized variables $\at_k$ by \BEA
326: \at_k=\frac{P_k-\i\wt_kQ_k}{\sqrt{2\wt_k}},\label{at} \EEA where
327: $\wt_k$ is an arbitrary function with the only restrictions \BEA
328: \wt_k>0,~~\wt_k=\wt_{N-k}.\label{canonical} \EEA One can show that,
329: these restrictions~(\ref{canonical}) provide a necessary and
330: sufficient condition for the transformation~(\ref{at}) to be
331: canonical. For the renormalized waves $\at_k$, we can compute \BEA
332: \la \at_k^*\at_l\ra&=&\frac{1}{2\wt_k}(\la|P_k|^2\ra+\wt_k^2\la|Q_k|^2\ra)\delta^k_l,\label{asat}\\
333: \la
334: \at_k\at_l\ra&=&\frac{1}{2\wt_k}(\la|P_k|^2\ra-\wt_k^2\la|Q_k|^2\ra)\delta^{k+l}_N.\label{aat}
335: \EEA Since we have the freedom of choosing any $\wt_k$ (with the
336: only restrictions~(\ref{canonical})), we can chose $\wt_k$ such that
337: $\la \at_k\at_{N-k}\ra$ vanishes. Thus, the renormalized variables
338: $\at_k$ for a strongly nonlinear system will behave like the bare
339: variables $a_k$ for a noninteracting system in terms of vanishing
340: correlations between waves. Therefore, we determine $\wt_k$ via \BEA
341: \la|P_k|^2\ra-\wt_k^2\la|Q_k|^2\ra=0.\label{vanish} \EEA Note that
342: the requirement~(\ref{vanish}) has the form of the virial theorem
343: for the free waves but with the renormalized linear dispersion
344: $\wt_k$. We rewrite Eq.~(\ref{vanish}) in terms of the kinetic and
345: quadratic potential parts of the energy of the mode $k$ as \BEA
346: \frac{\wt_k}{\w_k}=\sqrt{\frac{\la K_k\ra}{\la
347: U_k\ra}}.\label{vanish2} \EEA The  in Eqs.~(\ref{Kkind})
348: and~(\ref{Qkind}) leads to the $k$ independence of the right-hand
349: side of Eq.~(\ref{vanish2}). This allows us to define \textit{the
350: renormalization factor} $\eta$ for \textit{all} $k$'s by \BEA
351: \eta\equiv\frac{\wt_k}{\w_k}=\sqrt{\frac{\la K\ra}{\la
352: U\ra}}.\label{eta} \EEA for dispersion $\w_k$. Here
353: $K=\sum_{k=1}^{N-1}K_k$ and $U=\sum_{k=1}^{N-1}U_k$ are the kinetic
354: and the quadratic potential parts of the total energy of the
355: system~(\ref{Ham_PQ}), respectively. Note that the way of
356: constructing the renormalized variables $\at_k$ via the precise
357: requirement of vanishing correlations between waves yields the exact
358: expression for the renormalization factor, which is valid for all
359: wave numbers $k$ and any strength of nonlinearity. The independence
360: of $\eta$ of the wave number $k$ is a consequence of the Gibbs
361: measure. This $k$ independence phenomenon has been observed in
362: previous numerical experiments~\cite{our_prl,Alabiso}. We will
363: elaborate on this point in the results of the  numerical experiment
364: presented in Section~\ref{sect_FPU}.
365: 
366: The immediate consequence of the fact that $\eta$ is independent of
367: $k$ is that the power spectrum of the renormalized waves possesses
368: the precise Rayleigh-Jeans distribution, i.e., \BEA
369: \tilde{n}_k=\frac{\T}{\wt_k},\label{nt} \EEA from Eq.~(\ref{asat}),
370: where $\tilde{n}_k=\la |\at_k|^2\ra$. Combining Eqs.~(\ref{intro_a})
371: and~(\ref{at}), we find the relation between the ``bare'' waves
372: $a_k$ and the renormalized waves $\at_k$ to be \BEA
373: a_k=\frac{1}{2}\left(\sqrt{\eta}+\frac{1}{\sqrt{\eta}}\right)\at_k+\frac{1}{2}\left(\sqrt{\eta}-\frac{1}{\sqrt{\eta}}\right)\at_{N-k}.\label{a2at}
374: \EEA Using Eq.~(\ref{a2at}), we obtain the following form of the
375: power spectrum for the bare waves $a_k$ \BEA
376: n_k=\frac{1}{2}\left(1+\frac{1}{\eta^2}\right)\frac{\T}{\w_k},\label{nkbare}
377: \EEA which is a modified Rayleigh-Jeans distribution due to the
378: renormalization factor $(1+1/\eta^2)/2$. Naturally, if the
379: nonlinearity becomes weak, we have $\eta\rightarrow 1$, and,
380: therefore, all the variables and parameters with tildes reduce to
381: the corresponding ``bare'' quantities, in particular,
382: $\wt_k\rightarrow\w_k$, $\at_k\rightarrow a_k$,
383: $\tilde{n}_k\rightarrow n_k$. It is interesting to point out that,
384: even in a strongly nonlinear regime, the ``free-wave'' form of the
385: Rayleigh-Jeans distribution is satisfied
386: \textit{exactly}~[Eq.~(\ref{nt})] by the renormalized waves. Thus,
387: we have demonstrated that even in the presence of strong
388: nonlinearity, the system in thermal equilibrium can still be viewed
389: statistically as a system of ``free'' waves in the sense of
390: vanishing correlations between waves and the power spectrum.
391: 
392: Note that, in the derivation of the formula for the renormalization
393: factor [Eq.~(\ref{eta})], we only assumed the nearest-neighbor
394: interactions, i.e., the potential is the function of $q_j-q_{j+1}$.
395: One of the well-known examples of such a system is the $\beta$-FPU
396: chain, where only the forth order nonlinear term in $V$ is present.
397: In the remainder of the article, we will focus on the $\beta$-FPU to
398: illustrate the framework of the renormalized waves $\at_k$.
399: \section{Numerical study of the $\beta$-FPU chain}
400: \label{sect_FPU} Since its introduction in the early 1950s, the
401: study of the FPU lattice~\cite{FPU} has led to many great
402: discoveries in mathematics and physics, such as soliton
403: theory~\cite{Toda}. Being non-integrable, the FPU system also became
404: intertwined with the celebrated Kolmogorov-Arnold-Moser
405: theorem~\cite{Licht}. Here, we extend our results of the thermalized
406: $\beta$-FPU chain, which were briefly reported in~\cite{our_prl}.
407: 
408: The Hamiltonian of the $\beta$-FPU chain is of the form \BEA
409: H=\sum_{j=1}^N\frac{1}{2}p_j^2+\frac{1}{2}(q_j-q_{j+1})^2+\frac{\beta}{4}(q_j-q_{j+1})^4,\label{H_FPU}
410: \EEA where $\beta$ is a parameter that characterizes the strength of
411: nonlinearity.
412: 
413: The canonical equations of motion of the $\beta$-FPU chain are \BEA
414: \dot{q}_j&=&\frac{\partial H}{\partial p_j}=p_j,\label{dyn_pq_1}\\
415: \dot{p}_j&=&-\frac{\partial H}{\partial q_j}=(q_{j-1}-2q_j+q_{j+1})\nonumber\\
416: &+&\beta\big[(q_{j+1}-q_j)^3-(q_j-q_{j-1})^3\big].\nonumber \EEA To
417: investigate the dynamical manifestation of the renormalized
418: dispersion $\wt_k$ of $\at_k$, we numerically integrate
419: Eq.~(\ref{dyn_pq_1}). Since we study the thermal equilibrium
420: state~\cite{Ford,Alabiso_fpu,Lepri_fpu,Carati} of the $\beta$-FPU
421: chain, we use random initial conditions, i.e., $p_j$ and $q_j$ are
422: selected at random from the uniform distribution in the intervals
423: $(-p_{\rm{max}},p_{\rm{max}})$ and $(-q_{\rm{max}},q_{\rm{max}})$,
424: respectively, with the two constraints that (i) the total momentum
425: of the system is zero and (ii) the total energy of the system $E$ is
426: set to be a specified constant. We have verified that the results
427: discussed in the paper do not depend on details of the initial data.
428: Note that the behavior of $\beta$-FPU for fixed $N$ is fully
429: characterized by only one parameter $\beta E$~\cite{Poggy}. We use
430: the sixth order symplectic Yoshida algorithm~\cite{Yoshida} with the
431: time step $dt=0.01$, which ensures the conservation of the total
432: system energy up to the ninth significant digit for a runtime
433: $\tau=10^6$ time units. In order to confirm that the system has
434: reached the thermal equilibrium state~\cite{Parisi}, the value of
435: the energy localization~\cite{Cretegny} was monitored via
436: $L(t)\equiv{N\sum_{j=1}^{N}G_j^2}/{(\sum_{j=1}^{N}G_j)^2}$, where
437: $G_j$ is the energy of the $j$-th particle defined as \BEA
438: G_j&=&\frac{1}{2}p_j^2+\frac{1}{4}\big[(q_j-q_{j+1})^2+(q_{j-1}-q_j)^2\big]\nonumber\\
439: &+&\frac{\beta}{8}\big[(q_j-q_{j+1})^4+(q_{j-1}-q_j)^4\big]. \EEA If
440: the energy of the system is concentrated around one site, then
441: $L(t)=O(N)$. Whereas, if the energy is uniformly distributed along
442: the chain, then $L(t)=O(1)$. In our simulations, in thermal
443: equilibrium states, $L(t)$ is fluctuating in the range of $1$-$3$.
444: Since our simulation is of microcanonical ensemble, we have
445: monitored various statistics of the system to verify that the
446: thermal equilibrium state that is consistent with the Gibbs
447: distribution~(canonical ensemble) has been reached. Moreover, we
448: verified that, for $N$ as small as 32 and up to as large as 1024,
449: the equilibrium distribution in the thermalized state in our
450: microcanonical ensemble simulation is consistent with the Gibbs
451: measure. We compared the renormalization factor~(\ref{eta}) by
452: computing the values of $\la K\ra$ and $\la U\ra$ numerically and
453: theoretically using the Gibbs measure and found the discrepancy of
454: $\eta$ to be within $0.1\%$ for $\beta=1$ and the energy density
455: $E/N=0.5$ for $N$ from 32 to 1024.
456: 
457: We now address numerically how the renormalized linear dispersion
458: $\wt_k$ manifests itself in the dynamics of the $\beta$-FPU system.
459: We compute the spatiotemporal spectrum $|\af_k(\w)|^2$, where
460: $\af_k(\w)$ is the Fourier transform of $\at_k(t)$. (Note that, for
461: simplicity of notation, we drop a tilde in $\af_k$.)
462: \begin{figure}
463: \includegraphics[width=3in, height=2.5in]{awk_num}
464: \caption{The spatiotemporal spectrum $|\af_k(\w)|^2$ in thermal
465: equilibrium. The chain was modeled for $N=256$, $\beta=0.5$, and
466: $E=100$. [$\max\{-8,\ln{|\af_k(\w)|^2}\}$, with corresponding gray
467: scale, is plotted for a clear presentation]. The solid curve
468: corresponds to the usual linear dispersion $\w_k=2\sin(\pi k/N)$.
469: The dashed curve shows the locations of the actual frequency peaks
470: of $|\af_k(\w)|^2$.} \label{Fig_awk_num}
471: \end{figure}
472: Figure~\ref{Fig_awk_num} displays the spatiotemporal spectrum of
473: $\at_k$, obtained from the simulation of the $\beta$-FPU chain for
474: $N=256$, $\beta=0.5$, and $E=100$. In order to measure the value of
475: $\eta$ from the spatiotemporal spectrum, we use the following
476: procedure. For the fixed wave number $k$, the corresponding
477: renormalization factor $\eta(k)$ is determined by the location of
478: the center of the frequency spectrum $|\af_k(\w)|^2$, i.e., \BEA
479: \eta(k)=\frac{\w_{c}(k)}{\w_k},~~\mbox{with}~~
480: \w_{c}(k)=\frac{{\int}
481: \w|\af_k(\w)|^2~d\w}{\int|\af_k(\w)|^2~d\w}.\nonumber \EEA The
482: renormalization factor $\eta(k)$ of each wave mode $k$ is shown in
483: Fig.~\ref{Fig_et}~(inset). The numerical approximation $\bar{\eta}$
484: to the value of $\eta$ is obtained by averaging all $\eta(k)$, i.e.,
485: \BEA \bar{\eta}=\frac{1}{N-1}\sum_{k=1}^{N-1}\eta(k)\nonumber. \EEA
486: The renormalization factor for the case shown in
487: Fig.~\ref{Fig_awk_num} is measured to be $\bar{\eta}\approx1.1824$.
488: It can be clearly seen in Fig.~\ref{Fig_et}~(inset) that $\eta(k)$
489: is nearly independent of $k$ and its variations around $\bar{\eta}$
490: are less than $0.3\%$. We also compare the renormalization factor
491: $\eta$ obtained from Eq.~(\ref{eta}) (solid line in
492: Fig.~\ref{Fig_et}~(inset)) with its numerically computed
493: approximation $\bar{\eta}$ (dashed line in
494: Fig.~\ref{Fig_et}~(inset)). Equation~(\ref{eta}) gives the value
495: $\eta\approx 1.1812$ and the difference between $\eta$ and
496: $\bar{\eta}$ is less then $0.1\%$, which can be attributed to the
497: statistical errors in the numerical measurement.
498: \begin{figure}
499: \includegraphics[width=3in, height=2.5in]{etak_in}
500: \caption{The renormalization factor as a function of the
501: nonlinearity strength $\beta$. The analytical prediction
502: [Eq.~(\ref{eta})] is depicted with a solid line and the numerical
503: measurement is shown with circles. The chain was modeled for
504: $N=256$, and $E=100$. Inset: Independence of $k$ of the
505: renormalization factor $\eta(k)$. The circles correspond to
506: $\eta(k)$  obtained from the spatiotemporal spectrum shown in
507: Fig.~\ref{Fig_awk_num} [only even values of $k$ are shown for
508: clarity of presentation]. The dashed line corresponds to the mean
509: value $\bar{\eta}$. For $\beta=0.5$, the mean value of the
510: renormalization factor is found to be $\bar{\eta}\approx1.1824$. The
511: variations of $\eta_k$ around $\bar{\eta}$ are less then $0.3\%$.
512: [Note the scale of the ordinate.] The solid line corresponds to the
513: renormalization factor $\eta$ obtained from Eq.~(\ref{eta}). For the
514: given parameters $\eta\approx 1.1812$. } \label{Fig_et}
515: \end{figure}
516: In Fig.~\ref{Fig_et}, we plot the value of $\eta$ as a function of
517: $\beta$ for the system with $N=256$ particles and the total energy
518: $E=100$. The solid curve was obtained using Eq.~(\ref{eta}) while
519: the circles correspond to the value of $\eta$ determined via the
520: numerical spectrum $|\af_k(\w)|^2$ as discussed above. It can be
521: observed that there is excellent agreement between the theoretic
522: prediction and numerically measured values for a wide range of the
523: nonlinearity strength $\beta$.
524: 
525: In the following Sections, we will discuss how the renormalization
526: of the linear dispersion of the $\beta$-FPU chain in thermal
527: equilibrium can be explained from the wave resonance point of view.
528: In order to give a wave description of the $\beta$-FPU chain, we
529: rewrite the Hamiltonian~(\ref{H_FPU}) in terms of the renormalized
530: variables $\at_k$~[Eq.~(\ref{at})] with $\wt_k=\eta\w_k$, \BEA
531: H&=&\sum_{k=1}^{N-1}\frac{\w_k}{2}\left(\eta+\frac{1}{\eta}\right)|\at_k|^2\nonumber\\
532: &+&\frac{\w_k}{4}\left(\eta-\frac{1}{\eta}\right)(\at_k^*\at_{N-k}^*+\at_k\at_{N-k})\nonumber\\
533: &+&\sum_{k,l,m,s=1}^{N-1}T^{kl}_{ms}\Big[\D_{ms}^{kl}\at_k\at_l\at_m^*\at_s^*\nonumber\\
534: &+&\left(\frac{2}{3}\D^{klm}_{s}\at_k\at_l\at_m\at_s^*+c.c.\right)\nonumber\\
535: &+&\left(\frac{1}{6}\D^{klms}_{0}\at_k\at_l\at_m\at_s+c.c.\right)\Big],\label{H_a}
536: \EEA where c.c. stands for complex conjugate, and \BEA
537: T^{kl}_{ms}=\frac{3\beta}{8N\eta^2}\sqrt{\w_k\w_l\w_m\w_s}\label{T_inter}
538: \EEA is the interaction tensor coefficient. Note that, due to the
539: discrete nature of the system of finite size, the wave space is
540: periodic and, therefore, the ``momentum'' conservation is guaranteed
541: by the following ``periodic'' Kronecker delta functions \BEA
542: \D^{kl}_{ms}&\equiv&\delta_{ms}^{kl}-\delta^{klN}_{ms}-\delta^{kl}_{msN},\label{D22}\\
543: \D^{klm}_{s}&\equiv&\delta^{klm}_{s}-\delta^{klm}_{sN}+\delta^{klm}_{sNN},\label{D31}\\
544: \D^{klms}_{0}&\equiv&\delta^{klms}_{NN}-\delta^{klms}_{N}-\delta^{klms}_{NNN}.\label{D40}
545: \EEA Here, the Kronecker $\delta$-function is equal to 1, if the sum
546: of all superscripts is equal to the sum of all subscripts, and 0,
547: otherwise.
548: \section{Dispersion relation and resonances}
549: \label{sect_Dispersion} In order to address how the renormalized
550: dispersion arises from wave interactions, we study the resonance
551: structure of our nonlinear waves. Since the system~(\ref{H_a}) is a
552: Hamiltonian system with four-wave interactions, we will discuss the
553: properties of the resonance manifold associated with the $\beta$-FPU
554: system described by Eq.~(\ref{H_a}) as a first step towards the
555: understanding of its long time statistical behavior. We comment that
556: the resonance structure is one of the main objects of investigation
557: in wave turbulence
558: theory~\cite{ZLF,Newell,Lvov,Cai,MMT,Rink1,Rink2}. The theory of
559: wave turbulence focuses on the specific type of interactions, namely
560: resonant interactions, which dominate long time statistical
561: properties of the system. On the other hand, the non-resonant
562: interactions are usually shown to have a total vanishing average
563: contribution to a long time dynamics.
564: 
565: In analogy with quantum mechanics, where $a^+$ and $a$ are creation
566: and annihilation operators, we can view $\at_k^*$ as the
567: \textit{outgoing} wave with frequency $\wt_k$ and $\at_k$ as the
568: \textit{incoming} wave with frequency $\wt_k$. Then, the nonlinear
569: term $\at_k^*\at_l^*\at_m\at_s\Delta^{kl}_{ms}$ in
570: system~(\ref{H_a}) can be interpreted as the interaction process of
571: the type $(2\rightarrow 2)$, namely, two outgoing waves with wave
572: numbers $k$ and $l$ are ``created'' as a result of interaction of
573: the two incoming waves with wave numbers $m$ and $s$. Similarly,
574: $\at_k^*\at_l\at_m\at_s\Delta_{k}^{lms}$ in system~(\ref{H_a})
575: describes the interaction process of the type $(3\rightarrow 1)$,
576: that is, one outgoing wave with wave number $k$ is ``created'' as a
577: result of interaction of the three incoming waves with wave numbers
578: $l$, $m$, and $s$, respectively. Finally,
579: $\at_k\at_l\at_m\at_s\Delta_{0}^{klms}$ describes the interaction
580: process of the type $(4\rightarrow 0)$, i.e., all four incoming
581: waves interact and annihilate themselves. Furthermore, the complex
582: conjugate terms $\at_k\at_l^*\at_m^*\at_s^*\Delta_{lms}^{k}$ and
583: $\at_k^*\at_l^*\at_m^*\at_s^*\Delta^{0}_{klms}$ describe the
584: interaction processes of the type $(1\rightarrow 3)$ and
585: $(0\rightarrow 4)$, respectively.
586: 
587: Instead of the processes with the ``momentum'' conservation given
588: via the usual $\delta^{kl}_{ms}$, $\delta^{klm}_s$, or
589: $\delta^{klms}_0$ functions for an infinite discrete system, the
590: resonant processes of the $\beta$-FPU chain of a \textit{finite}
591: size are constrained to the manifold given by $\D^{kl}_{ms}$,
592: $\D^{klm}_s$, or $\D^{klms}_0$, respectively. Next, we describe
593: these resonant manifolds in detail. As will be pointed out in
594: Section~\ref{sect_Width}, there is a consequence of this
595: \textit{finite} size effect to the properties of the renormalized
596: waves.
597: 
598: The resonance manifold that corresponds to the $(2\rightarrow 2)$
599: resonant processes in the discrete periodic system, therefore, is
600: described by \BEA \BC
601: k+l\eqN m+s,\\
602: \wt_k+\wt_l=\wt_m+\wt_s, \EC\label{2to2} \EEA where we have
603: introduced the notation $g\eqN h$, which means that $g=h$, $g=h+N$,
604: or $g=h-N$ for any $g$ and $h$. The first equation in
605: system~(\ref{2to2}) is the ``momentum'' conservation condition in
606: the periodic wave number space. This ``momentum'' conservation comes
607: from $|\D^{kl}_{ms}|=1$. (Note that $|\D^{kl}_{ms}|$ can assume only
608: the value of $1$ or $0$.) Similarly, from $|\D^{klm}_{s}|=1$ and
609: $|\D^{klms}_{0}|=1$, the resonance manifolds corresponding to the
610: resonant processes of types $(3\rightarrow 1)$ and $(4\rightarrow
611: 0)$ are given by \BEA \BC
612: k+l+m\eqN s,\\
613: \wt_k+\wt_l+\wt_m=\wt_s, \EC\label{3to1} \EEA and \BEA \BC
614: k+l+m+s\eqN 0,\\
615: \wt_k+\wt_l+\wt_m+\wt_s=0, \EC\label{4to0} \EEA respectively. For
616: the processes of type $(3\rightarrow 1)$, the notation $g\eqN h$
617: means that $g=h$, $g=h+N$, or $g=h+2N$. For the $(4\rightarrow 0)$
618: processes, $g\eqN h$ means that $g=h+N$, $g=h+2N$, or $g=h+3N$.
619: 
620: To solve system (\ref{2to2}), we rewrite it in a continuous form
621: with $x={k}/{N}$, $y={l}/{N}$, $z={m}/{N}$, $v={s}/{N}$, which are
622: real numbers in the interval $(0,1)$. By recalling that
623: $\wt_k=2\eta\sin({\pi k}/{N})$, we have \BEA \BC
624: x+y\eq1 z+v,\\
625: \sin(\pi x)+\sin(\pi y)=\sin(\pi z)+\sin(\pi v). \EC\label{2to2cont}
626: \EEA Thus, any rational quartet that satisfies Eq.~(\ref{2to2cont})
627: yields a solution for Eq.~(\ref{2to2}). There are two distinct types
628: of the solutions of Eq.~(\ref{2to2cont}). The first one corresponds
629: to the case \BEA x+y=z+v,\nonumber \EEA whose only solution is given
630: by \BEA \BC
631: x=z,\\
632: y=v, \EC \mbox{or}~~ \BC
633: x=v,\\
634: y=z,\label{trivial} \EC \EEA i.e., these are \textit{trivial}
635: resonances, as we mentioned above. The second type of the resonance
636: manifold of the $(2\rightarrow 2)$-type interaction processes
637: corresponds to \BEA x+y=z+v\pm 1,\nonumber \EEA the solution of
638: which can be described by the following two branches \BEA
639: z_1&=&\frac{x+y}{2}+\frac{1}{\pi}\arcsin(A)+2j,\label{folded1}\\
640: z_2&=&\frac{x+y}{2}-1-\frac{1}{\pi}\arcsin(A)+2j,\label{folded2}
641: \EEA where
642: $A\equiv\tan\left(\pi({x+y})/{2}\right)\cos\left(\pi({x-y})/{2}\right)$
643: and $j$ is an integer. The second type of resonances arises from the
644: discreteness of our model of a finite length, leading to
645: \textit{non-trivial} resonances. For our linear dispersion here,
646: non-trivial resonances are only those resonances that involve wave
647: numbers crossing the first Brillouin zone. As mentioned above, in
648: the setting of the phonon physics, these non-trivial resonant
649: processes are also known as the umklapp scattering processes. In
650: Fig.~\ref{Fig_resonances}, we plot the solution of
651: Eq.~(\ref{2to2cont}) for $x={k}/{N}$ with the wave number $k=90$ for
652: the system with $N=256$ particles (the values of $k$ and $N$ are
653: chosen merely for the purpose of illustration). We stress that
654: \textit{all} the solutions of the system~(\ref{2to2cont}) are given
655: by the Eqs.~(\ref{trivial}),~(\ref{folded1}), and~(\ref{folded2}),
656: and that the non-trivial solutions arise only as a consequence of
657: discreteness of the finite chain. The curves in
658: Fig.~\ref{Fig_resonances} represent the loci of $(z,y)$,
659: parametrized by the fourth wave number $v$, i.e., $x$, $y$, $z$, and
660: $v$ form a resonant quartet, where $z={m}/{N}$, and $y={l}/{N}$.
661: Note that the fourth wave number $v$ is specified by the
662: ``momentum'' conservation, i.e., the first equation in
663: Eq.~(\ref{2to2cont}). The two straight lines in
664: Fig.~\ref{Fig_resonances} correspond to the trivial solutions, as
665: given by Eq.~(\ref{trivial}). The two curves (dotted and dashed)
666: depict the non-trivial resonances. Note that the dotted part of
667: non-trivial resonance curves corresponds to the
668: branch~(\ref{folded1}), and the dashed part corresponds to the
669: branch~(\ref{folded2}), respectively. An immediate question arises:
670: how do these resonant structures manifest themselves in the FPU
671: dynamics in the thermal equilibrium? By examining the
672: Hamiltonian~(\ref{H_a}), we notice that the resonance will control
673: the contribution of terms like
674: $\at_k^*\at_l^*\at_m\at_s\D^{kl}_{ms}$ in the long time limit.
675: Therefore, we address the effect of resonance by computing long time
676: average, i.e., $\la\at_k^*\at_l^*\at_m\at_s\ra\D^{kl}_{ms}$, and
677: comparing this average (Fig.~\ref{Fig_resonances_num}) with
678: Fig.~\ref{Fig_resonances}.
679: \begin{figure}
680: \includegraphics[width=2.5in, height=2.5in]{resonances}
681: \caption{The solutions of Eq.~(\ref{2to2cont}). The solid straight
682: lines correspond to the trivial resonances [solutions of
683: Eq.~(\ref{trivial})]. The solutions are shown for fixed $x=k/N$,
684: $k=90$, $N=256$ as the fourth wave number $v$ scans from ${1}/{N}$
685: to $(N-1)/{N}$ in the resonant quartet Eq.~(\ref{2to2cont}). The
686: non-trivial resonances are described by the dotted or dashed curves.
687: The dotted branch of the curves corresponds to the non-trivial
688: resonances described by Eq.~(\ref{folded1}) and and the dashed
689: branch corresponds to the non-trivial resonances described by
690: Eq.~(\ref{folded2}).} \label{Fig_resonances}
691: \end{figure}
692: \begin{figure}
693: \includegraphics[width=3in, height=2.5in]{numerical_resonance}
694: \caption{The long time average
695: $|\la\at_k^*\at_l^*\at_m\at_s\ra\D^{kl}_{ms}|$ of the $\beta$-FPU
696: system in thermal equilibrium. The parameters for the FPU chain are
697: $N=256$, $\beta=0.5$, and $E=100$.
698: $\la\at_k^*\at_l^*\at_m\at_s\ra\D^{kl}_{ms}$ was computed for fixed
699: $k=90$. The darker grayscale corresponds to the larger value of
700: $\la\at_k^*\at_l^*\at_m\at_s\ra\D^{kl}_{ms}$. The exact solutions of
701: Eq.~(\ref{2to2cont}), which are shown in Fig.~\ref{Fig_resonances},
702: coincide with the locations of the peaks of
703: $|\la\at_k^*\at_l^*\at_m\at_s\ra\D^{kl}_{ms}|$. Therefore, the
704: darker areas represent the near-resonance structure of the finite
705: $\beta$-FPU chain. (The two white lines show the locations, where
706: $s=0$ and, therefore, $\at_k^*\at_l^*\at_m\at_s\D^{kl}_{ms}=0$.)
707: [$\max\{2,\ln(|\la\at_k^*\at_l^*\at_m\at_s\ra\D^{kl}_{ms}|)\}$ with
708: the corresponding grayscale is plotted for a clean presentation].}
709: \label{Fig_resonances_num}
710: \end{figure}
711: To obtain Fig.~\ref{Fig_resonances_num}, the $\beta$-FPU system was
712: simulated with the following parameters: $N=256$, $\beta=0.5$,
713: $E=100$, and the averaging time window $\tau=400\tilde{t}_1$, where
714: $\tilde{t}_1$ is the longest linear period, i.e.,
715: $\tilde{t}_1={2\pi}/{\wt_1}$. In Fig.~\ref{Fig_resonances_num}, mode
716: $k$ was fixed with $k=90$ and the mode $s$, a function of $k$, $l$,
717: and $m$, is obtained from the constraint $k+l\eqN m+s$, i.e.,
718: $|\D^{kl}_{ms}|=1$. Note that we do not impose here the condition
719: $\wt_k+\wt_l=\wt_m+\wt_s$, therefore,
720: $|\la\at_k^*\at_l^*\at_m\at_s\ra\D^{kl}_{ms}|$ is a function of $l$
721: and $m$. By comparing Figs.~\ref{Fig_resonances} and
722: \ref{Fig_resonances_num}, it can be observed that the locations of
723: the peaks of the long time average
724: $|\la\at_k^*\at_l^*\at_m\at_s\ra\D^{kl}_{ms}|$ coincide with the
725: loci of the $(2\rightarrow 2)$-type resonances. This observation
726: demonstrates that, indeed, there are nontrivial $(2\rightarrow
727: 2)$-type resonances in the finite $\beta$-FPU chain in thermal
728: equilibrium. Furthermore, it can be observed in
729: Fig.~\ref{Fig_resonances_num} that, in addition to the fact that the
730: resonances manifest themselves as the locations of the peaks of
731: $|\la\at_k^*\at_l^*\at_m\at_s\ra\D^{kl}_{ms}|$, the structure of
732: \textit{near-resonances} is reflected in the finite width of the
733: peaks around the loci of the exact resonances. Note that, due to the
734: discrete nature of the finite $\beta$-FPU system, only those
735: solutions $x$, $y$, $z$, and $v$ of Eq.~(\ref{2to2cont}), for which
736: $Nx$, $Ny$, $Nz$, and $Nv$ are integers, yield solutions $k$, $l$,
737: $m$, and $s$ for Eq.~(\ref{2to2}). In general, the rigorous
738: treatment of the exact integer solutions of Eq.~(\ref{2to2}) is not
739: straightforward. For example, for $N=256$, we have the following two
740: exact quartets $\vec{k}=\{k,l,m,s\}$:
741: $\vec{k}=\{k,N/2-k,N/2+k,N-k\}$, $\vec{k}=\{k,N/2-k,N-k,N/2+k\}$ for
742: $k<N/2$, and $\vec{k}=\{k,3N/2-k,k-N/2,N-k\}$,
743: $\vec{k}=\{k,3N/2-k,N-k,k-N/2\}$ for $k>N/2$. We have verified
744: numerically that for $N=256$ there are no other exact integer
745: solutions of Eq.~(\ref{2to2}). In the analysis of the resonance
746: width in Section~\ref{sect_Width}, we will use the fact that the
747: number of exact non-trivial resonances [Eq.~(\ref{2to2})] is
748: significantly smaller than the total number of modes.
749: 
750: The broadening of the resonance peaks in
751: Fig.~\ref{Fig_resonances_num} suggests that, to capture the
752: near-resonances for characterizing long time statistical behavior of
753: the $\beta$-FPU system in thermal equilibrium, instead of
754: Eq.~(\ref{2to2}), one  needs to consider the following effective
755: system \BEA \BC
756: k+l\eqN m+s,\\
757: |\wt_k+\wt_l-\wt_m-\wt_s|<\Delta\w, \EC\label{2to2near} \EEA where
758: $0<\Delta\w\ll\wt_k$ for any $k$, and $\Delta\w$ characterizes the
759: resonance width, which results from the near-resonace structure.
760: Clearly, $\D\w$ is related to the broadening of the spectral peak of
761: each wave $\at_{\a}(t)$ with $\a=k,l,m$, or $s$ in the quartet, and
762: this broadening effect will be studied in detail in
763: Section~\ref{sect_Width}. Note that the structure of near-resonances
764: is a common characteristic of many periodic discrete nonlinear wave
765: systems~\cite{Lvov2,Colm, Pushkarev}.
766: 
767: Further, it is easy to show that the dispersion relation of the
768: $\beta$-FPU chain does not allow for the occurrence of
769: $(3\rightarrow 1)$-type resonances, i.e., there are no solutions for
770: Eq.~(\ref{3to1}), and, therefore, all the nonlinear terms
771: $\at_k^*\at_l\at_m\at_s\D^{k}_{lms}$ are non-resonant and their long
772: time average $\la\at_k^*\at_l\at_m\at_s\ra\D^{k}_{lms}$ vanishes. As
773: for the resonances of type $(4\rightarrow 0)$, since the dispersion
774: relation is non-negative, one can immediately conclude that the
775: solution of the system (\ref{4to0}) consists only of zero modes.
776: Therefore, the processes of type $(4\rightarrow 0)$ are also
777: non-resonant, giving rise to
778: $\la\at_k\at_l\at_m\at_s\ra\D_{0}^{klms}=0$. In this article, we
779: will neglect the higher order effects of the near-resonances of the
780: types $(3\rightarrow 1)$ and $(4\rightarrow 0)$.
781: 
782: In the following sections, we will study the effects of the resonant
783: terms of type $(2\rightarrow 2)$, namely, the linear dispersion
784: renormalization and the broadening of the frequency peaks of
785: $\at_k(t)$. It turns out that, the former is related to the trivial
786: resonance of type $(2\rightarrow 2)$ and the latter is related to
787: the near-resonances, as will be seen below.
788: \section{Self-consistency approach to frequency renormalization}
789: \label{sect_Renormalization} We now turn to the discussion of how
790: the trivial resonances give rise to the dispersion renormalization.
791: This question was examined in~\cite{our_prl} before. There, it was
792: shown that the renormalization of the linear dispersion of the
793: $\beta$-FPU chain arises due to the collective effect of the
794: nonlinearity. In particular, the trivial resonant interactions of
795: type $(2\rightarrow 2)$, i.e., the solutions of Eq.~(\ref{trivial}),
796: enhance the linear dispersion (the renormalized dispersion relation
797: takes the form $\wt_k=\eta\w_k$ with $\eta>1$), and effectively
798: weaken the nonlinear interactions. Here, we further address this
799: issue and present a self-consistency argument to arrive at an
800: approximation for the renormalization factor $\eta$. As it was
801: mentioned above, the contribution of the non-resonant terms have a
802: vanishing long time effect to the statistical properties of the
803: system, therefore, in our self-consistent approach, we ignore these
804: non-resonant terms. By removing the non-resonant terms and using the
805: canonical transformation \BEA \at_k=\frac{P_k-\i\eta_{sc}\w_k
806: Q_k}{\sqrt{2\eta_{sc}\w_k}},\label{a_sc} \EEA where $\eta_{sc}$ is a
807: factor to be determined, we arrive at a simplified effective
808: Hamiltonian from Eq.~(\ref{H_a}) for the finite $\beta$-FPU system
809: \BEA
810: H_{\rm{eff}}&=&\sum_{k=1}^{N-1}\frac{\w_k}{2}\left(\eta_{sc}+\frac{1}{\eta_{sc}}\right)|\at_k|^2\label{H_a_res}\\
811: &+&\sum_{k,l,m,s=1}^{N-1}T^{kl}_{ms}\Delta^{kl}_{ms}\at_k^*\at_l^*\at_m\at_s.\nonumber
812: \EEA The ``off-diagonal'' quadratic terms $\at_k\at_{N-k}$ from
813: Eq.~(\ref{H_a}) are not present in Eq.~(\ref{H_a_res}), since
814: $\at_k$ are chosen so that $\la\at_k\at_{N-k}\ra=0$ (see
815: Section~\ref{sect_Hformalism}). The contribution of the
816: \textit{trivial} resonances in $H_{\rm{eff}}$ is \BEA
817: H_4^{\rm{tr}}=4\sum_{k,l=1}^{N-1}T^{kl}_{kl}|\at_l|^2|\at_k|^2,\label{H4_diag}
818: \EEA which can be ``linearized'' in the sense that averaging the
819: coefficient in front of $|\at_k|^2$ in $H_4^{\rm{tr}}$ gives rise to
820: a quadratic form \BEA
821: H_2^{\rm{tr}}\equiv\sum_{k=1}^{N-1}\left(4\sum_{l=1}^{N-1}T^{kl}_{kl}\la|\at_l|^2\ra\right)|\at_k|^2.\nonumber
822: \EEA Note that the subscript $2$ in $H_2^{\rm{tr}}$ emphasizes the
823: fact that $H_2^{\rm{tr}}$ now can be viewed as a Hamiltonian for the
824: free waves with the familiar effective linear dispersion
825: $\Omega_k=4\sum_{l=1}^{N-1}T^{kl}_{kl}\la|\at_l|^2\ra$~\cite{our_prl,ZLF}.
826: This linearization is essentially a mean-field approximation, since
827: the long-time average of trivial resonances in Eq.~(\ref{H4_diag})
828: is approximated by the interaction of waves $\at_k$ with background
829: waves $\la |\at_l|\ra$. The self consistency condition, which
830: determines $\eta_{sc}$, can be imposed as follows: the quadratic
831: part of the Hamiltonian (\ref{H_a_res}), combined with the
832: ``linearized'' quadratic part, $H_2^{\rm{tr}}$, of the quartic
833: $H_4^{\rm{tr}}$, should be equal to an effective quadratic
834: Hamiltonian $\tilde{H}_2=\sum_{k=1}^{N-1}\wt_k|\at_k|^2$ for the
835: renormalized waves, i.e., \BEA
836: & &\sum_{k=1}^{N-1}\frac{\w_k}{2}\left(\eta_{sc}+\frac{1}{\eta_{sc}}\right)|\at_k|^2\nonumber\\
837: &
838: &+\sum_{k=1}^{N-1}\left(4\sum_{l=1}^{N-1}T^{kl}_{kl}\la|\at_l|^2\ra\right)|\at_k|^2=\sum_{k=1}^{N-1}\wt_k|\at_k|^2,\nonumber\\\label{self}
839: \EEA where $\wt_k$ is the renormalized linear dispersion, which is
840: used in the definition of our renormalized wave, Eq.~(\ref{a_sc}),
841: and $\wt_k=\eta_{sc}\w_k$. Equating the coefficients of
842: $\w_k|\at_k|^2$ on both sides for every wave number $k$ yields \BEA
843: \frac{1}{2}\left(\eta_{sc}+\frac{1}{\eta_{sc}}\right)+4\sum_{l=1}^{N-1}\frac{3\beta}{8N\eta_{sc}^2}\w_l\langle
844: |\at_l|^2\rangle=\eta_{sc},\nonumber \EEA where use is made of
845: Eq.~(\ref{T_inter}). After algebraic simplification, we have the
846: following equation for $\eta_{sc}$ \BEA
847: \eta_{sc}^3-\eta_{sc}=\frac{3\beta}{N}\sum_{l=1}^{N-1}\w_l\langle|\at_l|^2\rangle.\label{eq_eta}
848: \EEA Using the property~(\ref{asat}) of the renormalized normal
849: variables $\at_k$, we find the following dependence of
850: $\la|\at_k|^2\ra$ on $\eta_{sc}$, \BEA
851: \la|\at_l|^2\ra=\frac{1}{2\eta_{sc}\w_l}\left(\la|P_l|^2\ra+\eta_{sc}^2\w_l^2\la|Q_l|^2\ra\right).\label{rhs}
852: \EEA Combining Eqs.~(\ref{eq_eta}) and (\ref{rhs}) leads to \BEA
853: \eta_{sc}^4-A\eta_{sc}^2-B=0,\label{eq_eta2} \EEA where \BEA
854: A&=&1+\frac{3\beta}{2N}\sum_{l=1}^{N-1}\w_l^2\langle|Q_l|^2\rangle=1+\frac{3\beta}{N}\la U\ra,\nonumber\\
855: B&=&\frac{3\beta}{2N}\sum_{l=1}^{N-1}\langle|P_l|^2\rangle=\frac{3\beta}{N}\la
856: K\ra.\nonumber \EEA The only physically relevant solution of
857: Eq.~(\ref{eq_eta2}) is \BEA
858: \eta_{sc}=\sqrt{\frac{A+\sqrt{A^2+4B}}{2}}.\label{new_eta} \EEA The
859: constants $A$ and $B$ can be easily derived using the Gibbs measure.
860: 
861: Next, we compare the renormalization factor $\eta$ [Eq.~(\ref{eta})]
862: with its approximation $\eta_{sc}$ [Eq.~(\ref{new_eta})] from the
863: self-consistency argument. In Appendix~\ref{app_FPU}, we study in
864: detail the behavior of both $\eta$ and $\eta_{sc}$ in the two
865: limiting cases, i.e., when nonlinearity is small ($\beta\rightarrow
866: 0$ with fixed total energy $E$), and when nonlinearity is large
867: ($\beta\rightarrow\infty$ with fixed total energy $E$). As is shown
868: in Appendix~\ref{app_FPU}, for the case of small nonlinearity, both
869: $\eta$ and $\eta_{sc}$ have the same asymptotic behavior in the
870: first order of the small parameter $\beta$, \BEA
871: \eta&=&1+\frac{3E}{2N}\beta+O(\beta^2),\label{small_eta}\\
872: \eta_{sc}&=&1+\frac{3E}{2N}\beta+O(\beta^2).\nonumber \EEA Moreover,
873: in the case of strong nonlinearity $\beta\rightarrow\infty$, both
874: $\eta$ and $\eta_{sc}$ scale as $\beta^\frac{1}{4}$, i.e., \BEA
875: \eta\sim\eta_{sc}\sim\beta^{\frac{1}{4}}\label{large_eta} \EEA (see
876: Appendix~\ref{app_FPU} for details). Note that, in~\cite{our_prl},
877: we numerically obtained the scaling $\eta\sim\beta^{0.2}$, which
878: differs from the exact analytical result~(\ref{large_eta}) due to
879: statistical errors in the numerical estimate of the power.
880: \begin{figure}
881: \includegraphics[width=3in, height=3in]{eta_in}
882: \caption{The renormalization factor as a function of the
883: nonlinearity strength $\beta$ for small values of $\beta$. The
884: renormalization factor $\eta$ [Eq.~(\ref{eta})] is shown with the
885: solid line. The approximation $\eta_{sc}$ [Eq.~(\ref{new_eta})](via
886: the self-consistency argument) is depicted with diamonds connected
887: with the dashed line. The small-$\beta$ limit
888: [Eq.~(\ref{small_eta})] is shown with the solid circles connected
889: with the dotted line. Note that, abscissa is of logarithmic scale.
890: Inset: The renormalization factor as a function of the nonlinearity
891: strength $\beta$ for large values of $\beta$. The renormalization
892: factor $\eta$ [Eq.~(\ref{eta})] is shown with the solid line.
893: $\eta_{sc}$ [Eq.(\ref{new_eta})] is depicted with diamonds connected
894: with the dashed line. The large-$\beta$ scaling
895: [Eq.~(\ref{large_eta})] is shown with the dashed-dotted line. Note
896: that, the plot is of log-log scale with base 10.}
897: \label{Fig_eta_small}
898: \end{figure}
899: In Fig.~\ref{Fig_eta_small}, we plot the renormalization factor
900: $\eta$ and its approximation $\eta_{sc}$ for the case of small
901: nonlinearity $\beta$ for the system with $N=256$ particles and total
902: energy $E=100$. The solid line shows $\eta$ computed via
903: Eq.~(\ref{eta}), the diamonds with the dashed line represent the
904: approximation via Eq.~(\ref{new_eta}), and the solid circles with
905: the dotted line correspond to the small-$\beta$
906: limit~(\ref{small_eta}). In Fig.~\ref{Fig_eta_small}~(inset), we
907: plot the renormalization factor $\eta$ and its approximation
908: $\eta_{sc}$ for the case of large nonlinearity $\beta$ for the
909: system with $N=256$ particles and total energy $E=100$. The solid
910: line shows $\eta$ computed via Eq.~(\ref{eta}), the diamonds with
911: the dashed line represent the approximation via Eq.~(\ref{new_eta}),
912: and the dashed-dotted line correspond to the large-$\beta$
913: scaling~(\ref{large_eta}). Figure~\ref{Fig_eta_small} shows good
914: agreement between the renormalization factor $\eta$ and its
915: approximation $\eta_{sc}$ from the self consistency argument for a
916: wide range of nonlinearity, from $\beta\sim 10^{-3}$ to $\beta\sim
917: 10^4$. This agreement demonstrates, that (i) the effect of the
918: linear dispersion renormalization, indeed, arises mainly from the
919: trivial four-wave resonant interactions, and (ii) our
920: self-consistency, mean-field argument is not restricted to small
921: nonlinearity.
922: \section{Resonance width}
923: \label{sect_Width} Finally, we address the question of how coherent
924: these renormalized waves are, i.e., we study how the nonlinear
925: interactions of waves in thermal equilibrium broaden the
926: renormalized dispersion. We will obtain an analytical formula for
927: the spatiotemporal spectrum $|\af_k(\w)|^2$ for the $\beta$-FPU
928: chain and compare the numerically measured width of the frequency
929: peaks with the predicted width.
930: 
931: In the Hamiltonian~(\ref{H_a_res}), the nonlinear terms
932: corresponding to the trivial resonances have been absorbed into the
933: quadratic part via the effective renormalized dispersion $\wt_k$.
934: Therefore, the new effective Hamiltonian is \BEA
935: \bar{H}=\sum_{k=1}^{N-1}\wt_k|\at_k|^2+\sum_{k,l,m,s=1}^{N-1}\Tt^{kl}_{ms}\Delta^{kl}_{ms}\at_k^*\at_l^*\at_m\at_s,\label{Hef}
936: \EEA where \BEA \BC
937: \Tt^{kl}_{ms}=T^{kl}_{ms}=\displaystyle{\frac{3\beta}{8N\eta^2}}\sqrt{\w_k\w_l\w_m\w_s},~k\neq m,~\mbox{and}~k\neq s\\
938: \Tt^{kl}_{ms}=0,~~\mbox{otherwise.} \EC\label{Tt} \EEA The new
939: interaction coefficient $\Tt^{kl}_{ms}$ ensures that the terms that
940: correspond to the interactions with trivial resonances are not
941: doubly counted in the Hamiltonian~(\ref{Hef}) and are removed from
942: the quartic interaction. This new interactions in the quartic terms
943: include the exact non-trivial resonant and non-trivial near-resonant
944: as well as non-resonant interactions of the $(2\rightarrow 2)$-type.
945: 
946: We change the variables to the interaction picture by defining the
947: corresponding variables $b_k$ via \BEA
948: b_k=\at_ke^{\i\wt_kt},\nonumber \EEA so that, the dynamics governed
949: by the Hamiltonian (\ref{Hef}) takes the familiar form \BEA \i\dot
950: b_k=2\sum_{l,m,s=1}^{N-1}\Tt^{kl}_{ms}\Delta^{kl}_{ms}b_l^*b_mb_se^{\i\wt^{kl}_{ms}
951: t},\label{b_dyn} \EEA where
952: $\wt^{kl}_{ms}=\wt_k+\wt_l-\wt_m-\wt_s$~\cite{Lvov}. Without loss of
953: generality, we consider only the case of $k<N/2$. As we have noted
954: before, only for a very small number of quartets does
955: $\wt^{kl}_{ms}$ vanish \textit{exactly}, i.e., $\wt^{kl}_{ms}=0$. We
956: separate the terms on the RHS of Eq.~(\ref{b_dyn}) into two kinds
957: --- the first kind with $\wt^{kl}_{ms}=0$ that corresponds to exact
958: non-trivial resonances, and the second kind that corresponds to
959: non-trivial near-resonances and non-resonances. Since, in the
960: summation, the first kind contains far fewer terms than the second
961: kind, and all the terms are of the same order of magnitude, we will
962: neglect the first kind in our analysis. Therefore, Eq.~(\ref{b_dyn})
963: becomes \BEA \i\dot b_k&=&2{\sum_{l,m,s=1}^{N-1}}^{\prime}
964: \Tt^{kl}_{ms}\Delta^{kl}_{ms}b_l^*b_mb_se^{\i\wt^{kl}_{ms}
965: t},\label{b_dyn2} \EEA where the prime denotes the summation that
966: neglects the exact non-trivial resonances.
967: 
968: The problem of broadening of spectral peaks now becomes the study of
969: the frequency spectrum of the dynamical variables $b_k(t)$ in
970: thermal equilibrium. This is equivalent to study the  two-point
971: correlation in time of $b_k(t)$ \BEA C_k(t)=\la b_k(t)b_k^*(0)\ra,
972: \EEA where the angular brackets denote the thermal average, since,
973: by Wiener-Khinchin theorem, the frequency spectrum \BEA
974: |b(\w)|^2=\mathfrak{F}^{-1}[C(t)](\w),\label{WKh} \EEA where
975: $\mathfrak{F}^{-1}$ is the inverse Fourier transform in time. Under
976: the dynamics (\ref{b_dyn2}), time derivative of the two-point
977: correlation becomes \BEA
978: \dot{C}_k(t)&=&\la \dot{b}_k(t)b_k^*(0)\ra\nonumber \\
979: &=&\la -2\i{\sum_{l,m,s}}^{\prime}\Tt^{kl}_{ms}b_l^*(t)b_m(t)b_s(t)e^{\i\wt^{kl}_{ms}t}\Delta^{kl}_{ms}b_k^*(0)\ra\nonumber\\
980: &=&-2\i{\sum_{l,m,s}}^{\prime}\Tt^{kl}_{ms}e^{\i\wt^{kl}_{ms}t}J^{kl}_{ms}(t)\Delta^{kl}_{ms}\label{cdot},
981: \EEA where \BEA J^{kl}_{ms}(t)\equiv\la
982: b_l^*(t)b_k^*(0)b_m(t)b_s(t)\ra.\nonumber \EEA In order to obtain a
983: closed equation for $C_k(t)$, we need to study the evolution of the
984: fourth order correlator $J^{kl}_{ms}(t)$. We utilize the weak
985: effective nonlinearity in Eq.~(\ref{Hef})~\cite{our_prl} as the
986: small parameter in the following perturbation analysis and obtain a
987: closure for $C_k(t)$, similar to the traditional way of deriving
988: kinetic equation, as in~\cite{ZLF,Benney}. We note that the
989: effective interactions of renormalized waves can be weak, as we have
990: shown in~\cite{our_prl}, even if the $\beta$-FPU chain is in a
991: strongly nonlinear regime. Our perturbation analysis is a multiple
992: time-scale, statistical averaging method. Under the near-Gaussian
993: assumption, which is applicable for the weakly nonlinear wave fields
994: in thermal equilibrium, for the four-point correlator, we obtain
995: \BEA
996: J^{kl}_{ms}(t)\Delta^{kl}_{ms}=C_k(t)C_l(0)(\delta^k_m\delta^l_s+\delta^k_s\delta^l_m).~~\label{4corr}
997: \EEA Combining Eqs.~(\ref{Tt}) and~(\ref{4corr}), we find that the
998: right-hand side of Eq.~(\ref{cdot}) vanishes because \BEA
999: \Tt^{kl}_{ms}J^{kl}_{ms}(t)\Delta^{kl}_{ms}=0. \EEA Therefore, we
1000: need to proceed to the higher order contribution of
1001: $J^{kl}_{ms}(t)$. Taking its time derivative yields \BEA
1002: \dot{J}^{kl}_{ms}(t)\D^{kl}_{ms}&=&\la[\dot{b}_l^*(t)b_m(t)b_s(t)+b_l^*(t)\dot{b}_m(t)b_s(t)\nonumber\\
1003:                 &+&b_l^*(t)b_m(t)\dot{b}_s(t)]b_k^*(0)\ra\D^{kl}_{ms}.\label{Jdot}
1004: \EEA Considering the right-hand side of Eq.~(\ref{Jdot}) term by
1005: term, for the first term, we have \BEA
1006: & &\la \dot{b}_l^*(t)b_m(t)b_s(t)b_k^*(0)\ra\D^{kl}_{ms}\nonumber\\
1007: & &=\Bigg\la \Big[2\i{\sum_{\a,\b,\g}}^{\prime}\Tt^{l\a}_{\b\g}b_{\a}(t)b_{\b}^*(t)b_{\g}^*(t)\nonumber\\
1008: & &\times
1009: e^{-\i\w^{l\a}_{\b\g}}\D^{l\a}_{\b\g}\Big]b_m(t)b_s(t)b_k^*(0)\Bigg\ra\D^{kl}_{ms}.\label{4b}
1010: \EEA We can use the near-Gaussian assumption to split the correlator
1011: of the sixth order in Eq.~(\ref{4b}) into the product of three
1012: correlators of the second order, namely, \BEA
1013: & &\la b_k^*(0)b_m(t)b_s(t)b_{\a}(t)b_{\b}^*(t)b_{\g}^*\ra \D^{kl}_{ms}\nonumber\\
1014: & &=C_k(t)n_m
1015: n_s\delta^k_{\a}(\delta_{\b}^m\delta^s_{\g}+\delta^m_{\g}\delta^s_{\b}),\nonumber
1016: \EEA where we have used that $n_m=C_m(0)$. Then, Eq.~(\ref{4b})
1017: becomes \BEA
1018: & &\la \dot{b}_l^*(t)b_m(t)b_s(t)b_k^*(0)\ra\D^{kl}_{ms}\nonumber\\
1019: & &=4\i\Tt^{lk}_{ms}C_k(t)n_m n_s
1020: e^{-\i\wt^{lk}_{ms}}\D^{kl}_{ms}.\label{4b1} \EEA Similarly, for the
1021: remaining two terms in Eq.~(\ref{Jdot}), we have \BEA
1022: & &\la b_l^*(t)\dot{b}_m(t)b_s(t)b_k^*(0)\ra\D^{kl}_{ms}\nonumber\\
1023: & &=-4\i\Tt^{ms}_{kl}C_k(t)n_l n_s
1024: e^{\i\wt^{ms}_{kl}}\D^{kl}_{ms},\label{4b2} \EEA and \BEA
1025: & &\la b_l^*(t)b_m(t)\dot{b}_s(t)b_k^*(0)\ra\D^{kl}_{ms}\nonumber\\
1026: & &=-4\i\Tt^{ms}_{kl}C_k(t)n_l n_m
1027: e^{\i\wt^{ms}_{kl}}\D^{kl}_{ms},\label{4b3} \EEA respectively.
1028: Combining Eqs.~(\ref{4b1}), (\ref{4b2}), and (\ref{4b3}) with
1029: Eq.~(\ref{Jdot}), we obtain \BEA
1030: \dot{J}^{kl}_{ms}(t)\D^{kl}_{ms}&=&4\i\Tt^{kl}_{ms}C_k(t)e^{-\i\wt^{kl}_{ms}t}\D^{kl}_{ms}\nonumber\\
1031: & &\times(n_m n_s-n_l n_m-n_l n_s).\label{Jdot2} \EEA
1032: Equation~(\ref{Jdot2}) can be solved for $J^{kl}_{ms}(t)$ under the
1033: assumption that the term $e^{-\i\wt^{kl}_{ms}t}$ oscillates much
1034: faster than $C_k(t)$. We numerically verify [Fig.~\ref{Fig_corr}
1035: below] the validity of this assumption of time-scale separation.
1036: Under this approximation, the solution of Eq.~(\ref{Jdot2})  becomes
1037: \BEA
1038: J^{kl}_{ms}(t)\D^{kl}_{ms}&=&4\Tt^{kl}_{ms}C_k(t)\D^{kl}_{ms}\frac{e^{-\i\wt^{kl}_{ms}t}-1}{-\wt^{kl}_{ms}}\nonumber\\
1039: & &\times(n_m n_s-n_l n_m-n_l n_s )\label{J} \EEA Plugging
1040: Eq.~(\ref{J}) into Eq.~(\ref{cdot}), we obtain the following
1041: equation for $C_k(t)$ \BEA
1042: \dot{C}_k(t)&=&8\i C_k(t) {\sum_{l,m,s}}^{\prime} \left(\Tt^{kl}_{ms}\right)^2\D^{kl}_{ms}\frac{1-e^{\i\w^{kl}_{ms}t}}{\w^{kl}_{ms}}\nonumber\\
1043: & &\times(n_m n_s-n_l n_s-n_l n_m).\label{cdot2} \EEA Since in the
1044: thermal equilibrium $n_k$ is known, i.e., $n_k=\la
1045: |b_k(t)|^2\ra=\theta/\wt_k$ [Eq.~(\ref{nt})], Eq.~(\ref{cdot2})
1046: becomes a closed equation for $C_k(t)$. The solution of
1047: Eq.~(\ref{cdot2}) yields the autocorrelation function $C_k(t)$ \BEA
1048: \ln\frac{C_k(t)}{C_k(0)}&=&8{\sum_{l,m,s}}^{\prime} \left(\Tt^{kl}_{ms}\right)^2\frac{e^{\i\w^{kl}_{ms}t}-1-\i\w^{kl}_{ms}t}{(\w^{kl}_{ms})^2}\nonumber\\
1049: & &\times(n_l n_s+n_l n_m-n_m n_s)\D^{kl}_{ms}.\label{C} \EEA Using
1050: this observation, together with Eq.~(\ref{Tt}), finally, we obtain
1051: for the thermalized $\beta$-FPU chain \BEA
1052: \ln\frac{C_k(t)}{C_k(0)}&=&\frac{9\beta^2\theta^2}{8N^2\eta^6}\w_k{\sum_{l,m,s}}^{\prime}(\w_m+\w_s-\w_l)\D^{kl}_{ms}\nonumber\\
1053: &
1054: &\times\frac{e^{\i\w^{kl}_{ms}t}-1-\i\w^{kl}_{ms}t}{(\w^{kl}_{ms})^2}.\label{C}
1055: \EEA Equation~(\ref{C}) gives a direct way of computing the
1056: correlation function of the renormalized waves $\at_k$, which, in
1057: turn, allows us to predict the spatiotemporal spectrum
1058: $|\af_k(\w)|^2$. In Fig.~\ref{Fig_awk_both}(a), we plot the
1059: analytical prediction (via Eq.~(\ref{C})) of the spatiotemporal
1060: spectrum
1061: $|\af_k(\w)|^2\equiv|b_k(\w-\wt_k)|^2=\mathfrak{F}^{-1}[C(t)](\w-\wt_k)$.
1062: By comparing this plot with the one presented in
1063: Fig.~\ref{Fig_awk_both}(b), in which the corresponding numerically
1064: measured spatiotemporal spectrum is shown, it can be seen that the
1065: analytical prediction of the frequency spectrum via Eq.~(\ref{C}) is
1066: in good qualitative agreement with the numerically measured one.
1067: However, to obtain a more detailed comparison of the analytical
1068: prediction with the numerical observation, we show, in
1069: Fig.~\ref{Fig_aw}, the numerical frequency spectra of selected wave
1070: modes with the corresponding analytical predictions. It can be
1071: clearly observed that the agreement is rather good.
1072: \begin{figure}
1073: \includegraphics[width=3in, height=2.5in]{awk_both}
1074: \caption{ (a) Plot of the analytical prediction for the
1075: spatiotemporal spectrum $|\af_k(\w)|^2$ via Eq.~(\ref{C}). (b) Plot
1076: of the numerically measured spatiotemporal spectrum $|\af_k(\w)|^2$.
1077: The parameters in both plots were $N=256$, $\beta=0.125$, $E=100$
1078: and $\eta=1.06$, $\theta=0.401$. $\eta$ and  $\theta$ were computed
1079: analytically via Gibbs measure. The darker gray scale correspond to
1080: larger values of $|\af_k(\w)|^2$ in $\w$-$k$ space.
1081: [$\max\{-8,\ln{|\af_k(\w)|^2}\}$ is plotted for clear
1082: presentation].} \label{Fig_awk_both}
1083: \end{figure}
1084: \begin{figure}
1085: \includegraphics[width=3in, height=2.5in]{aw}
1086: \caption{Temporal frequency spectrum $|\af_k(\w)|^2$ for $k=30$
1087: (left peak) and $k=50$ (right peak). The numerical spectrum is shown
1088: with pluses and the analytical prediction [via Eq.~(\ref{C})] is
1089: shown with solid line. The parameters were $N=256$, $\beta=0.125$,
1090: $E=100$.
1091: %Both analytical and numerical peaks for each $k$ are centered at the same frequencies $\w$, which are equal to the numerically observed dispersion $\wt_k$.
1092: } \label{Fig_aw}
1093: \end{figure}
1094: One of the important characteristics of the frequency spectrum is
1095: the width of the spectrum. We compute the width $W(f)$ of the
1096: spectrum $f(\w)$ by \BEA W(f)=\frac{\int f(\w)~d\w}{\max_\w f(\w)}.
1097: \EEA In Fig.~\ref{Fig_widths}, we compare the width, as a function
1098: of the wave number $k$, of the frequency peaks from the numerical
1099: observation with that obtained from the analytical predictions. We
1100: observe that, for weak nonlinearity ($\beta=0.125$), the analytical
1101: prediction and the numerical observation are in excellent agreement.
1102: In the weakly nonlinear regime, this agreement can be attributed to
1103: the validity of (i) the near-Gaussian assumption, and (ii) the
1104: separation between the linear dispersion time scale and the time
1105: scale of the correlation $C_k(t)$. This separation was used in
1106: deriving the analytical prediction [Eq.~(\ref{C})]. However, when
1107: the nonlinearity becomes larger ($\beta=0.25$ and $\beta=0.5$), the
1108: discrepancy between the numerical measurements and the analytical
1109: prediction increases, as can be seen in Fig.~\ref{Fig_widths}.
1110: Nevertheless, it is important to emphasize that, even for very
1111: strong nonlinearity, our prediction is still qualitatively valid, as
1112: seen in Fig.~\ref{Fig_widths}. In order to find out the effect of
1113: the umklapp scattering due to the finite size of the chain, we also
1114: computed the correlation [Eq.~(\ref{C})] with the ``conventional''
1115: $\delta$-function $\delta^{kl}_{ms}$ (i.e., without taking into
1116: account the umklapp processes) instead of our ``periodic'' delta
1117: function $\D^{kl}_{ms}$. It turns out that the correlation time is
1118: approximately $30\%$ larger if it is computed without umklapp
1119: processes taken into account for the case $N=256$, $\beta=0.5$,
1120: $E=100$. It demonstrates that the influence of the non-trivial
1121: umklapp resonances is important and should be considered when one
1122: describes the dynamics of the finite length chain of particles.
1123: \begin{figure}
1124: \includegraphics[width=3in, height=2.5in]{widths}
1125: \caption{Frequency peak width $W(|\af_k(\w)|^2)$ as a function of
1126: the wave number $k$. The analytical prediction via Eq.~(\ref{C}) is
1127: shown with a dashed line and the numerical observation is plotted
1128: with solid circles. The parameters were $N=256$, $E=100$. The upper
1129: thick lines correspond to $\beta=0.5$, the middle fine lines
1130: correspond to $\beta=0.25$, and the lower solid circle and dashed
1131: line (almost overlap) correspond to $\beta=0.125$.}
1132: \label{Fig_widths}
1133: \end{figure}
1134: Finally, in Fig.~\ref{Fig_corr}, we verify the time scale separation
1135: assumption used in our derivation, i.e., the correlation time of the
1136: wave mode $k$ is sufficiently larger than the corresponding linear
1137: dispersion period $\tt_k=2\pi/\wt_k$. In the case of small
1138: nonlinearity ($\beta=0.125$), the two-point correlation changes over
1139: much slower time scale than the corresponding linear oscillations
1140: --- the correlation time is nearly two orders of magnitude larger than the corresponding linear oscillations for weak nonlinearity $\beta=0.125$, and
1141: nearly one order of magnitude larger than the corresponding linear
1142: oscillations for stronger nonlinearity $\beta=0.25$ and $\beta=0.5$.
1143: This demonstrates that the renormalized waves have long lifetimes,
1144: i.e., they are coherent over time-scales that are much longer than
1145: their oscillation time-scales.
1146: \begin{figure}
1147: \includegraphics[width=3in, height=2.5in]{awk_corr}
1148: \caption{ Ratio, as a function of $k$, of the correlation time
1149: $\tau_k$ of the mode $k$ to the corresponding linear period
1150: $\tt_k=2\pi/\wt_k$. Circles, squares, and diamonds represent the
1151: analytical prediction for $\beta=0.5$, $\beta=0.25$, and
1152: $\beta=0.125$ respectively. Stars, pentagrams, and triangles
1153: correspond to the numerical observation for $\beta=0.5$,
1154: $\beta=0.25$, and $\beta=0.125$ respectively. The parameters were
1155: $N=256$, $E=100$. The ratio is sufficiently large for all wave
1156: numbers $k$ even for relatively large $\beta=0.5$, which validates
1157: the time-scale separation assumption used in deriving Eq.~(\ref{J}).
1158: The comparison also suggests that for smaller $\beta$ the analytical
1159: prediction should be closer to the numerical observation, as is
1160: confirmed in Fig.~\ref{Fig_widths}.} \label{Fig_corr}
1161: \end{figure}
1162: \section{Conclusions}
1163: \label{sect_Conclusions} We have studied the statistical behavior of
1164: the nonlinear periodic lattice with the nearest neighbor
1165: interactions in thermal equilibrium. We have extended the notion of
1166: normal modes to the nonlinear system by showing that regardless of
1167: the strength of nonlinearity, the system in thermal equilibrium can
1168: still be effectively characterized by a complete set of renormalized
1169: waves, in the sense that those renormalized waves possess the
1170: Rayleigh-Jeans distribution and vanishing correlations between
1171: different wave modes. In addition, we have studied the property of
1172: dispersion relation of the renormalized waves. The results we
1173: obtained in Section~\ref{sect_Hformalism} are general and can be
1174: applied to the large class of nonlinear systems with the nearest
1175: neighbor interactions in thermal equilibrium.
1176: 
1177: We have further focused our attention on the particular system with
1178: the nearest neighbor interactions --- the famous FPU chain. We have
1179: confirmed that the general renormalization framework that we
1180: discussed above is consistent with the numerical observations. In
1181: particular, we have shown that the renormalized dispersion of the
1182: thermalized $\beta$-FPU chain is in excellent agreement with the
1183: numerical one for a wide range of the nonlinearity strength. We have
1184: further demonstrated that the renormalized dispersion is a direct
1185: consequence of the trivial resonant interactions of the renormalized
1186: waves. Using a self-consistency argument, we have found an
1187: approximation of the renormalization factor via a mean-field
1188: approximation. In addition, we have used the multiple time-scale,
1189: statistical averaging method to obtain the theoretical prediction of
1190: the spatiotemporal spectrum and demonstrated that the renormalized
1191: waves have long lifetimes. We note that the results obtained here
1192: can be extended to general nonlinear potentials with the nearest
1193: neighbor interactions.
1194: 
1195: The scenario of the wave behavior in the thermal equilibrium we
1196: obtained here may suggest a theoretical framework for the
1197: application of the wave turbulence to $\beta$-FPU in the situation
1198: of \textit{near}-equilibrium.
1199: \begin{acknowledgments}
1200: We thank Sergey Nazarenko and Naoto Yokoyama for discussions. Y.L.
1201: was supported by NSF CAREER DMS 0134955 and D.C. was supported by
1202: NSF DMS 0507901.
1203: \end{acknowledgments}
1204: \appendix
1205: \section{Limiting behaviors of $\eta$ for the thermalized $\beta$-FPU chain}
1206: \label{app_FPU} We change variables $y_j=q_j-q_{j+1}$ in the
1207: Hamiltonian~(\ref{H_FPU}) for $\beta$-FPU to obtain \BEA
1208: H(p,y)=\sum_{j=1}^N\left[\frac{1}{2}p_j^2+\frac{1}{2}y_j^2+\frac{\beta}{4}y_j^4\right].\label{app2_Ham_py}
1209: \EEA Next, we compute the pdf's for the momentum and displacement.
1210: Any $p_j$ is distributed with the Gaussian pdf
1211: $z_p=C_p\exp(-{\T}^{-1}p^2/2)$ and any $y_j$ is distributed with the
1212: pdf $z_y=C_y\exp(-{\T}^{-1}(y^2+\beta y^4/2)/2)$, where $C_p$ and
1213: $C_y$ are the normalizing constants. As we have discussed, the
1214: renormalization factor $\eta$ of the $\beta$-FPU system in thermal
1215: equilibrium is given by Eq.~(\ref{eta}), and its approximation via
1216: the self-consistency argument $\eta_{sc}$ is given by
1217: Eq.~(\ref{new_eta}). Here, we compare the behavior of both formulas
1218: in two limiting cases, i.e., the case of small nonlinearity
1219: $\beta\rightarrow 0$ and the case of strong nonlinearity
1220: $\beta\rightarrow\infty$. We will use the following expressions for
1221: the average density of kinetic, quadratic potential and quartic
1222: potential parts of the total energy of the system \BEA
1223: \frac{\la K\ra}{N}&=&\frac{1}{N}\sum_{j=1}^N\frac{\la p_j^2\ra}{2}=\frac{1}{2}\T,\label{app2_Hp2}\\
1224: \frac{\la U\ra}{N}&=&\frac{1}{N}\sum_{j=1}^N\frac{\la y_j^2\ra}{2}=\frac{1}{2}\la y^2\ra,\label{app2_Hy2}\\
1225: \frac{\la V\ra}{N}&=&\frac{\beta}{4N}\sum_{j=1}^N\la
1226: y_j^4\ra=\frac{\beta}{4}\la y^4\ra.\label{app2_Hy4} \EEA In a
1227: canonical ensemble, the temperature of a system is given by the
1228: temperature of the heat bath. By identifying the average energy
1229: density of the system with $\eb=E/N$ in our simulation (a
1230: microcanonical ensemble), we can determine $\T$ as a function of
1231: $\eb$ and $\beta$ by the following equation \BEA \frac{1}{N}\big(\la
1232: K\ra+\la U\ra+\la V\ra\big)=\eb.\label{app2_eqT} \EEA We start with
1233: the case of small nonlinearity $\beta\rightarrow 0$. Suppose in the
1234: first order of the small parameter $\beta$ the temperature has the
1235: following form \BEA \T(\beta)=\T_0+\beta \T_1\label{app2_Tsm}, \EEA
1236: where $\T_0=O(1)$ and $\T_1=O(1)$. We find the values of $\T_0$ and
1237: $\T_1$ using the constraint~(\ref{app2_eqT}). We use the following
1238: expansions in the small parameter $\beta$ \BEA
1239: & &\int_{-\infty}^{\infty}e^{-\frac{1}{2\T(\beta)}(y^2+\beta\frac{y^4}{2})}~dy\nonumber\\
1240: &
1241: &=\sqrt\frac{\pi}{8}\sqrt{\T_0}\left(4+\Big(\frac{2\T_1}{\T_0}-3\T_0\Big)\beta\right)+O(\beta^2),\label{app2_y0}
1242: \EEA \BEA
1243: & &\int_{-\infty}^{\infty}y^2e^{-\frac{1}{2\T(\beta)}(y^2+\beta\frac{y^4}{2})}~dy\nonumber\\
1244: &
1245: &=\sqrt\frac{\pi}{8}\sqrt{\T_0}\left(4\T_0+(6\T_1-15\T_0^2)\beta\right)+O(\beta^2),\label{app2_y2}
1246: \EEA \BEA
1247: & &\int_{-\infty}^{\infty}(y^2+\frac{\beta}{2}y^4)e^{-\frac{1}{2\T(\beta)}(y^2+\beta\frac{y^4}{2})}~dy\nonumber\\
1248: &
1249: &=\sqrt\frac{\pi}{8}\sqrt{\T_0}\left(4\T_0+(6\T_1-9\T_0^2)\beta\right)+O(\beta^2).\label{app2_y4}
1250: \EEA Then, in the first order in $\beta$, Eq.~(\ref{app2_eqT})
1251: becomes \BEA & &\T_0+\beta
1252: \T_1+\frac{\sqrt\frac{\pi}{8}\sqrt{\T_0}\left(4\T_0+(6\T_1-9\T_0^2)\beta\right)}{\sqrt\frac{\pi}{8}\sqrt{\T_0}\left(4+
1253: \Big(\frac{2\T_1}{\T_0}-3\T_0\Big)\beta\right)}=2\eb,\nonumber%\label{app2_eqT2}
1254: \EEA and we obtain $\T_0=\eb$ and $\T_1=(3/4)\eb^2$. Therefore, for
1255: the average kinetic energy density, we have \BEA \frac{\la
1256: K\ra}{N}=\frac{1}{2}\eb+\frac{3}{8}\eb^2\beta+O(\beta^2),\label{app2_Kb}
1257: \EEA and, for the average quadratic potential energy density, we
1258: have \BEA \frac{\la
1259: U\ra}{N}&=&\frac{1}{2}\frac{\sqrt\frac{\pi}{8}\sqrt{\T_0}\left(4\T_0+(6\T_1-15\T_0^2)\beta\right)}{\sqrt\frac{\pi}{8}\sqrt{\T_0}
1260: \left(4+\Big(\frac{2\T_1}{\T_0}-3\T_0\Big)\beta\right)}\nonumber\\
1261: &=&\frac{1}{2}\eb-\frac{9}{8}\eb^2\beta+O(\beta^2).\label{app2_U2b}
1262: \EEA Finally, we obtain Eq.~(\ref{small_eta}), i.e., for small
1263: $\beta$ \BEA
1264: \eta=1+\frac{3}{2}\eb\beta+O(\beta^2).\label{app2_eta_approx} \EEA
1265: Similarly, from Eq.~(\ref{new_eta}), we find the small $\beta$ limit
1266: of the approximation $\eta_{sc}$ \BEA
1267: \eta_{sc}=1+\frac{3}{2}\eb\beta+O(\beta^2).\nonumber \EEA Now, we
1268: consider the case of strong nonlinearity $\beta\rightarrow\infty$.
1269: From Eq.~(\ref{app2_eqT}), we conclude that temperature in the large
1270: $\beta$ limit, which we denote as $\tinf$, stays bounded, i.e., $
1271: 0<\tinf<2\eb, $ and, in the limit of large $\beta$, we obtain for
1272: Eq.~(\ref{app2_eqT}) \BEA
1273: \tinf+\frac{\int_{-\infty}^{\infty}\frac{\beta}{2}y^4e^{-\frac{\beta}{4\tinf}y^4}~dy}{\int_{-\infty}^{\infty}e^{-\frac{\beta}{4\tinf}y^4}~dy}=2\eb.
1274: \label{app2_eqT3} \EEA After performing the integration, we obtain $
1275: \tinf=(4/3)\eb, $ and the average kinetic energy density becomes $
1276: \la K\ra/N=(2/3)\eb $. For the average quadratic potential energy
1277: density, we have \BEA
1278: \frac{\la U\ra}{N}=\frac{1}{2}\frac{\int_{-\infty}^{\infty}y^2e^{-\frac{\beta}{4\tinf}y^4}~dy}{\int_{-\infty}^{\infty}e^{-\frac{\beta}{4\tinf}y^4}~dy}%=\nonumber\\
1279: %\frac{1}{2}\frac{\sqrt{2}\Gamma(\frac{3}{4})/\left(\frac{\beta}{\tinf}\right)^{\frac{3}{4}}}
1280: %{\Gamma(\frac{1}{4})/\Big(\sqrt{2}\left(\frac{\beta}{\tinf}\right)^{\frac{1}{4}}\Big)}
1281: =\frac{\Gamma(\frac{3}{4})}{\Gamma(\frac{1}{4})}\left(\frac{4\eb}{3\beta}\right)^{\frac{1}{2}}
1282: \EEA For the renormalization factor, we obtain the following large
1283: $\beta$ scaling \BEA
1284: \eta=\sqrt{\frac{\Gamma(\frac{3}{4})}{\sqrt{3}\Gamma(\frac{1}{4})}}\eb^\frac{1}{4}\beta^\frac{1}{4}.\label{app2_etainf}
1285: \EEA Similarly, for the approximation of $\eta_{sc}$, we obtain
1286: $A=C\sqrt{\eb\beta}$, $B=4\eb\beta$, and
1287: $C=2\sqrt{3}\Gamma(3/4)/\Gamma(1/4)$. Therefore, the large $\beta$
1288: scaling of $\eta_{sc}$ becomes \BEA
1289: \eta_{sc}=\sqrt{\frac{C+\sqrt{C^2+16}}{2}}\eb^\frac{1}{4}\beta^\frac{1}{4},\label{app2_etascinf}
1290: \EEA which yields Eq.~(\ref{large_eta}).
1291: \begin{thebibliography}{2000}
1292: 
1293: \bibitem{fpu50} Focus issue: The Fermi-Pasta-Ulam problem, the first 50 years, Chaos {\bf 15} (2005).
1294: 
1295: \bibitem{Peyrard} M. Peyrard, Nonlinearity {\bf 17}, R1 (2004), and references therein.
1296: 
1297: \bibitem{Toda} M. Toda, Theory of Nonlinear Lattice (Springer-Verlag, New York, 1989).
1298: 
1299: \bibitem{LL5} L. Landau and E. Lifshitz, {\it Statistical Physics}, Course of Theoretical Physics Vol. 5 (Pergamon, Oxford, 1981).
1300: 
1301: \bibitem{ZLF} V.E. Zakharov, V.S. Lvov,  and G. Falkovich, Kolmogorov Spectra of Turbulence (Springer-Verlag, Berlin, 1992).
1302: 
1303: \bibitem{Alabiso} C. Alabiso, M. Casartelli, P. Marenzoni, J. Stat. Phys. {\bf 79}, 451 (1995).
1304: 
1305: \bibitem{Lepri_renorm} S. Lepri, Phys. Rev. E {\bf 58}, 7165 (1998)
1306: 
1307: \bibitem{Umklapp} U. Z\"ulicke and A. H. MacDonald, Physica (Amsterdam) {\bf 6E}, 104 (2000) and references therein.
1308: 
1309: \bibitem{Kramer} J. Biello, P. Kramer, Y. Lvov, Proc. of the fourth international conf. on dyn. sys. and diff. eqns., 113 (2001);
1310: P. Kramer, J. Biello, Y. Lvov, \textit{ibid.}, 482 (2003).
1311: 
1312: \bibitem{LL1} L. Landau and E. Lifshitz, {\it Mechanics}, Course of Theoretical Physics Vol. 1 (Pergamon, Oxford, 1976).
1313: 
1314: \bibitem{Licht} A. J. Lichtenberg and M. A. Lieberman, {\it Regular and Chaotic Dynamics}, 2nd ed. (Springer-Verlag, Berlin, 1992).
1315: 
1316: \bibitem{our_prl} B. Gershgorin, Y.V. Lvov, D. Cai, Phys. Rev. Lett. {\bf 95}, 264302 (2005).
1317: 
1318: \bibitem{FPU} E. Fermi, J. Pasta, and S. Ulam, Los Alamos Scientific Laboratory Report No. LA-1940 (reprinted
1319: in Fermi E. Collected papers by University of Chicago Press,
1320: Chicago, 1965, Vol II, p 978).
1321: 
1322: \bibitem{Ford} J. Ford, Phys. Rep., {\bf 213}, 271, (1992).
1323: 
1324: \bibitem{Alabiso_fpu} C. Alabiso and M. Casartelli, J. Phys. A: Math. Gen. {\bf 34}, 1223 (2001).
1325: 
1326: \bibitem{Lepri_fpu} S. Lepri, R. Livi, A. Politi, Chaos {\bf 15}, 015118 (2005).
1327: 
1328: \bibitem{Carati} A. Carati, L. Galgani, A. Giorgilli, Chaos {\bf 15}, 015105 (2005).
1329: 
1330: \bibitem{Poggy} P. Poggy and S. Ruffo, Physica (Amsterdam) {\bf 103D}, 251 (1997).
1331: 
1332: \bibitem{Yoshida} H. Yoshida, Phys. Lett. A {\bf 150}, 262 (1990).
1333: 
1334: \bibitem{Parisi} G. Parisi, Europhysics Lett. 40, 357 (1997).
1335: 
1336: \bibitem{Cretegny} T. Cretegny et al., Physica (Amsterdam) {\bf 121D}, 109 (1998).
1337: 
1338: \bibitem{Newell} A. Newell, S. Nazarenko, and L. Biven, Physica (Amsterdam) {\bf 152D}, 520 (2001)
1339: % Wave turbulence and intermittency
1340: 
1341: \bibitem{Lvov} Y.V. Lvov and S. Nazarenko, Phys. Rev. E {\bf 69}, 066608 (2004);
1342:                Y. Choi, Y.V. Lvov, S. Nazarenko, and B. Pokorni, Phys. Lett. A {\bf 339}, 361 (2005).
1343: 
1344: \bibitem{Cai} D. Cai et al., Physica (Amsterdam) {\bf 152D}, 551 (2001).
1345: %Dispersive wave turbulence in one dimension
1346: 
1347: \bibitem{MMT} A.J. Majda, D.W. McLaughlin, and E.G. Tabak, J. Nonlinear Sci. {\bf 6}, 9 (1997).
1348: 
1349: \bibitem{Rink1} B. Rink, Comm. Math. Phys. {\bf 218(3)}, 665 (2001).
1350: % Symmetry and resonances in periodic FPU chains
1351: 
1352: \bibitem{Rink2} B. Rink,  Physica (Amsterdam) {\bf 175D}, 31 (2003).
1353: %Symmetric invariant manifolds in the Fermi-Pasta-Ulam lattice
1354: 
1355: \bibitem{Lvov2} Y.V. Lvov, S. Nazarenko, B. Pokorni, Physica (Amsterdam) {\bf 218D}, 24 (2006).
1356: 
1357: \bibitem{Colm} C. Connaughton, S.V. Nazarenko, and A.N. Pushkarev, Phys. Rev. E. {\bf 63}, 046306 (2001).
1358: 
1359: \bibitem{Pushkarev} A. Pushkarev, Eur. J. Mech. B/fluids {\bf 18(3)}, 345 (1999).
1360: %On the Kolmogorov frozen turbulence in numerical simulation of capillary waves
1361: 
1362: \bibitem{Benney} J. Benney and A. Newell, Stud. in Appl. Math. {\bf 48}, 29 (1969).
1363: % random wave closure
1364: \end{thebibliography}
1365: \end{document}
1366: