1: \documentclass[runningheads]{svmult}
2: \usepackage{epsfig}
3: \usepackage{makeidx} % allows index generation
4: \usepackage{graphicx} % standard LaTeX graphics tool
5: % for including eps-figure files
6: \usepackage{subeqnar} % subnumbers individual equations
7: % within an array
8: \usepackage{multicol} % used for the two-column index
9: %\usepackage{cropmark} % cropmarks for pages without
10: % pagenumbers - only needed when manuscript
11: % is printed from paper and not from data
12: \usepackage{physprbb} % modified textarea for proceedings,
13: % lecture notes, and the like.
14: \makeindex % used for the subject index
15: % please use the style sprmidx.sty with
16: % your makeindex program
17:
18: %%upright Greek letters (example below: upright "mu")
19: %%\newcommand{\greeksym}[1]{{\usefont{U}{psy}{m}{n}#1}}
20: %%\newcommand{\umu}{\mbox{\greeksym{m}}}
21: %%\newcommand{\udelta}{\mbox{\greeksym{d}}}
22: %%\newcommand{\uDelta}{\mbox{\greeksym{D}}}
23: %%\newcommand{\uPi}{\mbox{\greeksym{P}}}
24: %%{\tiny} {\scriptsize} {\footnotesize} {\small} {\normalsize}
25: %%{\large} {\Large} {\LARGE}
26: %\oddsidemargin=-.8cm
27:
28:
29: % FIGURES
30: \def\chevrons{\epsfig{figure=instab.ps,height=13truecm}}%,width=3truecm}}
31: \def\schock{\epsfig{figure=schock.ps,height=6truecm}}
32: \def\density{\epsfig{figure=density.ps,height=6truecm}}
33: \def\spirale{\epsfig{figure=spirale.ps,height=7truecm}}
34: \def\chevronsburgers{\epsfig{figure=chevronsburgers.ps,height=12truecm}}
35: \def\herisson{\epsfig{figure=herisson.ps,height=8truecm}}
36: \def\excitationparametrique{\epsfig{figure=excitationparametrique.ps,height=8truecm}}
37: \def\excitationparametriquebis{\epsfig{figure=excitationparametriquebis.ps,height=5truecm}}
38:
39:
40: \begin{document}
41: \title*{The Hamiltonian Mean Field Model:
42: from Dynamics to Statistical Mechanics and back}
43:
44: %\toctitle{The Hamiltonian Mean Field Model: from Dynamics to
45: % Statistical Mechanics and back}
46:
47: \titlerunning{The Hamiltonian Mean Field Model}
48:
49: \author{Thierry Dauxois\inst{1}
50: \and Vito Latora\inst{2} \and Andrea Rapisarda\inst{2} \and
51: Stefano Ruffo\inst{1,3}
52: \and Alessandro Torcini\inst{3,4}}
53:
54: \authorrunning{T. Dauxois et al.}
55: % if there are more than two authors,
56: % please abbreviate author list for running head
57: \institute{Laboratoire de Physique, UMR CNRS 5672, ENS Lyon, 46,
58: all{\'e}e d'Italie, F-69007 Lyon, France \and Dipartimento di Fisica,
59: Universit{\'a} di Catania, and Istituto Nazionale di Fisica Nucleare,
60: Sezione di Catania, Corso Italia 57, I-95129 Catania, Italy \and
61: Dipartimento di Energetica "S. Stecco", Universit{\`a} di Firenze, via
62: S. Marta, 3, INFM and INFN, I-50139 Firenze, Italy \and
63: UMR-CNRS 6171, Universit{\'e} d'Aix-Marseille III -
64: Av. Esc. Normandie-Niemen, F-13397 Marseille Cedex 20, France.}
65:
66:
67:
68: \maketitle % typesets the title of the contribution
69:
70: \begin{abstract}
71: The thermodynamics and the dynamics of particle systems with
72: infinite-range coupling display several unusual and
73: new features with respect to systems with short-range
74: interactions. The Hamiltonian Mean Field (HMF) model
75: represents a paradigmatic example of this class of systems.
76: The present study addresses both attractive and repulsive
77: interactions, with a particular emphasis on the description
78: of clustering phenomena from a thermodynamical as well as
79: from a dynamical point of view. The observed
80: clustering transition can be first or
81: second order, in the usual thermodynamical sense. In the former
82: case, ensemble inequivalence naturally arises close to the
83: transition, i.e. canonical and microcanonical ensembles give
84: different results. In particular, in the microcanonical ensemble
85: negative specific heat regimes and temperature jumps are
86: observed.
87: Moreover, having access to dynamics one can study non-equilibrium
88: processes. Among them, the most striking is
89: the emergence of coherent structures in the repulsive model,
90: whose formation and dynamics can be studied either by using the tools of
91: statistical mechanics or as a manifestation of the
92: solutions of an associated Vlasov equation.
93: The chaotic character of the HMF model has been also analyzed in terms
94: of its Lyapunov spectrum.
95: \end{abstract}
96:
97:
98: \section{Introduction}
99: Long-range interactions appear in the domains of
100: gravity~\cite{paddyhmf,chavanishmf} and of plasma physics~\cite{elskenshmf} and
101: make the statistical treatment extremely complex. Additional features
102: are present in such systems at short distances: the
103: gravitational potential is singular at the origin and screening
104: phenomena mask the Coulomb singularity in a plasma. This justifies the
105: introduction of simplified toy models that retain only the long-range
106: properties of the force, allowing a detailed description of the
107: statistical and dynamical behaviors associated to this feature.
108: In this context a special role is
109: played by mean-field models, i.e. models where all particles interact
110: with the same strength. This constitutes a dramatic reduction of
111: complexity, since in such models the spatial coordinates have no role,
112: since each particle is equivalent. However, there are several
113: preliminary indications that behaviors found in mean-field models
114: extend to cases where the two-body potential decays at large distances
115: with a power smaller than space dimension~\cite{tamahmf,BMRhmf,campa}.
116:
117:
118: The
119: Blume-Emery-Griffiths (BEG) mean-field model is discussed in
120: this book~\cite{BMRhmf} and represents an excellent benchmark to discuss
121: relations between canonical and microcanonical ensembles. Indeed,
122: this model is exactly solvable in both ensembles and is, at the same
123: time, sufficiently rich to display such interesting features as
124: negative specific heat and temperature jumps in the microcanonical
125: ensemble. Since these effects cannot be present in the canonical
126: ensemble, this rigorously proves ensemble inequivalence. However, the
127: BEG model has no dynamics and only the thermodynamical behavior can be
128: investigated. Moreover, it is a spin model where variables take discrete
129: values. It would therefore be amenable to introduce a model that
130: displays all these interesting thermodynamical effects, but for which
131: one would also dispose of an Hamiltonian dynamics with continuous
132: variables, whose equilibrium states could be studied both in the
133: canonical and in the microcanonical ensemble. Having access to
134: dynamics, one could moreover study non equilibrium features and
135: aspects of the microscopic behavior like sensitivity to initial
136: conditions, expressed by the Lyapunov spectrum~\cite{Ruelleeckman}.
137: Such a model has been introduced in Ref.~\cite{Antoni} and has been
138: called the Hamiltonian Mean Field (HMF) model. In the simpler version,
139: it represents a system of particles moving on a circle, all coupled by
140: an equal strength attractive or repulsive cosine interaction. An
141: extension of it to the case in which particles move on a 2D torus has
142: been introduced in Ref.~\cite{at} and it has been quite recently
143: realized that all such models are particular cases of a more general
144: Hamiltonian~\cite{art}.
145:
146: The HMF model, that we introduce in Section 2,
147: is exactly solvable in the canonical ensemble by a
148: Hubbard-Stratonovich transformation. The solution in the
149: microcanonical ensemble can be obtained only under certain hypotheses
150: that we will discuss in Section 3, but detailed information on the
151: behavior in the microcanonical ensemble can be obtained by direct
152: molecular dynamics (MD) simulations. The model has first and second
153: order phase transitions and tricritical points. Its rich phase diagram
154: allows to test the presence of ensemble inequivalence
155: \index{ensemble inequivalence} near canonical first order phase transitions and,
156: indeed, we find negative specific heat and temperature jumps in the
157: microcanonical ensemble. Having access to dynamics, one can study
158: metastability of out-of-equilibrium states. This is done in Section 4,
159: where we analyze the emergence of a coherent structure in the
160: repulsive HMF at low energy. Similar features are also discussed in
161: another chapter of this book~\cite{tsallisrap} for the attractive case
162: near the second order phase transition. Section 5 is devoted to the
163: study of the spectrum of Lyapunov exponents. The maximal exponent has
164: a peak near the phase transition~\cite{at,lat1hmf,celiahmf} and vanishes
165: when increasing the number of particles with a universal scaling law
166: in the whole high energy disordered phase. In a low energy range the
167: Lyapunov spectrum has a thermodynamic limit distribution similar to
168: the one observed for systems with short-range
169: interaction~\cite{livipolruffo}.
170:
171:
172: \section{The HMF models}
173:
174: A generic two-body potential in a two dimensional square box of
175: side $2\pi$ with periodic boundary conditions, a 2D torus, can be Fourier
176: expanded as
177: \begin{equation}
178: \label{fourierexpan}
179: V(x,y)=\sum_{{\bf k}=(k_x,k_y)} \exp \left( i {\bf k} \cdot {\bf r}
180: \right) V(k_x,k_y)\quad.
181: \end{equation}
182: A sufficiently rich family of potential functions is obtained if,
183: we restrict to the first two momentum shells $|k|=1$ and
184: $|k|=\sqrt{2}$, we require that the potential is only invariant
185: under discrete rotations by all multiples of $\pi/4$, and we assume
186: that the Fourier coefficients on each shell are the same.
187: This amounts to perform a truncation in the Fourier expansion
188: of the potential (\ref{fourierexpan}), as done in studies of
189: spherically symmetric gravitational systems in another chapter
190: of this book~\cite{grossdh}.
191: We get
192: \begin{equation}
193: \label{limitexpansion}
194: V(x,y)=a+b\cos x +b\cos y +c\cos x\cos y\;.
195: \end{equation}
196: As the constant $a$ is arbitrary and scaling $b$ is equivalent to
197: scale the energy, $c$ remains the only free parameter. We consider
198: $N$ particles interacting through the two-body potential $V(x,y)$ and
199: we adopt the Kac prescription~\cite{Kachmf}\index{Kac prescription}
200: \index{extensivity} to scale the equal strength coupling among the
201: particles by their number $N$. This scaling allows to perform safely
202: the thermodynamic limit\index{thermodynamic limit}, since both the
203: kinetic and the potential energy increase proportional to
204: $N$\footnote{Kac prescription is, however, unphysical and it would be
205: important to find a viable alternative.}. By appropriately
206: redefining the constants $a=2\varepsilon+A,b=-\varepsilon,c=-A$ in formula
207: (\ref{limitexpansion}) and using Kac prescription one gets the
208: following potential energy
209: \begin{eqnarray}
210: \label{eqHA}
211: V_A &=& \frac{1}{2N} \sum^{N}_{i,j=1}
212: \varepsilon \left( 1- \cos ( x_i-x_j ) \right)
213: +\varepsilon \left( 1-\cos ( y_i-y_j ) \right) \nonumber \\
214: && \hskip 2truecm+A \left( 1 - \cos ( x_i-x_j ) \cos ( y_i-y_j ) \right) ~,
215: \end{eqnarray}
216: with $(x_{i},y_{i})\in ]-\pi, \pi ] \times ]-\pi, \pi ]$ representing the coordinates
217: of $i$-th particle and $(p_{x,i},p_{y,i}) \) its conjugated momentum.
218: The Hamiltonian of the HMF model is now the sum of this potential
219: energy with the kinetic energy
220: \begin{equation}
221: \label{kinetic}
222: K= \sum _{i=1}^{N}\left(\frac{p_{x,i}^{2}+p_{y,i}^{2}}{2}\right)~.
223: \end{equation}
224: We get
225: \begin{equation}
226: \label{hamHMF}
227: H_{HMF}=K+V_A~.
228: \end{equation}
229: In the following, we will consider model (\ref{hamHMF}) for
230: $A=0$ and both $\varepsilon$ positive (attractive case) and
231: negative (repulsive case). The $A\neq0$ case will always have
232: $A>0$ and $\varepsilon=1$.
233:
234: \section{Equilibrium Thermodynamics}
235:
236: In this section, we discuss the equilibrium thermodynamical
237: results for model~(\ref{eqHA}) in the canonical and
238: microcanonical ensembles. Canonical results will be obtained
239: analytically while, for the microcanonical ones, we will mostly rely on
240: molecular dynamics (MD) simulations.
241:
242: \subsection{Canonical Ensemble for $A=0$}\index{canonical!ensemble}
243:
244: \label{Aeq0}
245: For pedagogical reasons, we will initially limit our
246: analysis to the case $A=0$, for which the model reduces to two
247: identical {\em uncoupled} systems: one describing the evolution of
248: the $\{ x_i,p_{x,i} \}$ variables and the other $\{y_i,p_{y,i} \}$.
249: Therefore let us rewrite the Hamiltonian
250: associated to one of these two sets of variables, named
251: $\theta_i$ in the following. We obtain
252: \begin{equation}
253: H_0 = \sum_{i=1}^N \frac{p_i^2}{2} + \frac{\varepsilon}{2N}
254: \sum_{i,j=1}^N [1-\cos(\theta_i - \theta_j)] = K_0 + V_0
255: \label{model0}
256: \end{equation}
257: where $\theta_i \in [-\pi;\pi[$ and $p_i$ are the corresponding
258: momenta. This model can be seen as representing particles moving
259: on the unit circle, or as classical $XY$-rotors with infinite
260: range couplings. For $\varepsilon>0$, particles attract each other
261: and rotors tend to align (ferromagnetic case), while for
262: $\varepsilon<0$, particles repel each other and spins tend to
263: anti-align (antiferromagnetic case). At short distances, we can
264: either think that particles cross each other or that they collide
265: elastically since they have the same mass.
266:
267: The physical meaning of this model is even clearer if one
268: introduces the mean field vector
269: \begin{equation}
270: {\bf M} = M {\rm e}^{i \phi} = \frac{1}{N} \sum_{i=1}^N {\bf m}_i
271: \label{m0}
272: \end{equation}
273: where ${\bf m}_i =(\cos \theta_i, \sin \theta_i)$. $M$ and $\phi$
274: represent the modulus and the phase of the order parameter, which
275: specifies the degree of clustering in the particle interpretation,
276: while it is the {\em magnetization} for the $XY$ rotors. Employing
277: this quantity, the potential energy can be rewritten as a sum of
278: single particle potentials $v_i$
279: \begin{equation}
280: V_0 = \frac{1}{2} \sum_{i=1}^N v_i \qquad {\rm with} \qquad v_i =
281: 1- M \cos(\theta_i -\phi) \quad . \label{v0}
282: \end{equation}
283: It should be noticed that the motion of each particle is coupled to all
284: the others, since the mean-field
285: variables $M$ and $\phi$ are determined at each time $t$ by the
286: instantaneous positions of all particles.
287:
288: The equilibrium results in the canonical ensemble can be obtained from
289: the evaluation of the partition function
290: \index{canonical!partition function}
291: \begin{equation}
292: Z = \int d^N p_i d^N \theta_i \exp{(-\beta H)}
293: \label{z0}
294: \end{equation}
295: where $\beta = 1/(k_B T)$, with $k_B$ the Boltzmann constant and
296: $T$ the temperature. The integration domain is extended
297: to the whole phase space. Integrating over momenta, one gets:
298: \begin{equation}
299: Z = \left( \frac{2 \pi}{\beta} \right)^{N/2} \int_{-\pi}^\pi d^N
300: \theta_i \exp{\left[\frac{-\beta \varepsilon N}{2} (1-{\bf
301: M}^2)\right]}\quad. \label{zz0}
302: \end{equation}
303: In order to evaluate this integral, we use the two dimensional
304: Gaussian identity
305: \begin{equation}
306: \exp{\left[\frac{\mu}{2} \bf x^2\right]}=
307: \frac{1}{\pi} \int_{-\infty}^\infty \int_{-\infty}^\infty
308: d{\bf y} \exp{[-{\bf y}^2+\sqrt{2 \mu} {\bf x}
309: \cdot {\bf y}]}
310: \label{gau}
311: \end{equation}
312: where ${\bf x}$ and ${\bf y}$ are two-dimensional vectors and
313: $\mu$ is positive. We can therefore rewrite Eq.~(\ref{zz0}) as
314: \begin{equation}
315: Z = \left( \frac{2 \pi}{\beta} \right)^{N/2}
316: \exp{\left[\frac{-\beta \varepsilon N}{2}\right]} J
317: \end{equation}
318: with
319: \begin{equation}
320: J= \frac{1}{\pi}
321: \int_{-\pi}^\pi d^N \theta_i
322: \int_{-\infty}^\infty \int_{-\infty}^\infty
323: d{\bf y} \exp{[-{\bf y}^2+\sqrt{2 \mu} {\bf M}
324: \cdot {\bf y}]}
325: \label{j0}
326: \end{equation}
327: and $\mu = \beta \varepsilon N$. We use now definition~(\ref{m0}) and
328: exchange the order of the integrals in (\ref{j0}), factorizing the
329: integration over the coordinates of the particles. Introducing
330: the rescaled variable $ {\bf y} \to {\bf y} \sqrt{N/2\beta
331: \varepsilon}$, one ends up with the following expression for $J$
332: \begin{equation}
333: J= \frac{N}{2 \pi \beta \varepsilon} \int_{-\infty}^\infty
334: \int_{-\infty}^\infty d{\bf y} \exp{\left[-N \left( \frac{y^2}{2
335: \beta \varepsilon} -\ln\left(2\pi I_0(y)\right) \right) \right] }
336: \label{j1}
337: \end{equation}
338: where $I_n$ is the modified Bessel function of order $n$ and $y$
339: is the modulus of ${\bf y}$. Finally, integral~(\ref{j1}) can be
340: evaluated by employing the saddle point technique in the
341: mean-field limit (i.e. for $N \to \infty$). In this limit, the
342: Helmholtz free energy
343: per particle $f$ reads as :
344: \begin{equation}
345: \beta f=-\lim_{N \to \infty} \frac{\ln Z}{N}=
346: -\frac{1}{2}\ln\left(\frac{2 \pi}{\beta}\right) +\frac{\varepsilon
347: \beta }{2} +\max_y\left(\frac{y^2}{2\beta\varepsilon}-\ln(2\pi
348: I_0(y))\right)\quad. \label{freehmf}
349: \end{equation}
350: The maximum condition leads to the consistency equation
351: \begin{equation}
352: \frac{y}{\beta \varepsilon}=\frac{I_1(y)}{I_0(y)}\quad.
353: \label{cons}
354: \end{equation}
355: For $\varepsilon<0$, there is a unique solution $\bar y=0$, which
356: means that the order parameter remains zero and there is no phase
357: transition (see Figs.~\ref{fe_anti}c,d). The particles are all the
358: time homogeneously distributed on the circle and the rotors have
359: zero magnetization. On the contrary, in the ferromagnetic case
360: ($\varepsilon>0$), the solution $\bar y=0$ is unstable for
361: $\beta \geq \beta_c = 2$. At $\beta=\beta_c$, two stable symmetric
362: solutions appear through a pitchfork bifurcation and a discontinuity
363: in the second derivative of the free energy is present, indicating a
364: second order phase transition\index{phase!transition}.
365:
366:
367: \begin{figure}[ht]
368: \begin{center}
369: \includegraphics[width=0.9\textwidth]{cc.eps}
370: \end{center}
371: \caption[]{Temperature and magnetization as a function of the
372: energy per particle $U$ for $A=0$ and $|\varepsilon|=1$ in the
373: ferromagnetic (a),(b) and in the antiferromagnetic case (c),(d).
374: Symbols refer to MD data for $N=10^2$ and $10^3$, while the solid
375: lines refer to the canonical prediction obtained analytically in
376: the mean-field limit. The vertical dashed line indicates the
377: critical energy in the ferromagnetic model, located at $U_c=0.75$,
378: $\beta_c=2$. The inset of panel
379: (d) shows the rescaled magnetization $M\times \sqrt{N}$.}
380: \label{fe_anti}
381: \end{figure}
382:
383:
384: These results are confirmed by an analysis of the order
385: parameter\footnote{This is obtained by adding to the Hamiltonian
386: an external field and taking the derivative of the free energy
387: with respect to this field, evaluated at zero field.}
388: \begin{equation}
389: M= \frac{I_1(\bar y)}{I_0(\bar y)} \quad. \label{mag0}
390: \end{equation}
391: For $\varepsilon >0$, the magnetization $M$ vanishes continuously
392: at $\beta_c$ (see Fig.~\ref{fe_anti}a,b), while it is always
393: identical to zero in the
394: antiferromagnetic case (see Fig.~\ref{fe_anti}c,d). Since $M$
395: measures the degree of clustering of the particles, we have for
396: $\varepsilon >0$ a transition from a clustered phase when
397: $\beta>\beta_c$ to a homogeneous phase when $\beta<\beta_c$. We
398: can obtain also the energy per particle
399: \begin{equation}
400: U= \frac{\partial(\beta f)}{\partial \beta}=
401: \frac{1}{2\beta}+\frac{\varepsilon}{2}\left(1-M^2\right)
402: \label{u0}
403: \end{equation}
404: which is reported for $|\varepsilon|=1$ in Fig. \ref{fe_anti}.
405: Panels (c) and (d) of Fig.~\ref{fe_anti} are limited to the range $U>0$
406: because in the antiferromagnetic model a non-homogeneous state, a
407: {\em bicluster}, can be generated for smaller energies. The
408: emergence of this state modifies all thermodynamical and dynamical
409: features as will be discussed in section~\ref{bicluster}.
410:
411:
412: The dynamics of each particle obeys the following
413: pendulum equation of motion
414: \begin{equation}
415: \ddot \theta_i= -M \sin(\theta_i-\phi) \quad , \label{pend}
416: \end{equation}
417: where $M$ and $\phi$ have a non trivial time dependence, related
418: to the motion of all the other particles in the system.
419: Equation~(\ref{pend}) has been very successfully used to describe
420: several features of particle motion, like for instance trapping
421: and untrapping mechanisms~\cite{Antoni}. There are also numerical
422: indications~\cite{Antoni} and preliminary theoretical
423: speculations~\cite{priv} that taking the mean-field limit before
424: the infinite time limit, the time-dependence would disappear and
425: the modulus and the phase of the magnetization become constant.
426: This implies, as we will discuss in Section 5, that chaotic
427: motion would disappear.
428: The inversion of these two limits is also discussed in the
429: contribution by Tsallis et al~\cite{tsallisrap}.
430:
431: \subsection{Canonical ensemble for $A\neq0$}\index{canonical!ensemble}
432:
433: As soon as $A > 0$, the evolution along the two spatial directions
434: is no more decoupled and the system cannot be described in terms
435: of a single order parameter. Throughout all this section
436: $\epsilon=1$. A complete description of the phase
437: diagram of the system requires now the introduction of two
438: distinct order parameters:
439: \begin{equation}
440: {\bf M}_z= \left(\frac{\sum_i \cos z_i}{N},\frac{\sum_i \sin z_i}{N}\right)=M_z
441: \exp{(i \phi_z)} \label{mz}
442: \end{equation}
443: where $z_i=x_i$ or $y_i$ and;
444: \begin{equation}
445: {\bf P}_{x \pm y}= \left(\frac{\sum_i \cos (x_i \pm y_i)}{N},
446: \frac{\sum_i \sin (x_i \pm y_i)}{N}\right)=P_z
447: \exp{(i \psi_z)}\quad. \label{pz}
448: \end{equation}
449: It can be shown that on average $M_x \simeq M_y \simeq M$
450: and $P_{x+y} \simeq P_{x-y} \simeq P$: therefore, we
451: are left with only two order parameters.
452:
453: Following the approach of section~\ref{Aeq0}, the canonical
454: equilibrium properties can be derived analytically in the
455: mean-field limit~\cite{at}. We obtain
456: \begin{equation}
457: \beta f=\frac{M^2+P^2}{\beta} -\ln
458: \left[\frac{G(M,P;A)}{\beta} \right]
459: \label{free2d}
460: \end{equation}
461: with
462: \begin{equation}
463: G= \int_0^{2\pi} ds \enskip I_0\left(M+\sqrt{2A} P \cos s\right)
464: \exp{(M\cos s)} \label{g2d}
465: \end{equation}
466: where $s$ is an integration variable.
467:
468: The energy per particle reads as
469: \begin{equation}
470: U = \frac{1}{\beta}+\frac{2+A-2 M^2-A P^2}{2}=K+V_A \quad .
471: \label{u2d}
472: \end{equation}
473: Depending on the value of the coupling constant $A$, the
474: single particle potential defined through
475: $V_A = \frac{1}{2} \sum_i v_i$ changes its shape,
476: inducing the different clustering phenomena described below.
477: For small values of $A$, $v_i$ exhibits a single minimum
478: per cell $(x,y)\in ([-\pi,\pi],[-\pi,\pi])$ (see Fig.~\ref{egg}a
479: for $A=1$), while for larger values of $A$ four minima can
480: coexist in a single cell (see Fig.~\ref{egg}b for $A=4$).
481: In the former case only one clustered equilibrium state can
482: exist at low temperatures, while in the latter case,
483: when all the four minima have the same depth a phase with
484: two clusters can emerge, as described in the following.
485:
486:
487: \begin{figure}[ht]
488: \begin{center}
489:
490:
491: \includegraphics[width=\textwidth]{contourpot2d.ps}
492: \end{center}
493: \caption[]{Single particle potential $v_i(x,y)$ for $A=1,\varepsilon=1$ (a) and
494: $A=4,\varepsilon=1$(b).}
495: \label{egg}
496: \end{figure}
497:
498: Since the system is ruled by two different order parameters $M$
499: and $P$, the phase diagram is more complicated than in the $A=0$
500: case and we observe two distinct clustered phases. In the very low
501: temperature regime, the system is in the clustered phase $
502: CP_{1}$: the particles have all the same location in a single
503: point-like cluster and $M\approx P \approx 1 $. In the very large
504: temperature range, the system is in a homogeneous phase ($HP$)
505: with particles uniformly distributed, $M\approx P=O(1/\sqrt{N})$.
506: For $A > A_2 \sim 3.5$, an intermediate two-clusters phase $CP_{2}
507: $ appears. In this phase, due to the symmetric location of the two
508: clusters in a cell, $M \sim O(1/\sqrt{N})$ while $P \sim O(1)$ \cite{art}.
509: We can gain good insights on the transitions by considering the
510: line $T_M$ (resp. $T_P$) where $M$ (resp. $P$) vanishes and the
511: phase $CP_1$ (resp. $CP_2$) looses its stability (see Fig.
512: \ref{pda} for more details).
513:
514: \begin{figure}[ht]
515: \begin{center}
516: \includegraphics[width=0.6\textwidth,angle=270]{figcri.eps}
517: \end{center}
518: \caption[]{Canonical phase diagram of model~(\ref{eqHA}) reporting the
519: transition temperatures versus the coupling parameter $A$. The
520: solid (resp. dashed) lines indicate the $T_M$ (resp. $T_P$) lines.
521: The dots the points where the nature of the transitions change. \protect\(
522: A_{1}\protect \), \protect\( A_{2}\protect \) and \protect\(
523: A_{3}\protect \) are the threshold coupling constants that
524: determine the transition scenario \protect\( I\to IV\protect \). }
525: \label{pda}
526: \end{figure}
527:
528:
529: We can therefore identify the following four different scenarios
530: depending on the value of $A$.
531:
532: \textbf{(I)} When $ 0 \leq A \leq A_1 = 2/5$, one observes a continuous
533: transition from the phase $CP_1$ to $HP$. The critical line is
534: located at $T_M=1/2$ ($U_M=3/2+A$) and a canonical tricritical point,
535: located at $A_1=2/5$, separates the 1st order from the second order phase
536: transition\index{phase!transition} regions\footnote{
537: The tricritical point has been identified by finding first the
538: value of ${\bar P}(M,A,\beta)$ that minimizes $f(M,P;A,\beta)$ and then
539: substituting the solution $\bar P$ in $f$. This reduces $f$ to a
540: function of $M$ only. Now, the standard
541: procedure described also for the BEG model \cite{BMRhmf} can be applied.
542: This consists in finding the value of $\beta$ and $A$ where both
543: the $M^2$ and the $M^4$ coefficient of the development of the free
544: energy in powers of $M$ vanishes
545: }
546:
547: \textbf{(II)} When \( A_{1}<A<A_{2}\approx 3.5 \), the transition
548: between \( CP_{1} \) and \( HP \) is first order with a finite
549: energy jump (latent heat). Inside this range of values of $A$ one
550: also finds microcanonical discontinuous transitions (temperature
551: jumps) as we will see in Section 3.2.
552:
553: \textbf{(III)} When \(A_{2}<A<A_{3}\approx 5.7 \), the third phase
554: begins to play a role and two successive transitions are observed:
555: first $CP_1$ disappears at $T_M$ via a first order transition that
556: gives rise to $CP_2$; then this two-clusters phase gives rise to the
557: $HP$ phase via a continuous transition. The critical line
558: associated to this transition is $T_P =A/4$ ($U_P =3A/4+1$).
559:
560: \textbf{(IV)} When \( A>A_{3}\), the transition connecting the
561: two clustered phases, $CP_1$ and $CP_2$, becomes second order.
562:
563:
564: \subsection{Microcanonical Ensemble}\index{microcanonical!ensemble}
565: \label{microens}
566:
567: As we have anticipated, our microcanonical results have been mostly
568: obtained via MD simulations, since we cannot estimate easily the
569: microcanonical entropy $S$ analytically for the HMF model\footnote{It
570: would be possible using large deviation
571: technique~\cite{barrethesis}.}. However, we show below how far we
572: can get, starting from the knowledge of the canonical free energy,
573: using Legendre transform or inverse Laplace transform techniques. The
574: following derivation is indeed valid in general, it does not refer to
575: any specific microscopic model.
576:
577: The relation that links the partition function $Z(\beta,N)$ to
578: the microcanonical phase-space density at energy $E=U \cdot N$
579: \begin{equation}
580: \omega(E,N) = \int d^N p_i d^N \theta_i \; \delta(E-H)
581: \label{w0}
582: \end{equation}
583: is given by
584: \begin{equation}
585: Z(\beta,N)=\int_0^\infty dE \;\omega(E,N) \; {\rm e}^{-\beta E}~,
586: \label{z0_w}
587: \end{equation}
588: where the lower limit of the integral ($E=0$) corresponds
589: to the energy of the ground state of the model.
590: Expression (\ref{z0_w}) can be readily rewritten as
591: \begin{equation}
592: Z(\beta,N)=N \int_0^\infty dU
593: \exp{\left[N(-\beta U + \frac{1}{N} \ln(\omega(E,N))
594: \right]}~,
595: \label{z0_w2}
596: \end{equation}
597: which is evaluated by employing the saddle-point technique in the
598: mean-field limit. Employing the definition of entropy per particle
599: in the thermodynamic limit\index{thermodynamic limit}
600: \begin{equation}
601: S(U)=\lim_{N\to \infty} \left[\frac{1}{N} \ln \omega(U,N)
602: \right]\quad, \label{e0}
603: \end{equation}
604: one can obtain the Legendre transform
605: that relates the free energy to the entropy:
606: \begin{equation}
607: -\beta F(\beta) = \max_U[-\beta U +S(U)] \qquad {\rm with} \qquad
608: \beta=\frac{\partial S}{\partial U} \label{f0_e0}\quad.
609: \end{equation}
610: Since a direct analytical evaluation of the entropy of the HMF model in the
611: microcanonical ensemble is not possible, we are rather interested
612: in obtaining the entropy from the free energy. This can be done
613: only if the entropy $S$ is a concave function of the energy.
614: Then, one can invert (\ref{f0_e0}), getting
615: \begin{equation}
616: S(U)=\min_{\beta>0} [\beta(U - F(\beta))] \qquad {\rm with} \qquad
617: U=\frac{\partial (\beta F)}{\partial \beta} \quad . \label{e0_f0}
618: \end{equation}
619: However, the assumption that $S$ is concave is not true for systems
620: with long range interactions near a canonical first order
621: transition, where a "convex intruder"\index{convex intruder} of $S$
622: appears~\cite{BMRhmf,grossdh}, which gives rise to a negative specific
623: heat\index{negative!specific heat} regime in the microcanonical
624: ensemble\footnote{A convex intruder is present also for short range
625: interactions in {\em finite} systems, but the entropy regains its
626: concave character in the thermodynamic limit~\cite{grossdh}.}.
627: For such cases, we have to rely on MD simulations or microcanonical
628: Monte-Carlo simulations~\cite{grossbook}.
629:
630: An alternative approach to the calculation of the entropy consists in
631: expressing the Dirac $\delta$ function in Eq.~(\ref{w0})
632: by a Laplace transform. One obtains
633: \begin{equation}\label{invLapla}
634: \omega(E,N) =\frac{1}{2i\pi}\int_{-i\infty}^{+i\infty} d\beta\;
635: e^{\beta E}\; Z(\beta)
636: \end{equation}
637: where one notices that $\beta$ is imaginary. As the partition function can
638: be estimated for our model, we would just have to analytically
639: continue to complex values of~$\beta$. By performing a rotation to the
640: real axis of the integration contour, one could then evaluate $\omega$ by
641: saddle point techniques. However, this rotation requires the
642: assumption that no singularity is present out of the real axis: this
643: is not in general guaranteed. It has been checked numerically for
644: the $A=0$ model~\cite{barrethesis}, allowing to obtain then an explicit
645: expression for $\omega$ and confirming ensemble equivalence for
646: this case.
647:
648: However, when $A> A_1$, a first order canonical transition occurs,
649: which implies the presence of a convex intruder and makes the
650: evaluation of $S(U)$ through Eq.~\ref{e0_f0} impossible
651: \index{ensemble inequivalence}. For these cases, a canonical
652: description is unable
653: to capture all the features associated with the phase transition. For
654: the microcanonical entropy, we have to rely here on MD simulations.
655:
656: The MD simulations have been performed adopting extremely
657: accurate symplectic integration schemes~\cite{mac}, with
658: relative energy conservations during the runs of order $\sim
659: 10^{-6}$. It is important to mention that the CPU time required
660: by our integration schemes, due to the mean-field nature of the
661: model, increases linearly with the number of particles.
662:
663:
664: Whenever we observe canonically continuous transitions (i.e. for
665: $A \leq A_1$), the MD results coincide with those obtained
666: analytically in the canonical ensemble, as shown in
667: Fig.~\ref{fe_anti} for $A=0$. The curve $T(U)$ is thus well
668: reproduced from MD data, apart from finite $N$ effects. It has
669: been however observed that starting from "water-bags" initial
670: conditions metastable states can occur in the proximity of the
671: transition~\cite{tsallisrap}.
672:
673:
674: \begin{figure}[ht]
675: \begin{center}
676: \includegraphics[width=0.7\textwidth,angle=270]{UvT.eps}
677: \end{center}
678: \caption[]{Temperature-energy relation in the coexistence region
679: for A=1 (a) and A=4 (b). Lines indicate canonical analytical
680: results, while circles correspond to microcanonical MD
681: simulations. Solid thick lines are equilibrium results, solid thin
682: lines metastable states and dashed thin lines unstable states. The
683: dash-dotted line is the Maxwell construction. Figure (a)
684: refers to a first order transition from $CP_1$ to $HP$, (b) to
685: discontinuous transition connecting the two clustered phases. In
686: (b) the second order transition from $CP_2$ to $HP$ associated to
687: the vanishing of $P$ is also shown. The MD results refer to model
688: (\ref{eqHA}) with $N=5000$ averaged over a time $t=10^6$.} \label{UvT}
689: \end{figure}
690:
691: When $A > A_1$, discrepancies between the results obtained in the two
692: ensembles are observable in Fig.~\ref{UvT}. For example, MD
693: results in the case $A=1$, reported in Fig.~\ref{UvT}(a), differ
694: clearly from canonical ones around the transition, exhibiting a regime
695: characterized by a negative specific heat\index{negative!specific heat}.
696: This feature is common to many models with
697: long-range~\cite{thir,compa} or power-law decaying
698: interactions~\cite{tamahmf} as well as to finite systems with
699: short-range forces~\cite{grossbook}. However, only recently a
700: characterization of all possible microcanonical transitions associated
701: to canonically first order ones has been initiated~\cite{BMRhmf,art}.
702:
703: For $A$ slightly above $A_1$, the transition is microcanonically
704: continuous, i.e. there is no discontinuity in the $T-U$ relation
705: (this regime presumably extends up to $A\sim1.2$). Before the
706: transition, one observes a negative specific heat regime (see
707: Fig.~\ref{UvT}a). In addition, as already observed for the
708: Blume-Emery-Griffiths model~\cite{BMRhmf}, microcanonically {\em
709: discontinuous} transitions can be observed in the "convex
710: intruder" region. This means that, at the transition energy, {\em
711: temperature jumps} exist in the thermodynamic limit. A complete
712: physical understanding of this phenomenon, which has also been
713: found in gravitational systems~\cite{chavanishmf}, has not been
714: reached. For $A> 1.2$, i.e. above the "microcanonical tricritical
715: point", our model displays temperature jumps. This situations is
716: shown in Fig. \ref{mdis} for $A=2$.
717:
718: \begin{figure}[ht]
719: \begin{center}
720: \includegraphics[width=0.5\textwidth,angle=270]{crit.eps}
721: \end{center}
722: \caption[]{Evolution of the temperature $T$ versus the energy $U$
723: in the case $A=2$. The symbols refer to MD results obtained by
724: successively cooling or warming a certain initial configuration.
725: Each simulation has been performed at constant total
726: energy and refers to a system of $N=4,000$ particles
727: integrated for a time $t=10^6$. The solid lines are computed
728: in the canonical ensemble and include also unstable and metastable
729: cases. The solid vertical line
730: indicating the transition energy has been estimated via
731: a Maxwell construction performed in the microcanonical ensemble
732: (for details see \protect\cite{BMRhmf}).}
733: \label{mdis}
734: \end{figure}
735:
736:
737: For $A>A_2$, we have again a continuous transition connecting the
738: two clustered phases $CP_1$ and $CP_2$. This is the first angular
739: point in the $T-U$ relation at $U\sim3.65$ in Fig.~\ref{UvT}(b).
740: The second angular point at $U\sim 4$ is the continuous
741: transition, connecting $CP_2$ to $HP$. The transition at lower
742: energy associated to the vanishing of $M$ is continuous in the
743: microcanonical ensemble with a negative specific heat, while
744: discontinuous in the canonical ensemble;
745: the dash-dotted line indicates the transition temperature in
746: the canonical ensemble,
747: derived using the Maxwell construction). The second transition
748: associated to the vanishing of $P$ is continuous in both
749: ensembles.
750:
751:
752:
753: \section{Dynamical Properties I: Out-of-equilibrium states}
754: \index{non!equilibrium phenomena}
755:
756: \subsection{Metastable states}\index{metastability}
757: Around the critical energy, relaxation to equilibrium depends in a
758: very sensitive way on the initial conditions adopted. When one starts
759: with out-of-equilibrium initial conditions in the ferromagnetic case,
760: one finds quasi-stationary (i.e. long lived) nonequilibrium
761: states\index{coherent!structures}. An example is represented by the
762: so-called ``water bag'' initial condition: all the particles are clustered
763: in a single point and the momenta are distributed according to a flat
764: distribution of finite width centered around zero. These states have
765: a lifetime which increases with the number of particles~$N$, and are
766: therefore stationary in the continuum limit. In correspondence of
767: these metastable states, anomalous diffusion and L{\'e}vy walks
768: \cite{levyhmf}, long living correlations in $\mu-$space \cite{lat0} and
769: zero Lyapunov exponents \cite{lat0a} have been found. In addition,
770: these states are far from the equilibrium caloric curve around the
771: critical energy, showing a region of negative specific heat and a
772: continuation of the high temperature phase (linear $T$ vs $U$
773: relation) into the low temperature one. It is very intriguing that
774: these out-of-equilibrium quasi-stationary states indicate a caloric
775: curve very similar to the one found in the region where one gets a
776: canonical first order phase transitions, but a continuous
777: microcanonical one, as discussed in section~\ref{microens}. In the
778: latter case, however, the corresponding states are stationary also at
779: finite $N$. The coexistence of different states in the continuum limit
780: near the critical region is a purely microcanonical effect, and arises
781: after the inversion of the $t \to \infty$ limit with the $N \to \infty$ one
782: \cite{lat0,lat0a,tsallisrap}.
783:
784: Similarly, the antiferromagnetic HMF ($A=0$, $\varepsilon =-1$),
785: where the particles interact through {\em repulsive} forces presents
786: unexpected dynamical properties in the
787: out-of-equili\-brium thermodynamics. On the first sight, the
788: thermodynamics of this model seems to be less interesting since no phase
789: transition occurs as discussed above. However, thermodynamical
790: predictions are again in some cases in complete disagreement with
791: dynamical results leading in particular to a striking localization
792: of energy. This aspect, as we will show below, is of course,
793: again, closely related to the long-range character of the
794: interaction and to the fact that such a dynamics is chaotic and
795: self-consistent. We mean by this that all particles give a contribution to the field
796: acting on each of them. One calls this phenomenon,
797: {\em self-consistent chaos}~\cite{Diego}. In addition to the toy model that we
798: consider here, we do think that similar emergence of
799: structures, but even more importantly, similar dynamical
800: stabilization of out-equilibrium states could be encountered in
801: other long-range systems, as we briefly describe at the end of this
802: section.
803:
804:
805: \subsection{The dynamical emergence of the bicluster in the antiferromagnetic case}
806: \label{bicluster}
807:
808: In the antiferromagnetic case, Eq.~(\ref{model0}) with $\varepsilon =-1$, the
809: intriguing properties appear in the region of very small energies. To
810: be more specific, if an initial state with particles evenly
811: distributed on the circle (i.e. close to the ground state predicted
812: by microcanonical or canonical thermodynamics) and with vanishingly
813: small momenta is prepared, this initial condition can lead to the
814: formation of unstable states. This process, discovered by chance, is
815: now characterized in full
816: detail~\cite{Antoni,barreepjB,drh,firpoleyvraz}.
817:
818: As shown by Fig.~\ref{instab}, the density of particles is initially
819: homogeneous. However a localization of particles do appear at a given
820: time, in two different points, symmetrically located with respect to
821: the center of the circle. This localized state, that we call
822: {\it bicluster}\index{coherent!structures}, is however unstable (as shown
823: again by Fig.~\ref{instab}), since both clusters are giving rise to
824: two smaller localized groups of particles: this is the reason for the
825: appearance of the first {\em chevron}. However, also this state is
826: unstable, so that the first chevron disappears to give rise again to a
827: localization of energy in two points. This state enhances the
828: formation of a chevron with a smaller width and this phenomenon
829: repeats until the width of the chevron is so small that one does not
830: distinguish anymore its destabilization.
831: Asymptotically one gets a density distribution displaying two sharp
832: peaks located at distance $\pi$ on the circle, a
833: dynamically stable bicluster.
834:
835: \begin{figure}[ht]
836: \begin{center}
837: \includegraphics[width=0.7\textwidth]{instab.ps}
838: \end{center}
839: \vskip-1truecm
840: \caption[]{Short-time evolution of the particle density
841: in grey scale: the darker the grey, the higher the density.
842: Starting from an initial condition with all the particles evenly
843: distributed on the circle, one observes a very rapid concentration
844: of particles, followed by the quasi periodic appearance of
845: {\em chevrons}, that shrink as time increases.} \label{instab}
846: \end{figure}
847:
848:
849: As we have shown in Ref.~\cite{barreepjB}, the emergence of the
850: bicluster is the signature of shock waves present in the
851: associated hydrodynamical equations. Indeed, we found a strikingly
852: good description of the dynamics of the particles by a non linear
853: analysis of the associated Vlasov equation, which is mathematically
854: justified~\cite{BraunHepp} in the infinite $N$ limit.
855: The physical explanation of this problem can
856: be summarized as follows. Once the Hamiltonian has been mapped to
857: the Vlasov equation, it is possible to introduce a density $\rho(\theta,t)$ and a
858: velocity field $v(\theta,t)$. Neglecting the dispersion in
859: momentum and relying on usual non linear dynamics hierarchy of
860: time-scales, one ends up with dynamical equations at different
861: orders in a multiscale analysis. The
862: first order corresponds to the linear dynamics and defines the
863: plasma frequency of order one. However, a second timescale appears
864: that is related to the previous one by the relationship
865: $\tau=\sqrt{U}\, t$, where $U$ is the energy per particle. When one
866: considers initial conditions with a very small energy density,
867: the two time scales are very different and clearly
868: distinguishable by considering particle trajectories: a typical trajectory
869: corresponds indeed to a very fast motion with a very small
870: amplitude, superimposed to a slow motion with a large amplitude.
871:
872: This suggests to average over the fast oscillations and leads to the
873: spatially forced Burgers equation
874: \begin{eqnarray}
875: \label{burgers} \frac{\partial u}{\partial \tau} +u\frac{\partial
876: u} {\partial \theta} & = & -\frac{1}{2}\sin 2\theta \qquad ,
877: \end{eqnarray}
878: once the average velocity $u(\theta,\tau)=\langle
879: v(\theta,t,\tau)\rangle_t$ is introduced. Due to the absence of dissipative
880: or diffusive terms, Equation~(\ref{burgers}) supports shock
881: waves and these can be related to the emergence of the bicluster.
882: By applying the methods of characteristics to solve
883: Equation~(\ref{burgers}), one obtains
884: \begin{eqnarray}
885: \frac{d^2\theta}{d\tau^2} + \frac{1}{2} \sin 2\theta & = & 0\;.
886: \label{burgerslag}
887: \end{eqnarray}
888: which is a pendulum-like equation.
889:
890: Fig.~\ref{schock} shows the trajectories, derived from
891: Eq.~(\ref{burgerslag}), for particles that are initially evenly
892: distributed on the circle. One clearly sees that two shock waves appear and
893: lead to an increase of the number of particles around two
894: particular sites, which depend on the initial conditions: this two
895: sites correspond to the nucleation sites of the bicluster. Because
896: of the absence of a diffusive term, the shock wave starts a spiral
897: motion that explains the destabilization of the first bicluster
898: and also the existence of the two arms per cluster, i.e. the
899: chevron.
900:
901: \begin{figure}[ht]
902: \begin{center}
903: \includegraphics[width=0.6\textwidth]{doubleshoc.ps}
904: \end{center}
905: \caption[]{Five successive snapshots of the velocity
906: profiles $u(\theta,t)$ are shown including the initial state when all
907: particles are uniformly
908: distributed in space with a small velocity dispersion.}
909: \label{schock}
910: \end{figure}
911:
912: This dynamical analysis allows an even more precise description,
913: since the methods of characteristics show that the
914: trajectories correspond to the motion of particles in the
915: double well periodic potential $V(\theta,\tau)=-(\cos 2\theta)/4$. This
916: potential is of mean field origin, since it represents the effect
917: of all interacting particles. Therefore, the particles
918: will have an oscillatory motion in one of the two wells.
919: One understands thus that particles starting close to
920: the minimum will collapse at the same time, whereas a particle
921: starting farther will have a larger oscillation period. This is
922: what is shown in Fig.~\ref{chevronsburgers} where
923: trajectories are presented for different starting positions. One
924: sees that the period of recurrence of the chevrons corresponds to
925: half the period of oscillations of the particles close to the minimum;
926: this fact is a direct consequence of the isochronism of the
927: approximate harmonic potential close to the minimum. The
928: description could even go one step further by computing the
929: caustics, corresponding to the envelops of the characteristics.
930: They are shown in Fig.~\ref{chevronsburgers} and testify the
931: striking agreement between this description and the real
932: trajectories: the chevrons of Fig.~\ref{instab} correspond to the
933: caustics.
934:
935: \begin{figure}[ht]
936: \begin{center}
937: \includegraphics[width=0.6\textwidth]{chevronsburgers.ps}
938: \end{center}
939: \caption[]{Superposition of the caustics (thick full lines) over the
940: characteristics (dotted lines) of
941: particles that are initially evenly distributed
942: between $-\pi/2$ and $\pi/2$.}
943: \label{chevronsburgers}
944: \end{figure}
945:
946:
947: \subsection{Thermodynamical predictions versus dynamical stabilization}
948: \label{thermo}
949:
950: If the above description is shown to be particularly accurate, it
951: does not explain why this state is thermodynamically preferred
952: over others. Indeed,
953: as shown before, thermodynamics predicts that the only
954: equilibrium state is homogeneous. This result has
955: been discussed in Section 3.1, where we have proved
956: that magnetization is zero at all energies for the antiferromagnetic
957: model and this has been also confirmed by Monte Carlo
958: simulations~\cite{drh}. However, since MD simulations
959: are performed at constant energy, it is important to derive
960: analytically the most probable state in the microcanonical
961: ensemble. Since this model does not present ensemble inequivalence,
962: we can obtain the microcanonical results by employing the inverse
963: Laplace transform (\ref{invLapla}) of the canonical partition function.
964: The microcanonical solution confirms that the maximal entropy state is
965: homogeneous on the circle.
966: It is therefore essential to see why the
967: bicluster state, predicted to be thermodynamically unstable is
968: instead {\it dynamically stable}.
969:
970: The underlying reason rely on the existence of the two very
971: different timesca\-les and the idea is again to average over the
972: very fast one. Instead of using the classical asymptotic expansion
973: on the equation of motions, it is much more appropriate to develop
974: an adiabatic approximation which leads to an effective Hamiltonian
975: that describes very well the long time dynamics. Doing statistical mechanics
976: of this averaged problem, one predicts the presence of the bicluter.
977:
978: The theory that we have developed relies on an application
979: of adiabatic theory, which in the case of the
980: HMF model is rather elaborate and needs lengthy
981: calculations~\cite{barreepjB} that we will not present here.
982: An alternative, but less powerful, method to derive similar
983: results has been used in Ref.~\cite{firpoleyvraz}.
984: On the contrary, we would like to present a qualitative explanations
985: of this phenomenon, using a nice (but even too simple !) analogy.
986:
987: This stabilization of unstable states can be described using the
988: analogy with the inverted pendulum, where the vertical unstable
989: equilibrium position can be made stable by the application of
990: a small oscillating force. One considers a rigid rod
991: free to rotate in a vertical plane and whose point-of-support is
992: vibrated vertically as shown by
993: Fig.~\ref{excitationparametrique}a. If the support oscillates
994: vertically above a certain frequency, one discovers the remarkable
995: property that the vertical position with the center of mass above
996: its support point is {\em stable}
997: (Fig.~\ref{excitationparametrique}b). This problem, discussed
998: initially by Kapitza~\cite{Kapitaza}, has strong similarities with
999: the present problem and allows a very simplified presentation of
1000: the averaging technique we have used.
1001: \begin{figure}[ht]
1002: \begin{center}
1003: \includegraphics[width=0.45\textwidth]{excitationparametrique.ps}
1004: \includegraphics[width=0.45\textwidth]{excitationparametriquebis.ps}
1005: \end{center}
1006: \vskip-1truecm
1007: \caption[]{Schematic picture of the inverted pendulum.} \label{excitationparametrique}
1008: \end{figure}
1009:
1010: The equation of motion of the vibrating pendulum is
1011: \begin{equation}\label{eqpendukuminverted}
1012: \frac{d^2 \theta}{dt^2}+\left(\frac{\omega_0^2}{\omega^2}-a\cos t\right)\sin
1013: \theta=0~,
1014: \end{equation}
1015: where $\omega_0$ is the proper linear frequency of the pendulum, $\omega$ the
1016: driving frequency of the support and $a$ the amplitude of
1017: excitation. Introducing a small parameter
1018: $\varepsilon=\omega_0/\omega$, one sees
1019: that~(\ref{eqpendukuminverted}) derives from the Lagrangian
1020: \begin{equation}\label{Lagrangianpendulum}
1021: {\cal L}=\frac{1}{2}\left(\frac{d \theta}{d t}\right)^2+\left(\varepsilon^2-a\cos
1022: t\right)\cos\theta\quad.
1023: \end{equation}
1024: Here the two frequencies $\omega$ and $\omega_0$ define two
1025: different time scales, in close analogy with the HMF model. Using
1026: the small parameter to renormalize the amplitude of the excitation as
1027: $A=a/\varepsilon$, and choosing the ansatz
1028: $\theta=\theta_0(\tau)+\varepsilon\,\theta_1(t,\varepsilon)$ where
1029: $\tau=\varepsilon t $, the Lagrangian equations for the function
1030: $\theta_1$, leads to the solution $\theta_1=-A\cos
1031: t\sin\theta_0(\tau)$. This result not only simplifies the above
1032: ansatz, but more importantly suggest to average the Lagrangian on
1033: the fast variable $t$ to obtain an effective Lagrangian $ {\cal
1034: L}_{eff}=\langle {\cal
1035: L}\rangle_t=\frac{1}{2}\left(\frac{d\theta_0}{d\tau}\right)^2+V_{eff}$~,
1036: where the averaged potential is found to be
1037: \begin{equation}\label{poteff}
1038: V_{eff}=-\cos\theta_0-\frac{A^2}{8}\cos2\theta_0+\mbox{Cste}\quad.
1039: \end{equation}
1040: It is now straightforward to show that the inverted position would
1041: be stable if $A^2>2$, i.e. if $a\omega>\sqrt{2}\omega_0$. As the
1042: excitation amplitude $a$ is usually small, this condition
1043: emphasizes that the two time scales should be clearly different,
1044: for the inverted position to be stable.
1045:
1046: The procedure for the HMF model is analogous, but of
1047: course it implies a series of tedious calculations.
1048: Since we would like to limit here to a pedagogical presentation,
1049: we will skip such details that can be found in
1050: Ref.~\cite{barreepjB}.
1051: It is however important to emphasize
1052: that the potential energy in the HMF model is self-consistently
1053: determined and depend on the position of all particles. The magic
1054: and the beauty is that, even if this is the potential energy of
1055: $N$ particles, it is possible to compute the statistical mechanics
1056: of the new effective Hamiltonian, derived directly from the
1057: effective Lagrangian via the Legendre transform. The main result
1058: is that the out-of-equilibrium state, (i.e. the bicluster
1059: shown in Fig.~\ref{instab}) corresponds to a statistical
1060: equilibrium of the effective mean-field dynamics.
1061: No external drive is present in this case, as for the inverted pendulum, but
1062: the time dependence of the mean field plays the role of the
1063: external drive.
1064:
1065: The HMF model represents presumably the simplest $N$-body
1066: system where out-of-equilibrium dynamically stabilized states
1067: can be observed and explained in detail.
1068: However, we believe that several systems with long range interactions
1069: should exhibit behaviours similar to the ones we have observed here.
1070: Moreover, this model represents a paradigmatic example for
1071: other systems exhibiting nonlinear interactions of rapid oscillations and
1072: a slower global motion. One of this is the piston problem~\cite{piston}:
1073: averaging techniques have been applied to the fast motion of gas particles in a
1074: piston which itself has a slow motion~\cite{sinairusse}. Examples
1075: can also be found in applied physics as for instance
1076: wave-particles interaction in plasma physics~\cite{plasma}, or the
1077: interaction of fast inertia gravity waves with the vortical motion
1078: for the rotating Shallow Water model~\cite{embidmajda}.
1079:
1080:
1081: \section{Dynamical Properties II: Lyapunov exponents}
1082:
1083: In this section, we discuss the chaotic features of the microscopic
1084: dynamics of the HMF model. We mainly concentrate on the $A=0$ case,
1085: presenting in details the behavior of the Lyapunov exponents and the
1086: Kolmogorov-Sinai entropy both for ferromagnetic and antiferromagnetic
1087: interactions. We also briefly discuss some peculiar mechanisms of chaos in the $A
1088: \neq 0$ case. The original motivation for the study of the chaotic
1089: properties of the HMF was to investigate the relation between phase
1090: transitions, which are macroscopic phenomena, and microscopic
1091: dynamics (see Ref.~\cite{laporeport} for a review) with the purpose
1092: of finding dynamical
1093: signatures of phase transitions\index{phase!transition}
1094: \cite{cmd}. Moreover, we wanted to check the scaling properties with
1095: the number of particles of the Lyapunov spectrum~\cite{livipolruffo} in
1096: the presence of long-range interactions.
1097:
1098:
1099: \subsection{The $A=0$ case}
1100:
1101: In the $A=0$ case (see formulae (\ref{model0})-(\ref{pend})), the Hamiltonian
1102: equations of motion are
1103: \begin{eqnarray}
1104: \dot \theta_i&=&p_i\\
1105: \dot p_i &=&-M \sin\left(\theta_i - \phi\right)\quad.\label{eqmoto}
1106: \end{eqnarray}
1107: The Largest Lyapunov Exponent (LLE) is defined as the limit
1108: \begin{equation}
1109: {\lambda_1} = \lim_{t\to \infty} {1\over t} \ln { d(t)\over
1110: d(0) }
1111: \end{equation}
1112: where $d(t) = \sqrt{ \sum_{i=1}^N (\delta \theta_i)^2 + (\delta
1113: p_i)^2 } $ is the Euclidean norm of the infinitesimal disturbance
1114: at time $t$. Therefore, in order to obtain the time evolution of
1115: $d(t)$, one has to integrate also the linearized equations of
1116: motion along the reference orbit
1117: \begin{equation}
1118: \label{lin1} {d\over dt} \delta \theta_i = \delta p_i ~~~~,~~~~
1119: {d\over dt} \delta p_i ~= -\sum_j {{\partial^2V}\over {\partial
1120: \theta_i \partial \theta_j} } \delta q_j~~,
1121: \end{equation}
1122: \noindent where the diagonal and off-diagonal terms of the Hessian are
1123: \begin{eqnarray}
1124: \label{diag} {\partial^2V\over{\partial \theta_i^2}} &=&M
1125: \cos(\theta_i-\Phi) - \frac{1}{N}\\
1126: \label{der2}
1127: {{\partial^2V}\over {\partial \theta_i \partial \theta_j} } &=&
1128: -{1\over{N}}\cos(\theta_i-\theta_j)\quad, ~~~i \neq j ~~.
1129: \end{eqnarray}
1130: To calculate the largest Lyapunov exponent we have used the
1131: standard method by Benettin et al~\cite{ben}. In Fig.~(\ref{llehmf}),
1132: we report the results obtained for four different sizes of the
1133: system (ranging from $N=100$ to $N=20,000$).
1134:
1135: In panel (a), we plot the largest Lyapunov exponent as a function
1136: of U. As expected, $\lambda_1$ vanishes in the limit of very small
1137: and very large energies, where the system is quasi-integrable.
1138: Indeed, the Hamiltonian reduces to weakly coupled harmonic
1139: oscillators in the former case or to free rotators in the latter.
1140: For $U < 0.2$, $\lambda_1$ is small and has no $N$-dependence.
1141: Then it changes abruptly and a region of ``strong chaos'' begins.
1142: It was observed~\cite{Antoni} that between $U=0.2$ and $U=0.3$, a
1143: different dynamical regime sets in and particles start to
1144: evaporate from the main cluster, in analogy with what was
1145: reported in other models~\cite{cmd,nayak}. In the region of strong
1146: chaoticity, we observe a pronounced peak already for
1147: $N=100$~\cite{yama}. The peak persists and becomes broader for
1148: $N=20,000$. The location of the peak is slightly below the
1149: critical energy and depends weakly on~$N$.
1150: %
1151: \begin{figure}[ht]
1152: \begin{center}
1153: \includegraphics[width=.7\textwidth]{lle_f.eps}
1154: \end{center}
1155: \caption[] {Largest Lyapunov exponent LLE and kinetic energy
1156: fluctuations $\Sigma= \sigma_K / \sqrt{N}$ as a function of U
1157: in the $A=0$ ferromagnetic case for different N sizes. The theoretical curve
1158: is shown as a full line, see text.}
1159: \label{llehmf}
1160: \end{figure}
1161:
1162: In panel (b), we report the standard deviation of the kinetic
1163: energy per particle $\Sigma$ computed from
1164: \begin{equation}
1165: \Sigma= \frac{\sigma_{K}}{\sqrt{N}}=
1166: {{\sqrt {\langle K^2\rangle -\langle K\rangle^2}} \over \sqrt{N}}\quad, \label{snum}
1167: \end{equation}
1168: where $\langle \bullet\rangle $ indicates the time average. The
1169: theoretical prediction for $\Sigma$, which is also reported
1170: in Fig.\ref{llehmf}, is~\cite{firpo,lat2hmf}
1171: \begin{equation}
1172: \Sigma= {T\over \sqrt{2}} \sqrt{ 1 -{ \left[1-2 M \left({ dM\over
1173: dT }\right) \right]}^{-1} }\quad, \label{steo}
1174: \end{equation}
1175: where $M(T)$ is computed in the canonical ensemble. Finite size
1176: effects are also
1177: present for the kinetic energy fluctuations, especially for
1178: $U>U_c$, but in general there is a good agreement with the
1179: theoretical formula, although the experimental points in
1180: Fig.~\ref{llehmf}b lies systematically below it. The figure
1181: emphasizes that the
1182: behavior of the Lyapunov exponent is strikingly correlated with
1183: $\Sigma$: in correspondence to the peak in the LLE, we observe
1184: also a sharp maximum of the kinetic energy fluctuations. The
1185: relation between the chaotic properties and the thermodynamics of
1186: the system, namely the presence of a critical point, can be made even more
1187: quantitative. An analytical formula, relating (in the $A=0$ model)
1188: the LLE to the second order phase transition undergone by the
1189: system, has been obtained~\cite{firpo} by means of the
1190: geometrical approach developed in Refs.~\cite{lapo1,laporeport}.
1191: Using a reformulation of Hamiltonian dynamics in the language of
1192: Riemannian geometry, they have found a general analytical
1193: expression for the LLE of a Hamiltonian many-body system in terms
1194: of two quantities: the average $\Omega_0$ and the variance
1195: $\sigma_{\Omega}$ of the Ricci curvature $ \kappa_R = \Delta V/N =
1196: 1/N~ \sum_{i=1}^N {\partial^2V\over{\partial q_i^2}} $, where $V$
1197: is the potential energy and $q_i$ are the coordinates of the
1198: system. Since in the particular case of the HMF model, we have
1199: \begin{equation}
1200: {1\over N} \sum_{i=1}^N {\partial^2V\over{\partial \theta_i^2}}
1201: = M^2 -{1\over N} = {2K\over N} + 1 - 2U - {1\over N} ~~,
1202: \end{equation}
1203: the two quantities
1204: $\Omega_0$ and $\sigma_{\Omega}$ can be expressed in terms of
1205: average values and fluctuations either of $M^2$ or of the kinetic
1206: energy $K$
1207: \begin{eqnarray}
1208: \label{omegazerohmf} \Omega_0 &=& \langle M^2\rangle -
1209: \frac{1}{N}
1210: = \frac{2}{N}\langle K\rangle + (1-2U) -\frac{1}{N}
1211: \nonumber\\
1212: \sigma^2_{\Omega} &=& N {\sigma^2_{M^2}}
1213: = \frac{4}{N} {\sigma^2_{K}~~.}
1214: \end{eqnarray}
1215: The formula obtained by Firpo~\cite{firpo} relates
1216: the LLE, a characteristic dynamical quantity, to thermodynamical quantities
1217: like $\langle M^2\rangle $ and ${\sigma_{M^2}}$,
1218: or $\langle K\rangle $ and ${\sigma_{K}}$, which
1219: characterize the macroscopic phase transition.
1220: For moderately small values of $U$, an approximation of the formula
1221: gives
1222: \begin{equation}
1223: \label{formulaprl}
1224: \lambda
1225: %\approx 2 \frac{1}{M} \frac{\sigma_{K}}{\sqrt{N}}
1226: %= 2 \frac{1}{M} \Sigma
1227: \propto \frac{\sigma_{K}}{\sqrt{N}}=\Sigma~~.
1228: \end{equation}
1229: This is in agreement with the proportionality between LLE and
1230: fluctuations of the kinetic energy found numerically in
1231: Fig.~(\ref{llehmf}). This implies also a connection between the LLE
1232: and the specific heat, another quantity which is directly related
1233: to the kinetic energy fluctuation. In fact the specific heat can
1234: be obtained from $\Sigma$ by means of the Lebowitz-Percus-Verlet
1235: formula~\cite{lpv}
1236: \begin{equation}
1237: C_V = { 1\over 2} \left[ 1- 2 \left({ \Sigma \over T} \right)^2
1238: \right] ^{ -1} \quad. \label{cv1}
1239: \end{equation}
1240: In Fig.~\ref{cvhmf}, we report the numerical results for
1241: the specific heat as a function of $U$ for a system made of $N=500$
1242: particles, and we compare them with the theoretical estimate.
1243:
1244:
1245: \begin{figure}[ht]
1246: \begin{center}
1247: \includegraphics[width=.7\textwidth]{cv.eps}
1248: \end{center}
1249: \vskip 0.25truecm
1250: \caption[]{Specific heat as a function of U
1251: in the $A=0$ ferromagnetic case.
1252: The numerical simulation at equilibrium for a system
1253: with $N=500$ is compared with the
1254: expected theoretical result (\protect\ref{cv1}).}
1255: \label{cvhmf}
1256: \end{figure}
1257: %
1258:
1259: In the HMF model and for a rather moderate size of the systems, it is
1260: possible to calculate not only the LLE but all the Lyapunov exponents,
1261: and from them the Kolmogorov-Sinai entropy.
1262: We give first a succinct definition of the spectrum
1263: of Lyapunov exponents (for more details see~\cite{Ruelleeckman}).
1264: Once the 2N-dimensional tangent vector
1265: ${\bf z}=(\delta\theta_1,\cdots,\delta\theta_N,\delta p_1,\cdots,\delta p_N)$
1266: is defined, with its dynamics given by Eqs.~(\ref{lin1}), one can formally
1267: integrate the motion in tangent space up to time $t$, since the
1268: equations are linear,
1269: \begin{equation}
1270: {\bf z}(t)=J^t{\bf z}(0)~,
1271: \label{jey}
1272: \end{equation}
1273: where $J^t$ is a $2N\times2N$ matrix that depends on time
1274: through the orbit $\theta_i(t),p_i(t)$.
1275: The first $k$ exponents of the spectrum $\lambda_1,\dots,\lambda_k$,
1276: which are ordered from the maximal to the minimal, are then given by
1277: \begin{equation}
1278: (\lambda_1 + \dots + \lambda_k) =
1279: \lim_{t \to \infty} \frac{1}{2t}
1280: \ln \mbox{Tr} J_k^t (J_k^t)^* \quad,
1281: \label{Lyapk}
1282: \end{equation}
1283: where $J_k^t$ is the matrix ($(J_k^t)^*$ its transpose)
1284: induced by $J^t$ that acts
1285: on the exterior product of $k$ vectors in the tangent space ${\bf
1286: z}_1 \land \dots \land {\bf z}_k$. The spectrum
1287: extends up to $k=2N$ and in our Hamiltonian system obeys
1288: the pairing rule
1289: \begin{equation}
1290: \lambda_i = - \lambda_{2N + 1 - i} \quad \mbox{for} \quad 1 \leq i \leq 2N~.
1291: \label{Pairing}
1292: \end{equation}
1293:
1294: The numerical evaluation of the spectrum of the Lyapunov exponents
1295: is a heavy computational task, in particular for the necessity to
1296: perform Gram-Schmidt orthonormalizations of the Lyapunov
1297: eigenvectors in order to maintain them mutually orthogonal during
1298: the time evolution. We have been able to compute the complete
1299: Lyapunov spectrum for system sizes up to $N=100$. In
1300: Fig.~\ref{spettro_f}, we report the positive part of the spectrum
1301: for different system sizes and an energy $U=0.1$ inside the weakly
1302: chaotic region. The negative part of the spectrum is symmetric due
1303: to the pairing rule (\ref{Pairing}). The limit distribution
1304: $\lambda(x)$, suggested for short range interactions,
1305: \begin{equation}
1306: \lambda(x) = \lim_{N \to \infty} \lambda_{xN}(N)~,
1307: \label{Distr}
1308: \end{equation}
1309: that is obtained by plotting $\lambda_i$ vs. $i/N$ and letting
1310: $N$ going to infinity, is found also here for the
1311: $N$ values that we have been able to explore.
1312: At higher energies, this scaling is not valid and a size-dependence
1313: is present~\cite{lat2hmf}.
1314: %
1315: \begin{figure}[ht]
1316: \begin{center}
1317: \includegraphics[width=.7\textwidth]{spettro_f.eps}
1318: \end{center}
1319: \caption[]{Scaling of the positive part of the spectrum of Lyapunov
1320: exponents in the A=0 ferromagnetic case for $U=0.1$.}
1321: \label{spettro_f}
1322: \end{figure}
1323: %
1324:
1325: The Kolmogorov-Sinai (K-S) entropy is, according to Pesin's
1326: formula~\cite{Ruelleeckman}, the sum of the positive Lyapunov
1327: exponents. In Fig.~\ref{sks_f}, we plot the entropy density
1328: $S_{KS}/N$ as a function of $U$ for different systems sizes.
1329:
1330: \begin{figure}[ht]
1331: \begin{center}
1332: \includegraphics[width=.7\textwidth]{sks_f.eps}
1333: \end{center}
1334: \caption[]{$S_{KS}/N$ as a function of $U$
1335: in the $A=0$ ferromagnetic case.Numerical calculations for different
1336: systems sizes ranging from $N=10$ to $N=100$ are shown. The dashed
1337: line indicates the critical energy.}
1338: \label{sks_f}
1339: \end{figure}
1340:
1341: As for the LLE, $S_{KS}/N$ shows a peak near the critical energy,
1342: a fast convergence to a limiting value as $N$ increases in the small
1343: energy limit, and a slow convergence to zero
1344: for $U \geq U_c$.
1345: A comparison of the ferromagnetic and
1346: antiferromagnetic cases is reported in Fig.~\ref{lle_fa}.
1347: %
1348: \begin{figure}[ht]
1349: \begin{center}
1350: \includegraphics[width=.7\textwidth]{lle_fa.eps}
1351: \end{center}
1352: \vskip0.5truecm
1353: \caption[]{LLE and $S_{KS}/N$ as a function of $U$
1354: in the $A=0$ ferromagnetic and
1355: antiferromagnetic cases for N=100.}
1356: \label{lle_fa}
1357: \end{figure}
1358: %
1359: Here, for N=100, we plot as a function of $U$ the LLE and the
1360: Kolmogorov-Sinai entropy per particle $S_{KS}/N$. In both the ferromagnetic
1361: and antiferromagnetic cases, the system is integrable in the limits of
1362: small and large energies. The main difference between the
1363: ferromagnetic and the antiferromagnetic model appears at intermediate
1364: energies. In fact, although both cases are chaotic (LLE and $S_{KS}/N$
1365: are positive), in the ferromagnetic system one observes a well defined
1366: peak just below the critical energy, because the dynamics feels the
1367: presence of the phase transition\index{phase!transition}. On the other
1368: hand, a smoother curve is observed in the antiferromagnetic case.
1369: %
1370: \begin{figure}[hb]
1371: \begin{center}
1372: \includegraphics[width=.7\textwidth]{lle_low_f.eps}
1373: \end{center}
1374: \vskip0.5truecm
1375: \caption[]{Scaling properties of the LLE at low energies in the
1376: $A=0$ ferromagnetic case. No N-dependence is observed for $U<0.2$.
1377: The dashed line indicates a power-law $U^{1/2}$.}
1378: \label{lle_low_f}
1379: \end{figure}
1380: %
1381: In the low energy regime, it is possible to work out~\cite{lat1hmf} a
1382: simple estimate $\lambda_1 \propto \sqrt{U}$, which is fully
1383: confirmed for the ferromagnetic case in Fig.~\ref{lle_low_f} for
1384: different system sizes in the range $N=100, 20000$. The same
1385: scaling law is also valid in the antiferromagnetic
1386: case~\cite{latoraprogtherophys} and for the $A=1$~\cite{at} HMF
1387: model.
1388:
1389: At variance with the $N$-independent behavior observed at small
1390: energy, strong finite size effects are present above the critical
1391: energy in the ferromagnetic case and for all energies for the
1392: antiferromagnetic case. In Fig.~\ref{lle_high_fa}(a), we show that
1393: the LLE is positive and $N$-independent below the transition (see
1394: the values $U=0.4, 0.5$), while it goes to zero with $N$ above.
1395: We also report in the same figure a calculation of the LLE using a
1396: random distribution of particle positions $\theta_i$ on the circle
1397: in Eq.~(\ref{lin1}) for the tangent vector. The agreement between
1398: the deterministic estimate and this random matrix calculation is
1399: very good. The LLE scales as $N^{-{1\over3}}$, as indicated by the
1400: fit reported in the figure. This agreement can be explained by means
1401: of an analytical result obtained for the LLE of product of random
1402: matrices~\cite{pari}. If the elements of the symplectic random
1403: matrix have zero mean, the LLE scales with the power $2/3$ of the
1404: perturbation. In our case, the latter condition is satisfied and
1405: the perturbation is the magnetization $M$. Since $M$ scales as
1406: $N^{-{1\over 2}}$, we get the right scaling of $\lambda_1$ with
1407: $N$. This proves that the system is integrable for $U \geq U_c $
1408: as $N \to \infty$. This result is also confirmed by the analytical
1409: calculations of Ref.~\cite{firpo} and, more recently, of
1410: Ref.~\cite{Vallejos}.
1411: %
1412: \begin{figure}[ht]
1413: \begin{center}
1414: \includegraphics[width=.9\textwidth]{lle_high_fa.eps}
1415: \end{center}
1416: \vskip1truecm
1417: \caption{Scaling of the LLE vs $N$ for the $A=0$ ferromagnetic and
1418: antiferromagnetic cases at various energies, see text. A power-law
1419: decreasing as $N^{-1/3}$ of the LLE is observed for overcritical energies
1420: in the ferromagnetic case and for all energies in the antiferromagnetic one.
1421: See text for further details.}
1422: \label{lle_high_fa}
1423: \end{figure}
1424: %
1425:
1426: In the antiferromagnetic case, the LLE goes to zero with
1427: system size as $N^{-{1\over 3}}$ for all values of $U$.
1428:
1429: Interesting scaling laws have also been found for the
1430: Kolmogorov-Sinai entropy in the ferromagnetic case: at small
1431: energies $S_{KS}/N \propto U^{3/4}$ with no size dependence, and
1432: $S_{KS}/N \propto N^{-1/5}$ for overcritical energy densities. The
1433: latter behavior has been found also in other
1434: models~\cite{tsuchia}. Concluding this section we would like to
1435: stress that the finite value of chaotic measures close to the
1436: critical point is strongly related to kinetic energy fluctuations and can
1437: be considered as a {\em microscopic dynamical indication} of the
1438: macroscopic equilibrium phase transition. This connection has been
1439: found also in other models and seems to be quite
1440: general~\cite{laporeport,cmd,lapo3,barrehmf}
1441:
1442: The behavior of the HMF model as a function of the range
1443: of the interaction~\cite{celiahmf,tamahmf,campa} and the dynamical
1444: features before equilibration \cite{lat0,lat0a} is discussed in a
1445: separate chapter of this volume in connection to Tsallis
1446: nonextensive thermodynamics \cite{tsallisrap}.
1447:
1448:
1449: \subsection{Mechanisms of chaos in the $A\neq 0$ case}
1450:
1451: For $A=0$, the origin of chaos is related to the non time-dependence
1452: of Eq.~(\ref{pend}), since it is obvious that if the
1453: phase $\phi$ and the magnetization $M$ would become constant the
1454: dynamics of the system will reduce to that of an integrable
1455: system. There are indeed preliminary indications~\cite{priv} that in the
1456: mean-field limit $N \to \infty$, $M$ and $\phi$ will become
1457: constant and $\lambda \to 0$ . It should be noticed that this is
1458: true if the mean-field limit is taken before the limit $t \to
1459: \infty$ in the definition of the maximal Lyapunov exponent, and
1460: numerical indications were reported in Ref.~\cite{lat0a}. When $A
1461: > 0$ we expect a quite different situation: indeed, even
1462: assuming that in the mean-field limit $M$ and $P$ and their
1463: respective phases will become constant, the dynamics will
1464: eventually take place in a $4$-dimensional phase space and chaos
1465: can in principle be observed.
1466:
1467: As already shown in~\cite{at}, for $A=1$ and $\varepsilon=1$, two
1468: different mechanisms of chaos are present in the system for $U <
1469: U_c$ : one acting on the particles trapped in the potential and
1470: another one felt by the particles moving in proximity of the
1471: separatrix. This second mechanism is well known and is related to
1472: the presence of a chaotic layer situated around the separatrix.
1473: The origin of the first mechanism is less clear, but presumably
1474: related to the erratic motion of the minimum of the potential
1475: well, i.e. to the time-dependent character of the equations ruling
1476: the dynamics of the single particle. Indications in this direction
1477: can be found by performing the following numerical experiment. Let
1478: us prepare a system with $N=200$ and $U=0.87$ (the critical energy
1479: is in this case $U_c \sim 2$) with a Maxwellian velocity
1480: distribution and with all particles in a single
1481: cluster.
1482:
1483: For an integration time $ t < 2 \times 10^6$, the Lyapunov
1484: exponent has a value $\lambda \simeq 0.13$. But when at time $t
1485: \sim 2 \times 10^6$, one particle escapes from the cluster, its
1486: value almost doubles (see Fig.~\ref{lyap}). The escaping of the
1487: particle from the cluster is associated to a decrease of the
1488: magnetization $M$ and of the kinetic energy~$K$. This last effect
1489: is related to the negative specific heat regime: the potential
1490: energy $V_A$ is minimal when all the particles are trapped, if one
1491: escapes then $V_A$ increases and due to the energy conservation
1492: $K$ decreases. As a matter of fact, we can identify a ``strong''
1493: chaos felt from the particles approaching the separatrix and a
1494: ``weak'' chaos associated to the orbits trapped in the potential
1495: well. We believe that the latter mechanism of chaotization should
1496: disappear (in analogy with the $A=0$ case) when the mean-field
1497: limit is taken before the $t \to \infty$ limit. Therefore we
1498: expect that for $N \to \infty$ the only source of chaotic
1499: behaviour should be related to the chaotic sea located around the
1500: separatrix. As already noticed in Ref.~\cite{at}, the degree of
1501: chaotization of a given system depends strongly on the initial
1502: condition (in particular in the mean-field limit). In the latter
1503: limit, for initial condition prepared in a clustered
1504: configuration, we expect that $\lambda =0$, until one particle will
1505: escape from the cluster.
1506:
1507: \begin{figure}[ht]
1508: \begin{center}
1509: \includegraphics[width=0.5\textwidth,angle=270]{f13n.eps}
1510: \end{center}
1511: \caption[]{Time evolution of the Largest Lyapunov Exponent (LLE)
1512: $\lambda_1$, of the magnetization $M$, and of the kinetic
1513: energy $K$ are shown for a clustered initial condition
1514: for the model with $A=1$ and $\varepsilon=1$ at $U=0.87$ and with $N=200$.
1515: }
1516: \label{lyap}
1517: \end{figure}
1518:
1519:
1520: \section{Conclusions}
1521:
1522: We have discussed the dynamical properties of the Hamiltonian Mean
1523: Field model in connection to ita thermodynamics. This apparently
1524: simple class of models has revealed a very rich and interesting
1525: variety of behaviours. Inequivalence of ensembles, negative specific
1526: heat, metastable dynamical states and chaotic dynamics are only some
1527: among them. During the past years these models have been of great
1528: help in understanding the connection between dynamics and
1529: thermodynamics when long-range interactions are present. Such
1530: kind of investigation is of extreme importance for self-gravitating
1531: systems and plasmas, but also for phase transitions in finite systems,
1532: such as atomic clusters or nuclei, and for the foundation of
1533: statistical mechanics. Several progresses have been done during these
1534: years. This contribution, although not exhaustive, is an
1535: effort to summarize some of the main results achieved so far.
1536: We believe that the problems which are still not understood
1537: will be hopefully clarified in the near future within a general
1538: theoretical framework. We list three important open questions that
1539: we believe can be reasonably addresses: the exact solution of the
1540: model in the microcanonical ensemble; the full characterization
1541: of the out-of-equilibrium states close to phase transitions; the
1542: clarification of the scaling laws of the maximal Lyapunov exponent
1543: and of the Lyapunov spectrum; the study of the single-particle
1544: diffusive motion~\cite{levyhmf} in the various non-equilibrium
1545: and equilibrium regimes.
1546:
1547:
1548:
1549: \section*{Acknowledgements}
1550: We would like to warmly thank our collaborators Mickael Antoni, Julien
1551: Barr{\'e}, Freddy Bouchet, Marie-Christine Firpo, Fran{\c c}ois Leyvraz and
1552: Constantino Tsallis for fruitful interactions. One of us (A.T.) would
1553: also thank Prof. Ing. P. Miraglino for giving him the opportunity to
1554: complete this paper. This work has been partially supported by the EU
1555: contract No. HPRN-CT-1999-00163 (LOCNET network), the French
1556: Minist{\`e}re de la Recherche grant ACI jeune chercheur-2001 N$^\circ$
1557: 21-311. This work is also part of the contract COFIN00 on {\it Chaos
1558: and localization in classical and quantum mechanics}.
1559:
1560:
1561:
1562: \begin{thebibliography}{8.}
1563: \addcontentsline{toc}{section}{References}
1564:
1565: \bibitem{paddyhmf} T. Padhmanaban, \emph{Statistical Mechanics of
1566: gravitating systems in static and cosmological
1567: backgrounds}, in ``Dynamics and Thermodynamics of Systems with
1568: Long Range Interactions'', T. Dauxois, S. Ruffo, E. Arimondo, M.
1569: Wilkens Eds., Lecture Notes in Physics Vol. 602, Springer (2002),
1570: (in this volume)
1571:
1572: \bibitem{chavanishmf} P.-H. Chavanis, \emph{Statistical mechanics of
1573: two-dimensional vortices and stellar systems},
1574: in ``Dynamics and Thermodynamics of Systems with Long Range
1575: Interactions'', T. Dauxois, S. Ruffo, E. Arimondo, M. Wilkens Eds.,
1576: Lecture Notes in Physics Vol. 602, Springer (2002), (in this volume)
1577:
1578:
1579: \bibitem{elskenshmf} Elskens, \emph{Kinetic theory for plasmas and
1580: wave-particle hamiltonian dynamics}, in ``Dynamics and
1581: Thermodynamics of Systems with Long Range Interactions'', T.
1582: Dauxois, S. Ruffo, E. Arimondo, M. Wilkens Eds., Lecture Notes in
1583: Physics Vol. 602, Springer (2002), (in this volume)
1584:
1585:
1586: \bibitem{tamahmf} F. Tamarit and C. Anteneodo,
1587: Physical Review Letters \textbf{84}, 208 (2000)
1588:
1589: \bibitem{BMRhmf} J. Barr{\'e}, D. Mukamel, S. Ruffo, \emph{Ensemble
1590: inequivalence in mean-field models of magnetism}, in ``Dynamics
1591: and Thermodynamics of Systems with Long Range Interactions'', T.
1592: Dauxois, S. Ruffo, E. Arimondo, M. Wilkens Eds., Lecture Notes in
1593: Physics Vol. 602, Springer (2002), (in this volume)
1594:
1595:
1596: \bibitem{campa}
1597: A. Campa, A. Giansanti, D. Moroni, Physical Review E
1598: \textbf{62}, 303 (2000) and Chaos Solitons and Fractals
1599: \textbf{13}, 407 (2002)
1600:
1601: \bibitem{Ruelleeckman}
1602: J.-P. Eckmann, D. Ruelle, Review of Modern Physics \textbf{57},
1603: 615 (1985)
1604:
1605: \bibitem{Antoni} S. Ruffo, in {\it Transport and Plasma Physics}, edited by
1606: S. Benkadda, Y. Elskens and F. Doveil (World Scientific, Singapore,
1607: 1994), pp. 114-119; M. Antoni, S. Ruffo, Physical Review E \textbf{52},
1608: 2361 (1995)
1609:
1610: \bibitem{at} M. Antoni and A. Torcini, Physical Review E \textbf{57},
1611: R6233 (1998); A. Torcini and M. Antoni, Physical Review E
1612: \textbf{59}, 2746 (1999)
1613:
1614: \bibitem{art} M. Antoni, S. Ruffo, A. Torcini, {\it First and second
1615: order phase transitions for a system with infinite-range
1616: attractive interaction}, Physical Review E, to appear (2002),
1617: [cond-mat/0206369]
1618:
1619:
1620: \bibitem{tsallisrap} C. Tsallis, A. Rapisarda, V. Latora and F.
1621: Baldovin \emph{Nonextensivity: from low-dimensional maps to
1622: Hamiltonian systems}, in ``Dynamics and Thermodynamics of Systems
1623: with Long Range Interactions'', T. Dauxois, S. Ruffo, E. Arimondo,
1624: M. Wilkens Eds., Lecture Notes in Physics Vol. 602, Springer (2002),
1625: (in this volume)
1626:
1627: \bibitem{lat1hmf} V. Latora, A. Rapisarda and S. Ruffo,
1628: Physical Review Letters \textbf{80}, 692 (1998)
1629:
1630: \bibitem{celiahmf} C. Anteneodo and C. Tsallis,
1631: Physical Review Letters \textbf{80}, 5313 (1998)
1632:
1633: \bibitem{livipolruffo}
1634: R. Livi, A. Politi, S. Ruffo, Journal of Physics A \textbf{19},
1635: 2033 (1986)
1636:
1637: \bibitem{grossdh} D.H. Gross, \emph{Thermo-Statistics or Topology of
1638: the Microcanonical Entropy Surface}, in ``Dynamics and
1639: Thermodynamics of Systems with Long Range Interactions'', T.
1640: Dauxois, S. Ruffo, E. Arimondo, M. Wilkens Eds., Lecture Notes in
1641: Physics Vol. 602, Springer (2002), (in this volume)
1642:
1643:
1644: \bibitem{Kachmf}
1645: M. Kac, G. Uhlenbeck and P.C. Hemmer, Journal Mathematical Physics
1646: (N.Y.) {\bf 4}, 216 (1963); {\bf 4}, 229 (1963); {\bf 5}, 60
1647: (1964)
1648:
1649: \bibitem{priv} S. Tanase-Nicola and J. Kurchan, Private Communication (2002)
1650:
1651: \bibitem{barrethesis}
1652: J. Barr{\'e}, PhD Thesis, ENS Lyon, unpublished (2002)
1653:
1654: \bibitem{grossbook} D.H.E. Gross, {\it Microcanonical thermodynamics,
1655: Phase transitions in ``small'' systems} (World Scientific,
1656: Singapore, 2000)
1657:
1658:
1659: \bibitem{mac}
1660: H. Yoshida, {Physics Letters A} \textbf{150}, 262 (1990); R. I.
1661: McLachlan and P. Atela, Nonlinearity {\bf 5}, 541 (1992)
1662:
1663: \bibitem{thir} P. Hertel and W. Thirring,
1664: {Annals of Physics} \textbf{63}, 520 (1971)
1665:
1666: \bibitem{compa}
1667: A. Compagner, C. Bruin, A. Roelse, {Physical Review} A
1668: \textbf{39}, 5989 (1989); H.A. Posch, H. Narnhofer, W. Thirring,
1669: Physical Review A {\bf 42}, 1880 (1990)
1670:
1671: \bibitem{levyhmf}
1672: V. Latora, A. Rapisarda and S. Ruffo,
1673: Physical Review Letters \textbf{83}, 2104 (1999)
1674:
1675: \bibitem{lat0} V. Latora, A. Rapisarda and C. Tsallis, Physical Review E
1676: \textbf{64}, 056124-1 (2001)
1677:
1678: \bibitem{lat0a} V. Latora, A. Rapisarda and C. Tsallis, Physica A
1679: \textbf{305}, 129 (2002)
1680:
1681: \bibitem{Diego} D. Del Castillo-Negrete, \emph{Dynamics and
1682: self-consistent chaos in a mean field Hamiltonian model}, in
1683: ``Dynamics and Thermodynamics of Systems with Long Range
1684: Interactions'', T. Dauxois, S. Ruffo, E. Arimondo, M. Wilkens Eds.,
1685: Lecture Notes in Physics Vol. 602, Springer (2002), (in this volume)
1686:
1687:
1688: \bibitem{barreepjB}
1689: J. Barr{\'e}, F. Bouchet, T. Dauxois, S.~Ruffo, \emph{Birth and
1690: long-time stabilization of out-of-equilibrium coherent
1691: structures}, submitted to European Physical Journal B (2002) and
1692: Physical Review Letters, to appear (2002)
1693:
1694: \bibitem{drh} T. Dauxois, P.~Holdsworth, S.~Ruffo,
1695: European Physical Journal B \textbf{16}, 659 (2000)
1696:
1697: \bibitem{firpoleyvraz} M.C. Firpo, F. Leyvraz and S. Ruffo,
1698: Journal of Physics A, \textbf {35}, 4413 (2002)
1699:
1700: \bibitem{BraunHepp} W. Braun, K. Hepp, Communications in Mathematical
1701: Physics \textbf{56}, 101 (1977)
1702:
1703: \bibitem{Kapitaza}
1704: P. L. Kapitza, in \emph{Collected Papers of P. L. Kapitza}, edited
1705: by D. Ter Harr (Pergamon, London, 1965), pp 714, 726
1706:
1707: \bibitem{piston} J. L. Lebowitz, J. Piasecki, Ya. Sinai "Scaling
1708: dynamics of a massive piston in an ideal gas" , in Hard ball
1709: systems and the Lorentz gas, 217-227, Encycl. Math. Sci., 101,
1710: Springer, Berlin (2000)
1711:
1712: \bibitem{sinairusse} Ya. Sinai, Theoretical and Mathematical Physics
1713: (in Russian), {\bf 121}, 110 (1999)
1714:
1715: \bibitem{plasma} M. Antoni, Y. Elskens, D. F. Escande, Physics of
1716: Plasmas {\bf 5}, 841 (1998); M-C. Firpo, \emph{Etude dynamique et
1717: statistique de l'interaction onde-particule}, PhD Thesis, Universit{\'e}
1718: de Marseille (1999)
1719:
1720: \bibitem{embidmajda} P. F. Embid, A. J. Majda, Communications in Partial
1721: Differential Equations {\bf 21}, 619 (1996)
1722:
1723: \bibitem{laporeport}
1724: L. Casetti, M. Pettini, E.G.D. Cohen, Physics Reports \textbf{337},
1725: 237 (2000)
1726:
1727: \bibitem{cmd}
1728: A. Bonasera, V. Latora and A. Rapisarda, {Physical Review Letters}
1729: \textbf{75}, 3434 (1995)
1730:
1731: \bibitem{ben}
1732: G. Benettin, L. Galgani and J.M. Strelcyn, {Physical Review} A
1733: \textbf{14}, 2338 (1976)
1734:
1735: \bibitem{nayak}
1736: S.K. Nayak, R.Ramaswamy and C. Chakravarty, Physical Review E
1737: \textbf{51}, 3376 (1995); V. Mehra, R. Ramaswamy, {Physical
1738: Review} E \textbf{56}, 2508 (1997)
1739:
1740: \bibitem{yama}
1741: Y.Y. Yamaguchi, {Progress of Theoretical Physics} \textbf{95},
1742: 717 (1996)
1743:
1744: \bibitem{firpo}
1745: M-C. Firpo, {Physical Review E} \textbf{57}, 6599 (1998)
1746:
1747: \bibitem{lat2hmf} V. Latora, A. Rapisarda and S. Ruffo,
1748: Physica D \textbf{131}, 38 (1999); Physica A \textbf{280}, 81 (2000)
1749:
1750: \bibitem{lapo1}
1751: L. Casetti, C. Clementi and M. Pettini, {Physical Review} E {\bf
1752: 54}, 5969 (1996); L. Caiani, L. Casetti, C. Clementi and M.
1753: Pettini, {Physical Review Letters} {\bf 79}, 4361, (1997)
1754:
1755: \bibitem{lpv}
1756: J.L. Lebowitz, J.K. Percus and L. Verlet, Physical Review A
1757: \textbf{153}, 250 (1967)
1758:
1759: \bibitem{latoraprogtherophys} V. Latora, A. Rapisarda, S. Ruffo,
1760: Progress of Theoretical Physics Supplement \textbf{139}, 204
1761: (2000)
1762:
1763:
1764: \bibitem{pari}
1765: G. Parisi and A. Vulpiani, Journal of Physics A \textbf{19}, L425
1766: (1986)
1767:
1768: \bibitem{tsuchia}
1769: T. Tsuchiya and N. Gouda, Physical Review E \textbf{61}, 948
1770: (2000)
1771:
1772: \bibitem{lapo3}
1773: L. Caiani, L. Casetti, C. Clementi, G. Pettini, M. Pettini and R.
1774: Gatto, {Physical Review E} \textbf{57}, 3886 (1998)
1775:
1776:
1777: \bibitem{barrehmf}
1778: J. Barr{\'e} and T. Dauxois, Europhysics Letters \textbf{55},
1779: 164 (2001)
1780:
1781: \bibitem{Vallejos} C. Anteneodo and R. Vallejos,
1782: Physical Review E \textbf{65}, 016210 (2002)
1783:
1784:
1785: \end{thebibliography}
1786:
1787: \end{document}
1788: