1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: % REVTEX FILE
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: \documentstyle[aps,pre,amsfonts,graphicx]{revtex}
5:
6: % \setlength{\textwidth}{6.5in}
7: % \setlength{\textheight}{9.0in}
8: % \setlength{\topmargin}{0.0in}
9: % \setlength{\oddsidemargin}{0.0in}
10:
11: \newcommand{\beq}{\begin{equation}}
12: \newcommand{\eeq}{\end{equation}}
13: \def\erre{\hbox{\rm\rlap{I}\kern.1em R}}
14:
15: \newcommand{\be}{\begin{equation}}
16:
17: \begin{document}
18:
19: \title{Hamiltonian dynamics and geometry of phase transitions in classical
20: XY models}
21:
22: \author{Monica Cerruti-Sola$^{1,3,}$\cite{moni},
23: Cecilia Clementi$^{2,}$\cite{cecilia},
24: and Marco Pettini$^{1,3,}$\cite{marco}}
25: \address{
26: $^1$Osservatorio Astrofisico di Arcetri, Largo E. Fermi 5, I-50125 Firenze,
27: Italy \\
28: $^2$Department of Physics, University of California at San Diego,
29: La Jolla, CA 92093-0319, USA \\
30: $^3$Istituto Nazionale per la Fisica della Materia, Unit\`a di Ricerca di
31: Firenze, Italy }
32:
33: \date{\today}
34:
35: \maketitle
36:
37: \begin{abstract}
38: The Hamiltonian dynamics associated to classical, planar, Heisenberg XY
39: models is investigated for two- and three-dimensional lattices.
40: Besides the conventional signatures of phase transitions, here obtained
41: through time averages of thermodynamical observables in place of ensemble
42: averages, qualitatively new information is derived from the
43: temperature dependence of Lyapunov exponents. A Riemannian geometrization
44: of newtonian dynamics suggests to consider other observables of geometric
45: meaning tightly related with the largest Lyapunov exponent. The numerical
46: computation of these observables - unusual in the study of phase transitions -
47: sheds a new light on the microscopic dynamical counterpart of
48: thermodynamics also pointing to the existence of some major change in the
49: geometry of the mechanical manifolds at the thermodynamical transition.
50: Through the microcanonical definition of the entropy, a relationship between
51: thermodynamics and the extrinsic geometry of the constant energy surfaces
52: $\Sigma_E$ of phase space can be naturally established. In this framework,
53: an approximate formula is worked out, determining a highly non-trivial
54: relationship between temperature and topology of the $\Sigma_E$. Whence it can
55: be understood that the appearance of a phase transition must be tightly
56: related to a suitable major topology change of the $\Sigma_E$. This contributes
57: to the understanding of the origin of phase transitions in the
58: microcanonical ensemble.
59: \end{abstract}
60: \pacs{PACS: 05.45.+b; 05.20.-y}
61: \newpage
62:
63: \section{Introduction}
64: \label{intro}
65: The present paper deals with the study of the microscopic Hamiltonian
66: dynamical phenomenology associated to thermodynamical phase transitions.
67: This general subject is addressed in the special case of planar, classical
68: Heisenberg XY models in two and three spatial dimensions.
69: A preliminary presentation of some of the results and ideas contained in
70: this paper has been already given in \cite{CCCP}.
71:
72: There are several reasons to tackle the Hamiltonian dynamical counterpart of
73: phase transitions. On the one hand, we might wonder whether our knowledge
74: of the already wide variety of dynamical properties of Hamiltonian systems
75: can be furtherly enriched by considering the dynamical signatures, if any,
76: of phase transitions. On the other hand, it is {\it a-priori} conceivable
77: that also the theoretical investigation of the phase transition phenomena
78: could benefit of a direct investigation of the natural microscopic dynamics.
79: In fact, from a very general point of view, we can argue that in those times
80: where microscopic dynamics was completely unaccessible to any kind of
81: investigation, statistical mechanics has been invented just to replace
82: dynamics. During the last decades, the advent of powerful computers
83: has made
84: possible, to some extent, a direct access to microscopic dynamics through the
85: so called molecular dynamical simulations of the statistical properties of
86: "macroscopic" systems.
87:
88: Molecular dynamics can be either considered as a mere alternative to Monte
89: Carlo methods in practical computations, or it can be also seen as a possible
90: link to concepts and methods (those of nonlinear Hamiltonian dynamics)
91: that could deepen our insight about phase transitions.
92: In fact, by construction,
93: the ergodic invariant measure of the Monte Carlo stochastic dynamics, commonly
94: used in numerical statistical mechanics, is
95: the canonical Gibbs distribution, whereas there is no general result that
96: guarantees the ergodicity and mixing of natural (Hamiltonian) dynamics.
97: Thus, the general interest for any contribution that helps in clarifying
98: under what conditions equilibrium statistical mechanics correctly describes
99: the average properties of a large collection of particles, safely replacing
100: their microscopic dynamical description.
101:
102: Actually, as it has been already shown and confirmed by the
103: results reported below, there are some intrinsically dynamical observables
104: that clearly signal the existence of a phase transition.
105: Notably, Lyapunov exponents appear as sensitive measurements for phase
106: transitions.
107: They are also probes of a hidden geometry of the dynamics, because Lyapunov
108: exponents depend on the geometry of certain ``mechanical manifolds''
109: whose geodesic flows coincide with the natural motions.
110: Therefore, a peculiar energy -- or temperature -- dependence of the largest
111: Lyapunov exponent at a phase transition point also reflects some important
112: change in the geometry of the mechanical manifolds.
113:
114: As we shall discuss throughout the present paper, also the topology of these
115: manifolds has been discovered to play a relevant role in the phase transition
116: phenomena (PTP).
117:
118: Another strong reason of interest for the Hamiltonian dynamical
119: counterpart of PTP is related to the equivalence problem of statistical
120: ensembles. Hamiltonian dynamics has its most natural and tight relationship
121: with microcanonical ensemble. Now, the well known equivalence among
122: all the statistical ensembles in the thermodynamic limit is valid in general
123: in the absence of thermodynamic singularities, i.e. in the absence of phase
124: transitions. This is not a difficulty for statistical mechanics
125: as it might seem at first sight
126: \cite{Gallavotti}, rather, this is a very interesting and intriguing point.
127:
128: The inequivalence of canonical and microcanonical ensembles in presence of
129: a phase transition has been analytically shown for a particular model by
130: Hertel and Thirring \cite{Thirring}, it is mainly revealed by the
131: appearance
132: of negative values of the specific heat and has been discussed by several
133: authors \cite{Lyndenbell,Gross}.
134:
135: The microcanonical description of phase transitions seems also to offer many
136: advantages in tackling first order phase transitions \cite{Gross2}, and
137: seems considerably
138: less affected by finite-size scaling effects with respect to the canonical
139: ensemble description \cite{Gross1}.
140: This non-equivalence problem, together with certain advantages of the
141: microcanonical ensemble, strenghtens the interest for the Hamiltonian dynamical
142: counterpart of PTP. Let us briefly mention the existing contributions in the
143: field.
144:
145: Butera and Caravati \cite{Butera},
146: considering an XY model in two dimensions, found that
147: the temperature dependence of the largest Lyapunov exponent changes just
148: near the critical temperature $T_c$ of the
149: Kosterlitz-Thouless phase transition. Other interesting aspects of the
150: Hamiltonian dynamics of the XY model in two dimensions have been extensively
151: considered in \cite{Leoncini}, where a very rich phenomenology is reported.
152: Recently, the behaviour of Lyapunov exponents
153: has been studied in Hamiltonian dynamical systems: {\it i)} with long-range
154: interactions \cite{Rapisarda,Ruffo,Antoni}, {\it ii)} describing either
155: clusters of particles or magnetic or gravitational models exhibiting phase
156: transitions, {\it iii)} in classical
157: lattice field theories with $O(1)$, $O(2)$ and $O(4)$ global symmetries
158: in two and three space dimensions \cite{CCP1,CCCPPG}, {\it iv)} in the XY
159: model in two and three space dimensions
160: \cite{CCCP}, {\it v)} in the "$\Theta$ - transition" of homopolymeric chains
161: \cite{polimeri}.
162: The pattern of $\lambda(T)$ close to the critical temperature $T_c$ is
163: model-dependent. The behaviour of Lyapunov exponents near the transition point
164: has been considered also in the case of first- order phase transitions
165: \cite{Dellago,Mehra}. It is also worth mentioning the very intriguing result
166: of Ref.\cite{Berry}, where a glassy transition is accompanied by a sharp
167: jump of $\lambda(T)$.
168:
169: $\lambda(T)$ always detects a phase transition and, even if
170: its pattern close to the critical temperature $T_c$ is
171: model-dependent, it can be used as an order parameter -- of dynamical origin --
172: also in the absence of a standard order parameter (as in the case of the
173: mentioned "$\Theta$-transition" of homopolymers and of the glassy transition
174: in amorphous materials).
175: This appears of great prospective interest also in the light of recently
176: developed analytical methods to compute Lyapunov exponents (see Section IV).
177:
178: Among Hamiltonian models with long-range
179: interactions exhibiting phase
180: transitions, the most extensively studied is the
181: mean-field XY model \cite{Ruffo,Firpo,Ruffo_prl,Ruffo_talk}, whose
182: equilibrium statistical mechanics
183: is exactly described, in the thermodynamic limit, by mean-field
184: theory \cite{Ruffo}. In this system,
185: the theoretically predicted temperature dependence of the largest Lyapunov
186: exponent $\lambda$ displays a non-analytic behavior at the phase transition
187: point.
188:
189: The aims of the present paper are
190: \begin{itemize}
191: \item
192: to investigate the dynamical phenomenology of Kosterlitz-Thouless and of
193: second order phase transitions in the $2d$ and $3d$ classical
194: Heisenberg XY models
195: respectively;
196: \item
197: to highlight the microscopic dynamical counterpart of phase transitions
198: through the temperature dependence of the Lyapunov exponents, also providing
199: some physical interpretation of abstract quantities involved in
200: the geometric theory of chaos (in particular among vorticity, Lyapunov
201: exponents and sectional curvatures of configuration space);
202: \item
203: to discuss the hypothesis that phase transition phenomena could be originated
204: by suitable changes in the topology of the constant energy hypersurfaces of
205: phase space, therefore hinting to a mathematical characterization of phase
206: transitions in the microcanonical ensemble.
207: \end{itemize}
208:
209: The paper is organized as follows: Sections $II$ and $III$ are devoted to
210: the dynamical investigation of the $2d$ and $3d$ XY models respectively.
211: In Section $IV$ the geometric description of chaos is considered, with
212: the analytic derivation of the temperature dependence of the largest Lyapunov
213: exponent, the geometric signatures of a second-order phase transition and
214: the topological hypothesis.
215: Section $V$ contains a presentation of the relationship between the extrinsic
216: geometry and topology of the energy hypersurfaces of phase space and
217: thermodynamics; the results of some numeric computations are also reported.
218: Finally, Section $VI$ is devoted to summarize the achievements reported in
219: the present paper and to discuss their meaning.
220:
221: \medskip
222: \section{$2d$ XY model}
223: \medskip
224:
225: We considered a system of planar, classical ``spins'' (in fact rotators)
226: on a square lattice of $N=n\times n$ sites, and interacting through
227: the ferromagnetic interaction $V=-\sum_{\langle i,j\rangle}J{\bf S}_{i}
228: \cdot {\bf S}_{j}$ (where $\vert{\bf S}_i\vert =1)$.
229: The addition of standard, i.e. quadratic, kinetic energy
230: term leads to the following choice of the Hamiltonian
231: \beq
232: H= \sum_{i,j=1}^{n} \left \{ \frac{p_{i,j}^2}{2}+J[2- \cos(q_{i+1,j}-q_{i,j})
233: -\cos(q_{i,j+1}-q_{i,j})]\right\}~~,
234: \label{xy2d}
235: \eeq
236: where $q_{i,j}$ are the angles with respect to a fixed direction on
237: the reference plane of the system.
238: In the usual definition of the XY model both the kinetic term and the
239: constant term $2JN$ are lacking; however, their contribution does not modify
240: the thermodynamic averages (because they usually depend only on the
241: configurational partition function,
242: $Z_C=\int\prod_{i=1}^N dq_i\exp[-\beta V(q)]$,
243: the momenta being trivially integrable when the kinetic energy is quadratic).
244: Thus, as we tackle classical systems, the choice of a quadratic
245: kinetic energy term
246: is natural because it corresponds to $\frac{1}{2}\sum_{i=1}^N \vert
247: {\bf\dot S}_i\vert^2$, written in terms of the momenta $p_{i,j}$ canonically
248: conjugated to the lagrangian coordinates $q_{i,j}$. The constant term $2JN$ is
249: introduced to make the low energy expansion of Eq. (\ref{xy2d}) coincident
250: with the
251: Hamiltonian of a system of weakly coupled harmonic oscillators.
252:
253: The theory predicts for this model a Kosterlitz-Thouless phase transition
254: occurring at a critical temperature
255: estimated around $T_c\sim J$. Many Monte Carlo
256: simulations of this model have been done in order to check the predictions
257: of the theory. Among them, we quote those of
258: Tobochnik and Chester \cite{TobChes} and of Gupta and Baillie \cite{Gupta}
259: which, on the basis of accurate numerical analysis, confirmed the predictions
260: of the theory and fixed the critical temperature at $T_c=0.89$ ($J=1$).
261:
262: The analysis of the present work is based on the numerical integration
263: of the equations of motion derived from Hamiltonian (\ref{xy2d}).
264: The numerical integration is performed by means of
265: a bilateral, third order, symplectic algorithm \cite{Lapo}, and it is
266: repeated at several values of the energy density
267: $\epsilon = E/N$ ($E$ is the total energy of the system which depends upon
268: the choice of the initial conditions).
269: While the Monte Carlo simulations perform statistical averages in the canonical
270: ensemble, Hamiltonian dynamics has its statistical counterpart in the
271: microcanonical ensemble. Statistical averages are here replaced by time
272: averages of relevant observables. In this perspective, from the microcanonical
273: definition of temperature $1/T=\partial S/
274: \partial E$, where $S$ is the entropy,
275: two definitions of temperature are available: $T=\frac{2}{N}\langle K\rangle$
276: (where $K$ is the kinetic energy per degree of freedom), if
277: $S=\log \int\prod_{i=1}^N dq_i dp_i \Theta (H(p,q) - E)$, where $\Theta
278: (\cdot)$ is the Heaviside step function, and
279: $\tilde T=\left[\left(\frac{N}{2}-1\right)\langle K^{-1}\rangle\right]^{-1}$,
280: if $S=\log \int\prod_{i=1}^N dq_i dp_i \delta (H(p,q) - E)$ \cite{Pearson}.
281: $T$ (or $\tilde T$) are numerically
282: determined by measuring the time average of the kinetic energy $K$
283: per degree of freedom (or its inverse), i.e.
284: $T=\lim_{t\rightarrow\infty}\frac{2}{N}
285: \frac{1}{t}\int_0^td\tau K(\tau)$ (and similarly for $\tilde T$).
286: There is no appreciable difference in the outcomes of the computations of
287: temperature according to these two definitions.
288:
289: \medskip
290: \subsection{Dynamical analysis of thermodynamical observables}
291: \medskip
292:
293: \subsubsection{Order parameter}
294:
295: The order parameter for a system of planar ``spins'' whose Hamiltonian
296: is invariant under the action of the group
297: $O(2)$, is the bidimensional vector
298: \beq
299: {\bf M}= (\sum_{i,j=1}^{n}{\bf S}^x_{i,j},
300: \sum_{i,j=1}^{n}{\bf S}^y_{i,j})\equiv (\sum_{i,j=1}^{n}\cos q_{i,j},
301: \sum_{i,j=1}^{n} \sin q_{i,j}),
302: \label{order_par}
303: \eeq
304: which describes the mean spin orientation field.
305: After the
306: Mermin-Wagner theorem, we know that no symmetry-breaking transition can
307: occur in one and two dimensional systems with a continuous symmetry and
308: nearest-neighbour interactions. This means that, at any non-vanishing
309: temperature,
310: the statistical average of the total magnetization vector is necessarily
311: zero in the thermodynamic limit.
312: However, a vanishing magnetization is not necessarily expected
313: when computed by means of Hamiltonian dynamics at
314: finite $N$.
315: In fact, statistical averages are equivalent to averages computed through
316: suitable markovian Monte Carlo dynamics that {\it a-priori} can reach
317: any region of phase space,
318: whereas in principle a true ergodicity breaking is possible in
319: the case of differentiable dynamics. Also an "effective" ergodicity breaking
320: of differentiable dynamics is possible, when the relaxation times -- of time
321: to ensemble averages -- are very fastly increasing with $N$ \cite{Palmer}.
322:
323: This model has two integrable limits:
324: coupled harmonic oscillators and free rotators,
325: at low and high temperatures respectively. Hereafter, $T$ is meant in units
326: of the coupling constant $J$.
327:
328: For a lattice of $N = 10 \times10$ sites,
329: Figure \ref{figura.spin2d.10e10} shows that
330: at low temperatures ($T<0.5$)-- being the system almost harmonic --
331: we can observe a persistent
332: memory of the total magnetization associated with the initial condition,
333: which, on the typical time scales of our numeric simulations ($10^6$ units
334: of proper time), looks almost frozen.
335:
336: By raising the temperature above a first threshold $T_0\simeq 0.6$,
337: the total magnetization vector -- observed on the same time scale --
338: starts rotating on the plane where it is confined.
339: A further increase of the temperature induces a faster rotation of the
340: magnetization vector together with a slight reduction of its average modulus.
341:
342: At temperatures slightly greater than $1$, we observe that already at
343: $N=10 \times10$ a random variation of the direction and of the modulus of
344: the vector ${\bf M}(t)$ sets in.
345:
346: At $T>1.2$, we observe a fast relaxation and, at high temperatures
347: ($T\simeq 10$), a sort
348: of saturation of chaos.
349:
350: At a first glance, the results reported in Fig.
351: \ref{figura.spin2d.10e10} could suggest
352: the presence of a phase transition associated with the breaking of the
353: $O(2)$ symmetry.
354: In fact, having in mind the Landau theory, the ring-shaped distribution of
355: the instantaneous magnetization shown
356: by Fig. \ref{figura.spin2d.10e10} is the typical signature of an
357: $O(2)$-broken symmetry phase and the spot-like patterns around zero are
358: proper to the unbroken symmetry phase.
359:
360: The apparent contradiction of these results with the Mermin-Wagner theorem
361: is resolved by checking whether the observed phenomenology is stable with
362: $N$. Thus, some simulations have been performed at larger values of $N$.
363: At any temperature, we found that the average modulus
364: $\langle\vert{\bf M}(t)\vert\rangle_t$ of the vector ${\bf M}(t)$,
365: computed along the trajectory, systematically
366: decreases by increasing $N$. However,
367: for temperatures lower than $T_0$, the $N$-dependence
368: of the order parameter is very weak, whereas,
369: for temperatures greater than $T_0$, the $N$-dependence
370: of the order parameter is rather strong.
371: In Fig. \ref{fig.spin2d.t0.74} two extreme cases
372: ($N=10 \times10$ and $N=200 \times200$) are shown for $T = 0.74$.
373: The systematic trend of $\langle\vert{\bf M}(t)\vert\rangle$
374: toward smaller values at increasing $N$ is
375: consistent with its expected vanishing in the limit $N\rightarrow\infty$.
376:
377: At $T = 1$, Fig. \ref{fig.spin2d.t1} shows that,
378: when the lattice dimension is greater than $50\times 50$,
379: ${\bf M}(t)$ displays random variations both in direction (in the interval
380: [0,$2\pi$]) and in
381: modulus (between zero and a value which is smaller at larger
382: $N$).
383:
384: \medskip
385: \subsubsection{Specific heat}
386: \medskip
387: By means of the recasting of a standard formula which relates the average
388: fluctuations of a generic observable computed in canonical and
389: microcanonical ensembles \cite{LPV}, and by specializing it to the
390: kinetic energy fluctuations, one obtains a microcanonical estimate
391: of the canonical specific heat
392: \begin{equation}
393: c_{V}(T)=\frac{C_V}{N}\rightarrow
394: \cases{
395: %\left\{\begin{array}{1}
396: c_{V}(\epsilon)=\displaystyle{ \frac{k_{B}d}{2}\left[1-\frac{N
397: d}{2}\frac{\langle K^{2}\rangle -\langle K\rangle
398: ^{2}}{\langle K\rangle ^{2}}\right]^{-1}} ~~,\cr
399: T=T(\epsilon)\cr }
400: %\end{array} \right.
401: \label{specheat}
402: \end{equation}
403: where $d$ is the number of degrees of freedom for each particle.
404: Time averages of the kinetic energy fluctuations computed at any
405: given value of the energy density $\epsilon$ yield $C_V(T)$, according
406: to its parametric definition in Eq.(\ref{specheat}).
407:
408: >From the microcanonical definition $1/C_V=\partial T(E)/\partial E$ of the
409: constant volume specific heat, a formula can be worked out \cite{Pearson},
410: which is exact at
411: {\it any} value of $N$ (at variance with the expression (\ref{specheat})).
412: It reads
413: \begin{equation}
414: c_V=\frac{C_V}{N}=[N - (N - 2)\langle K\rangle\langle K^{-1}\rangle]^{-1}
415: \label{cvmicro}
416: \end{equation}
417: and it is the natural expression to be used in Hamiltonian
418: dynamical simulations of
419: finite systems.
420:
421: The numerical simulations of the Hamiltonian dynamics of the $2d$
422: XY model -- computed with both Eqs.(\ref{specheat}) and (\ref{cvmicro}) --
423: yield a cuspy pattern for $c_V(T)$ peaked at $T\simeq 1$
424: (Fig. \ref{calspec_2d}).
425: This is in good agreement with the outcomes of canonical Monte Carlo
426: simulations reported in Ref. \cite{TobChes,Gupta},
427: where a pronounced peak of $c_{V}(T)$ was detected at $T \simeq 1.02$.
428:
429: By varying the lattice dimensions, the peak height remains
430: constant, in agreement with the absence of a symmetry-breaking phase
431: transition.
432: \medskip
433: \subsubsection{Vorticity}
434: \medskip
435: Another thermodynamic observable which can be studied
436: is the vorticity of the system. Let us briefly recall that if
437: the angular differences of nearby ``spins'' are small, we can suppose the
438: existence of a continuum limit function $\theta({\bf r})$ that conveniently
439: fits a given spatial configuration of the system.
440: Spin waves correspond to regular patterns of $\theta({\bf r})$, whereas the
441: appearance of a singularity in $\theta({\bf r})$ corresponds to a topological
442: defect, or a vortex, in the ``spin'' configuration. When such a defect is
443: present, along any closed path ${\cal C}$ that contains the centre of the
444: defect, one has
445: \begin{equation}
446: \oint_{\cal C}\nabla\theta({\bf r})\cdot d{\bf r}= 2\pi q~,~~~~q=0,\pm 1,\pm 2,
447: \dots
448: \end{equation}
449: indicating the presence of a vortex whose intensity is $q$. For
450: a lattice model with periodic boundary conditions, there is an equal number
451: of vortices and antivortices (i.e. vortices rotating in opposite directions).
452: Thus, the vorticity of our model can be defined as the mean total number of
453: equal sign vortices per unit volume.
454: In order to compute the vorticity ${\cal V}$ as a function of temperature,
455: we have averaged the number of positive vortices along the numerical phase
456: space trajectories. On the lattice, ${\bf r}$ is replaced by the multi-index
457: ${\bf i}$ and $\nabla_\mu\theta_i= q_{{\bf i}+\mu} - q_{\bf i}$, then the
458: number of elementary vortices is counted: the discretized version of
459: $\oint_\square\nabla\theta\cdot d{\bf r}=1$ amounts to one elementary vortex
460: on a plaquette. Thus ${\cal V}$ is obtained by summing over all the plaquettes.
461:
462: Our results are in agreement with the values obtained
463: by Tobochnik and Chester \cite{TobChes}
464: by means of Monte Carlo simulations with $N=60\times 60$.
465:
466: As shown in Fig. \ref{fig.vort2d},
467: on the $10\times 10$ lattice,
468: the first vortex shows up at $T\sim0.6$
469: and on the $40\times 40$ lattice
470: at $T\sim0.5$, when the
471: system changes its dynamical behavior,
472: increasing its chaoticity (see next Subsection).
473: At lower temperatures, vortices are less probable,
474: due to the fact
475: that the formation of vortex has a minimum energy cost.
476: Below $T\sim1$, the vortex density steeply grows
477: with a power law ${\cal V}(T)\sim T^{10}$.
478: The growth of ${\cal V}$ then slows down, until the saturation is reached at
479: $T\sim10$.
480:
481: \medskip
482: \subsection{Lyapunov exponents and chaoticity}
483: \medskip
484:
485: The values of the largest Lyapunov exponent $\lambda_{1}$
486: have been computed using the standard tangent dynamics equations [see Eqs.
487: (\ref{eqdintang}) and (\ref{bgs})], and are reported
488: in Fig. \ref{xy2d.lyap.num.fig}.
489:
490: Below $T\simeq 0.6$,
491: the dynamical behavior is nearly the same as that of harmonic
492: oscillators and the excitations of the system are only ``spin-waves''.
493:
494: In the interval $[0., 0.6]$, the observed temperature dependence
495: $\lambda_1(T) \sim T^2$ is equivalent to the
496: $\lambda_1(\epsilon) \sim\epsilon^2$ dependence (since at low temperature
497: $T(\epsilon) \propto \epsilon$), already found -- analytically and
498: numerically --
499: in the quasi-harmonic regime of other systems and characteristic of weakly
500: chaotic dynamics \cite{CCP}.
501:
502: Above $T \simeq0.6$, vortices begin to form and correspondingly the largest
503: Lyapunov exponent signals a "qualitative" change of the dynamics through a
504: steeper increase vs. $T$.
505:
506: At $T \simeq0.9$, where the theory predicts a Kosterlitz - Thouless phase
507: transition, $\lambda_1(T)$ displays an inflection point.
508:
509: Finally, at high temperatures, the power law $\lambda_{1}(T)\sim T^{-1/6}$
510: is found.
511:
512:
513: \medskip
514: \section{$3d$ XY model}
515: \medskip
516:
517: In order to extend the dynamical investigation
518: to the case of second-order phase transitions, we have
519: studied a system described by an Hamiltonian having at the same time
520: the main characteristics
521: of the $2d$ model and the differences necessary to the appearance
522: of a spontaneous symmetry-breaking below a certain critical temperature.
523: The model
524: we have chosen is such that the spin rotation is constrained on a plane and
525: only the lattice dimension has been increased, in order to elude the
526: ``no go'' conditions of the Mermin-Wagner theorem.
527: This is simply achieved by tackling a system
528: defined on a cubic lattice of $N=n\times n\times n$ sites
529: and described by the Hamiltonian
530: \begin{eqnarray}
531: H&=& \sum_{i,j,k=1}^{n}
532: \{ \frac{p_{i,j,k}^{2}}{2}+J[3-\cos(q_{i+1,j,k}-q_{i,j,k})-\nonumber\\
533: &-&\cos(q_{i,j+1,k}-q_{i,j,k})-\cos(q_{i,j,k+1}-q_{i,j,k})] \}~~.
534: \end{eqnarray}
535:
536: \medskip
537: \subsection{Dynamical analysis of thermodynamical observables}
538: \medskip
539: The basic thermodynamical phenomenology of a second-order phase transition is
540: characterized
541: by the existence of equilibrium
542: configurations that make the order parameter bifurcating away from zero
543: at some critical temperature $T_{c}$ and by a divergence of
544: the specific heat $c_{V}(T)$ at
545: the same $T_{c}$.
546: Therefore, this is the obvious starting point for the Hamiltonian dynamical
547: approach.
548:
549: \medskip
550: \subsubsection{Order parameter}
551: \medskip
552: Below a critical value of the temperature,
553: the symmetry-breaking in a system invariant under the action of the
554: $O(2)$ group, appears
555: as the selection
556: -- by the average magnetization vector of Eq. (\ref{order_par})--
557: of a preferred direction
558: among all the possible, energetically equivalent choices.
559: By increasing the lattice dimension, the symmetry breaking is therefore
560: characterized by a sort of simultaneous "freezing"
561: of the direction of the order parameter ${\bf M}$ and of the
562: convergence of its modulus to a non-zero value.
563:
564: Figure \ref{mag3d.e2} shows that in the $3d$ lattice,
565: at $T < 2$, i.e. in the broken-symmetry phase (as we shall see in the
566: following),
567: the dynamical simulations yield a thinner spread of the longitudinal
568: fluctuations by increasing $N$ -- that is, $| {\bf M} |$
569: oscillates by exhibiting a trend
570: to converge to a
571: non-zero value -- and that the transverse
572: fluctuations damp, ``fixing'' the direction of the oscillations.
573: This direction depends on the initial conditions.
574:
575: Moreover, the dynamical analysis provides us with a better detail than a
576: simple distinction between regular and chaotic dynamics.
577: In fact, it is possible to
578: distinguish between three different dynamical regimes
579: (Fig. \ref{fig.1.spin3d.9}).
580:
581: At low temperatures, up to $T \simeq 0.8$, one observes the
582: persistency of the initial direction and of an equilibrium value of the
583: modulus $| {\bf M} |$ close to one.
584:
585: At $0.8 < T < 2.2$, one observes transverse
586: oscillations, whose amplitude increases with temperature.
587:
588: At $T > 2.2$,
589: the order parameter exhibits the features typical of an unbroken symmetry
590: phase.
591: In fact, it displays fluctuations peaked at zero, whose
592: dispersion decreases by increasing the temperature (bottom of
593: Fig. \ref{fig.1.spin3d.9}) and,
594: at a given temperature, by increasing the lattice volume (Fig.
595: \ref{3d.fig.spin.altat}a,b).
596:
597: We can give an estimate of the order parameter by evaluating the average
598: of the modulus $\langle | {\bf M}(t)|\rangle
599: = \rho(T)$. At $T < 2.2$, the $N$-dependence is given mainly by
600: the rotation of the vector, while the longitudinal oscillations are moderate,
601: as shown in Fig. \ref{parord.3d.fig}. At temperatures above $T \simeq2.2$,
602: we observe the squeezing of $\rho(T)$ to a small value.
603:
604: The existence of a second order phase transition can be recognized
605: by comparing the temperature behavior and the $N$-dependence of
606: the thermodynamic observables computed for
607: the $2d$ and the $3d$ models.
608: Both systems exhibit
609: the rotation of the magnetization
610: vector and small fluctuations of its modulus when they are considered on small
611: lattices.
612: In the $2d$ model the average modulus of the order parameter is theoretically
613: expected to vanish logarithmically with $N$, what seems qualitatively
614: compatible with the weak $N$ dependence shown in
615: Fig. \ref{fig.spin2d.t0.74}, whereas in the $3d$ model we observe a stability
616: with $N$ of $\langle| {\bf M} |\rangle$, suggesting the convergence
617: to a non-zero value of the order parameter also in the limit
618: $N\rightarrow\infty$, as shown in Fig. \ref{mag3d.e2}.
619:
620: $T \simeq2.2$ is an approximate value of the critical temperature $T_c$ of the
621: second-order phase transition. This value will be refined in the following
622: Subsection. No finite-size scaling analysis has been performed for two
623: different reasons: {\it i)} our main concern is a qualitative phenomenological
624: analysis of the Hamiltonian dynamics of phase transitions rather than a
625: very accurate quantitative analysis, {\it ii)} finite-size effects are much
626: weaker in the microcanonical ensemble than in the canonical ensemble
627: \cite{Gross1}.
628:
629: \medskip
630: \subsubsection{Specific heat}
631: \medskip
632: As in the $2d$ model, numerical simulations of the Hamiltonian dynamics
633: have been performed with both Eqs.(\ref{specheat}) and (\ref{cvmicro}).
634: The outcomes
635: show a cusplike pattern of the specific heat, whose peak
636: makes possible a better determination of
637: the critical temperature. By increasing the lattice
638: dimension up to $N=15\times 15\times 15$,
639: the cusp becomes more pronounced,
640: at variance with the case of the $2d$ model.
641: Fig. \ref{calspec.3d.fig} shows that this occurs
642: at the temperature $T_{c}\simeq2.17$.
643:
644: \medskip
645: \subsubsection{Vorticity}
646: \medskip
647: The definition of the vorticity in the $3d$ case is not a simple extension
648: of the $2d$ case.
649: Vortices are always defined on a plane and if all the ``spins'' could freely
650: move
651: in the three-dimensional space, the concept of vortices would be meaningless.
652: For the $3d$ planar (anisotropic) model considered here, vortices can be
653: defined and studied on two-dimensional
654: subspaces of the lattice. The variables $q_{i,j,k}$ do not contain any
655: information about the position of the plane where the reference direction
656: to measure the angles $q_{i,j,k}$ is assigned.
657: Dynamics is completely independent of this choice, which has no effect on the
658: Hamiltonian. Moreover, as the Hamiltonian is symmetric with respect to the
659: lattice axes, the three coordinate-planes are equivalent. This
660: equivalence implies that vortices can contemporarily exist on three
661: orthogonal planes. Though the usual pictorial representation of a vortex
662: can hardly be maintained, its mathematical definition is the same as in the
663: $2d$ lattice case. Hence three vorticity functions exist and their
664: average values - at a given temperature - should not differ, what is actually
665: confirmed by numerical simulations.
666:
667: The vorticity function vs. temperature is plotted in
668: Fig. \ref{vort.fig3d}.
669: On a lattice of $10\times 10\times 10$ spins, the first vortex is
670: observed at $T \simeq0.8$.
671: The growth of the average density of vortices is very
672: fast up to the critical temperature, above which the saturation is reached.
673:
674: \medskip
675: \subsection{Lyapunov exponents and symmetry-breaking phase transition}
676: \medskip
677: A quantitative analysis of the dynamical chaoticity is provided by the
678: temperature dependence
679: of the largest Lyapunov exponent.
680:
681: Figure \ref{lyap.3d.fig} shows the results of this computation.
682: At low temperatures, in the limit of quasi-harmonic oscillators, the scaling
683: law is again found to be
684: $\lambda_{1}(T)\sim T^{2}$ and, at high
685: temperatures, the scaling law is again $\lambda_{1}(T)\sim T^{-1/6}$,
686: as in the $2d$ case.
687: In the temperature range intermediate between $T \simeq0.8$ and
688: $T_c \simeq2.17$, there is a linear growth of $\lambda_1(T)$.
689: At the critical temperature, the Lyapunov exponent exhibits an angular
690: point. This makes a remarkable difference between this
691: system undergoing a second order phase transition and
692: its $2d$ version, undergoing a Kosterlitz-
693: Thouless transition. In fact, the analysis of the $2d$ model has
694: shown a mild transition between the different regimes of $\lambda_{1}(T)$
695: (inset of Fig. \ref{vort.fig3d}),
696: whereas in the $3d$ model this transition
697: is sharper (inset of Fig. \ref{lyap.3d.fig}).
698:
699: We have also computed the temperature dependence of the largest Lyapunov
700: exponent of Markovian random processes which replace the true dynamics on the
701: energy surfaces $\Sigma_E$ (see Appendix).
702: The results are shown in Fig. \ref{randyn.3d.fig}.
703: The dynamics is considered
704: strongly chaotic in the temperature range where the patterns
705: $\lambda_1(T)$ are the same for both random and
706: differentiable dynamics, i.e. when differentiable dynamics mimics,
707: to some extent,
708: a random process.
709: The dynamics is considered weakly chaotic when the value
710: $\lambda_1$ resulting from
711: random dynamics is larger than the value $\lambda_1$
712: resulting from differentiable
713: dynamics.
714: The transition from weak to strong chaos is quite abrupt.
715: Figure \ref{randyn.3d.fig} shows that
716: the pattern of the largest Lyapunov exponent
717: computed by means of the random dynamics
718: reproduces that of the true Lyapunov exponent
719: at temperatures $T \geq T_c$. This means that the setting in
720: of strong thermodynamical disorder corresponds to the setting in of strong
721: dynamical chaos.
722: The ``window'' of strong chaoticity starts at
723: $T_c$ and ends at $T \sim 10$.
724: The existence of a second transition from strong to weak chaos is
725: due to the existence, for $T\rightarrow\infty$,
726: of the second integrable limit (of free
727: rotators), whence chaos cannot remain strong at any $T>T_c$.
728:
729: \medskip
730: \section{Geometry of dynamics and phase transitions}
731: \medskip
732: Let us briefly recall that the geometrization of the dynamics
733: of $N$-degrees-of-freedom systems defined by a Lagrangian
734: ${\cal L} = K - V$, in which the kinetic energy is quadratic in the velocities:
735: $K=\frac{1}{2}a_{ij} \dot{q}^i\dot{q}^j~$, stems from the fact that
736: the natural motions are the extrema
737: of the Hamiltonian action functional ${\cal S}_H =
738: \int {\cal L} \, dt$,
739: or of the Maupertuis' action
740: ${\cal S}_M = 2 \int K\, dt$.
741: In fact, also the geodesics of Riemannian and pseudo-Riemannian
742: manifolds are the extrema of a functional, the arc-length
743: $\ell=\int ds$, with $ds^2={g_{ij}dq^i dq^j}$.
744: Hence, a suitable choice of the metric tensor allows for the
745: identification of the arc-length with either ${\cal S}_H$ or
746: ${\cal S}_M$, and of the geodesics with the natural motions of the
747: dynamical system. Starting from ${\cal S}_M$, the ``mechanical manifold''
748: is the accessible configuration space endowed with
749: the Jacobi metric \cite{Pettini}
750: \beq
751: (g_J)_{ij} = [E - V(q)]\,a_{ij}~~,
752: \label{jacobi_metric}
753: \eeq
754: where $V(q)$ is the potential energy and $E$ is the total energy.
755: A description of the extrema of Hamilton's
756: action ${\cal S}_H$ as geodesics of a ``mechanical manifold''
757: can be obtained using Eisenhart's metric
758: \cite{Eisenhart} on an enlarged configuration spacetime
759: ($\{q^0\equiv t,q^1,\ldots,q^N\}$
760: plus one real coordinate $q^{N+1}$), whose arc-length is
761: \begin{equation}
762: ds^2 = -2V(\{ q \}) (dq^0)^2 + a_{ij} dq^i dq^j + 2 dq^0
763: dq^{N+1}~~.
764: \label{ds2E}
765: \end{equation}
766: The manifold has a Lorentzian structure and the dynamical
767: trajectories are those geodesics satisfying the condition
768: $ds^2 = C dt^2$, where $C$ is a positive constant.
769: In the geometrical framework, the (in)stability
770: of the trajectories is the (in)stability
771: of the geodesics, and it is completely determined by the
772: curvature properties of the underlying manifold according to
773: the Jacobi equation \cite{Pettini,doCarmo}
774: \begin{equation}
775: \frac{\nabla^2 \xi^i}{ds^2} + R^i_{~jkm}\frac{dq^j}{ds} \xi^k
776: \frac{dq^m}{ds} = 0~~,
777: \label{eqJ}
778: \end{equation}
779: whose solution $\xi$, usually called Jacobi or geodesic variation field,
780: locally measures the distance between nearby geodesics;
781: $\nabla/ds$ stands for the covariant derivative
782: along a geodesic and $R^i_{~jkm}$ are the components of
783: the Riemann curvature tensor.
784: Using the Eisenhart metric (\ref{ds2E}),
785: the relevant part of the Jacobi equation
786: (\ref{eqJ}) is \cite{CCP}
787: \begin{equation}
788: \frac{d^2 \xi^i}{dt^2} + R^i_{~0k0}\xi^k = 0~~,~~~~i=1,\dots ,N
789: \label{eqdintang}
790: \end{equation}
791: where the only non-vanishing components of the curvature tensor are
792: $R_{0i0j}=\partial^2 V/\partial q_i \partial q_j $. Equation
793: (\ref{eqdintang}) is the tangent dynamics equation, which is commonly used to
794: measure Lyapunov exponents in standard Hamiltonian systems.
795: Having recognized its geometric origin,
796: it has been
797: devised in Ref.\cite{CCP} a geometric reasoning
798: to derive from Eq.(\ref{eqdintang})
799: an {\it effective} scalar stability equation that, {\it independently} of the
800: knowledge of dynamical trajectories, provides an average measure of their
801: degree of instability.
802: An intermediate step in this derivation yields
803: \beq
804: \frac{d^2 \xi^j}{dt^2} + k_R(t) \xi^j + \delta K^{(2)}{(t)} \xi^j = 0~~,
805: \label{sectional}
806: \eeq
807: where $k_R=K_R/N$ is the Ricci curvature along a geodesic defined as
808: $K_R = \frac{1}{v^2} R_{ij} {\dot{q}^i}{\dot{q}^j}$,
809: with $v^2 = {\dot{q}^i}{\dot{q}_i}$ and
810: $R_{ij} = R^{k}_{~ikj}$, and $\delta K^{(2)}$ is the local
811: deviation of sectional
812: curvature from its average value \cite{CCP}.
813: The sectional curvature is defined as $K^{(2)} = R_{~ijkl} \xi^i \dot{q}^j
814: \xi^k \dot{q}^l/ \parallel\xi\parallel^2 \parallel\dot{q}\parallel^2$.
815:
816: Two simplifying assumptions are made:
817: $(i)$ the ambient manifold is {\em almost isotropic}, i.e.
818: the components of the curvature tensor --- that for an isotropic manifold
819: (i.e. of constant curvature)
820: are $R_{ijkm}=k_0(g_{ik} g_{jm} - g_{im} g_{jk})$, $k_0=const$
821: -- can be approximated by
822: $R_{ijkm} \approx k(t)
823: (g_{ik} g_{jm} - g_{im} g_{jk})$
824: along a generic geodesic $\gamma(t)$; $(ii)$
825: in the large $N$ limit, the ``effective curvature''
826: $k(t)$ can be modeled by a gaussian and $\delta$-correlated stochastic
827: process.
828: Hence, one derives
829: an effective
830: stability equation, independent of the dynamics and in the
831: form of a stochastic oscillator equation
832: \cite{CCP},
833: \begin{equation}
834: \frac{d^2\psi}{dt^2} + [k_0 + \sigma_k \eta(t)] \, \psi = 0~~,
835: \label{eqpsi}
836: \end{equation}
837: where $\psi^2 \propto |\xi|^2$.
838: The mean $k_0$ and variance $\sigma_k$ of $k(t)$
839: are given by
840: $k_0 = {\langle K_R \rangle}/{N}$
841: and $\sigma^2_k = {\langle (K_R - \langle K_R \rangle)^2 \rangle}/{N}$,
842: respectively, and the averages $\langle\cdot\rangle$ are geometric averages,
843: i.e. integrals computed on the mechanical manifold. These averages are
844: directly related with microcanonical averages, as it will be seen at the end
845: of Section V.
846: $\eta(t)$ is a gaussian $\delta$-correlated random process of
847: zero mean and unit variance.
848:
849: The main source of instability of the solutions of Eq.(\ref{eqpsi}),
850: and therefore the main source of Hamiltonian chaos, is
851: parametric resonance, which is
852: activated by the variations of the Ricci curvature
853: along the geodesics and which takes place also on positively curved manifolds
854: \cite{CerrutiPettini}. The dynamical instability can be enhanced
855: if the geodesics encounter regions
856: of negative sectional curvatures, such that $k_R + \delta K^{(2)}
857: < 0$, as it is evident from Eq. (\ref{sectional}).
858:
859: In the case of Eisenhart metric,
860: it is
861: $K_R\equiv \Delta V = \sum_{i=1}^N ({\partial^2 V}/{\partial q_i^2})$
862: and $K^{(2)} = R_{~0i0j} \xi^i \xi^j/ \parallel \xi \parallel^2 \equiv
863: (\partial^2V/\partial q^i\partial q^j)\xi^i\xi^j/\Vert\xi\Vert^2$.
864: The exponential growth rate $\lambda$ of the
865: quantity $\psi^2+\dot\psi^2$ of the solutions of Eq. (\ref{eqpsi}),
866: is therefore an estimate of the largest Lyapunov exponent that can be
867: analytically computed. The final result reads \cite{CCP}
868: \begin{equation}
869: \lambda = \frac{\Lambda}{2} - \frac{2 k_0}{3 \Lambda}\,,~~
870: \Lambda = \left(2\sigma_k^2 \tau +
871: \sqrt{\frac{64 k_0^3}{27} + 4\sigma_k^4 \tau^2}~\right)^\frac{1}{3}~,
872: \label{lambda}
873: \end{equation}
874: where
875: $\tau = \pi\sqrt{k_0}/(2\sqrt{k_0(k_0 + \sigma_k)}
876: +\pi\sigma_k )$;
877: in the limit $\sigma_k/k_0 \ll 1$ one finds $\lambda \propto \sigma_k^2~$.
878:
879: \subsection{Signatures of phase transitions from geometrization of
880: dynamics}
881:
882: In the geometric picture, chaos is mainly originated by the parametric
883: instability activated by the fluctuating curvature felt by geodesics,
884: i.e. the fluctuations of the (effective)
885: curvature are the source of the instability of the dynamics.
886: On the other hand, as it is witnessed by the derivation of Eq. (\ref{eqpsi})
887: and by the equation itself, a statistical-mechanical-like treatment of the
888: average degree of chaoticity is made possible by the geometrization of the
889: dynamics. The relevant curvature properties of the mechanical manifolds
890: are computed, at the formal level, as statistical averages, like other
891: thermodynamic observables. Thus, we can expect that some precise relationship
892: may exist between geometric, dynamic and thermodynamic quantities.
893: Moreover, this implies that phase transitions should correspond to peculiar
894: effects in the geometric observables.
895:
896: In the particular case of the $2d$ XY model,
897: the microcanonical average kinetic energy $\langle K \rangle$
898: and the average Ricci curvature
899: $\langle K_R \rangle$ computed
900: with the Eisenhart metric are linked by the equation
901: \beq
902: K_{R}=
903: \left \langle \sum _{i,j=1}^{N}\frac{\partial^{2} V}{\partial^2 q_{i,j}}
904: \right \rangle = 2J
905: \sum_{i,j=1}^{N}
906: \left\langle\cos(q_{i+1,j}-q_{i,j})
907: +\cos(q_{i,j+1}-q_{i,j}) \right\rangle =2(J-\langle V \rangle )~~,
908: \label{kinRicci}
909: \eeq
910: so that
911: \beq
912: H=N\epsilon=\langle K \rangle + \langle V \rangle \mapsto\frac{ \langle K
913: \rangle }{N}=\epsilon-2J+\frac{1}{2}\frac{\langle K_{R} \rangle}{N}~~.
914: \label{cRicci.Uinterna}
915: \eeq
916: Being the temperature defined as $T=2\langle K \rangle /N$
917: (with $k_{B}=1$) and being $d=1$ (because each spin has only one
918: rotational degree of freedom), from Eq.(\ref{specheat}) it follows that
919:
920: \beq
921: c_{V}=\frac{1}{2}\left(1-\frac{1}{2}\frac{\sigma^{2}_k/N}{T^2}\right)^{-1}.
922: \label{fluttKr.calspec}
923: \eeq
924:
925: In the special case of these XY systems,
926: it is possible to link the specific heat and the Ricci curvature
927: by inserting Eq.(\ref{cRicci.Uinterna}) into the usual expression for the
928: specific heat at constant volume. Thus, one obtains the equation
929: \beq
930: c_{V}=-\frac{1}{2N}\frac{\partial \langle K_{R} \rangle(T)}{\partial T}~.
931: \label{calspec,Ricci}
932: \eeq
933: The appearance of a peak in the specific heat function at the critical
934: temperature
935: has to correspond to a suitable temperature dependence of the Ricci curvature.
936:
937: In the $3d$ model, the
938: potential energy and the Ricci curvature
939: are proportional, according to:
940: $\frac{1}{N} \langle V \rangle = 3 - \frac{1}{2 N} \langle K_{R} \rangle~$.
941:
942: Another interesting point is the relation between a geometric observable
943: and the vorticity function in both models.
944: As already seen in previous sections, the
945: vorticity function is a useful signature of the dynamical chaoticity
946: of the system. From the geometrical point of view,
947: the enhancement of the instability of the dynamics
948: with respect to the parametric instability due to curvature fluctuations,
949: is linked
950: to the probability of obtaining negative sectional curvatures
951: along the geodesics (as discussed for $1d$ XY model in Ref.\cite{CCP}).
952: In fact, when vortices are present in the system,
953: there will surely be two neighbouring spins with an orientation difference
954: greater than $\pi/2$, such that, if $i,j$ and $i+1,j$ are their coordinates
955: on the lattice, it follows that
956: \beq
957: q_{i+1,j}-q_{i,j} > \frac{\pi}{2} \rightarrow \cos(q_{i+1,j}-q_{i,j})<0~.
958: \label{coord}
959: \eeq
960: The sectional curvature relative to the plane defined
961: by the velocity ${\bf v}$ along
962: a geodesic and a generic vector ${\bf \xi}\perp {\bf v}$ is
963: \beq
964: K^{(2)}= \sum_{i,j,k,l=1}^{N} \frac{\partial^2V}
965: {\partial q_{i,j}\partial q_{k,l}}
966: \frac{\xi^{i,j}\xi^{k,l}}{\|{\bf \xi}\|^{2}}~~.
967: \label{kappa2}
968: \eeq
969: For the $2d$ XY model, it is
970: \beq
971: K^{(2)}= \frac{J}{\|{\bf \xi}\|^{2}}\sum_{i,j=1}^{N}\{\cos(q_{i+1,j}-q_{i,j})
972: [\xi^{i+1,j}-\xi^{i,j}]^{2} +\cos(q_{i,j+1}-q_{i,j})[\xi^{i,j+1}-\xi^{i,j}]^{2}
973: \}~.
974: \label{curvsezXY}
975: \eeq
976: Thus, a large probability of
977: having a negative value of the cosine of the difference among the directions
978: of two close spins corresponds to a larger probability of obtaining
979: negative values of the sectional curvatures along the geodesics; here for $\xi$
980: the geodesic separation vector of Eq.(\ref{eqdintang}) is chosen.
981:
982: In the $3d$ model, the sectional curvature relative to
983: the plane defined by the velocity ${\bf v}$ and
984: a generic vector ${\bf \xi}\perp {\bf v}$ is
985: \begin{eqnarray}
986: K^{(2)}&=& \frac{J}{\|\xi\|^{2}}\sum_{i,j,k=1}^{N} \{
987: \cos(q_{i+1,j,k}-q_{i,j,k})[\xi^{i+1,j,k}-\xi^{i,j,k}]^{2}+
988: \nonumber\\
989: &+&\cos(q_{i,j+1,k}-q_{i,j,k})[\xi^{i,j+1,k}-\xi^{i,j,k}]^{2}+
990: \cos(q_{i,j,k+1}-q_{i,j,k})[\xi^{i,j,k+1}-\xi^{i,j,k}]^{2}\}
991: \label{curvsez3d}
992: \end{eqnarray}
993: and again the probability of finding negative values of $K^{(2)}$
994: along a trajectory is limited to the probability of finding vortices.
995:
996: The mean values of the geometric quantities entering Eq.(\ref{eqpsi})
997: can be numerically computed by means of Monte Carlo simulations
998: or by means of time averages along the
999: dynamical trajectories. In fact, due to the lack of an explicit expression
1000: for the canonical partition function of the system,
1001: these averages are not analytically computable.
1002: For sufficiently high
1003: temperatures, the potential energy becomes
1004: negligible with respect to the kinetic energy, and each spin is free to move
1005: independently from the others. Thus, in the limit of high temperatures,
1006: one can estimate the configurational partition function
1007: $Z_C = \int_{-\pi}^{\pi} \prod_{\bf i} dq_{\bf i} e^{-\beta V(q)}$
1008: by means of the expression
1009: \begin{eqnarray}
1010: Z_C& =& e^{-2 \beta JN}
1011: \int_{-\pi}^{\pi} \prod_{i,j=1}^{N}
1012: dq_{i,j}\exp\{\beta J \sum_{i,j=1}^{N}[
1013: \cos(q_{i+1,j}-q_{i,j})+\cos(q_{i,j+1}-q_{i,j})]\}\nonumber\\
1014: &\sim& e^{-2 \beta JN}\int_{-\pi}^{\pi} \prod_{i,j=1}^{N}du_{i,j}dv_{i,j}
1015: \exp\{\beta J \sum_{i,j=1}^{N}[
1016: \cos(u_{i,j})+\cos(v_{i,j})]\}
1017: \label{partition}
1018: \end{eqnarray}
1019: after the introduction of $u_{i,j}= q_{i+1,j}-q_{i,j}$ and
1020: $v_{i,j}= q_{i,j+1}-q_{i,j}$ as independent variables.
1021: In this way, some analytical estimates of the average Ricci
1022: curvature $k_{0}(T)$ and of its r.m.s. fluctuations
1023: $\sigma^2_k(T)$ have been obtained for the $2d$ model
1024: (Fig. \ref {fig.Ricci2d}).
1025: For temperatures above the temperature of the Kosterlitz-Thouless
1026: transition, these estimates
1027: are in agreement
1028: with the numerical computations on a $N=10\times 10$ lattice.
1029: It is confirmed that Hamiltonian dynamical simulations,
1030: already on rather small lattices,
1031: are useful to predict, with a good approximation, the thermodynamic limit
1032: behavior of relevant observables.
1033: Moreover, the good quality of the high temperature estimate gives
1034: a further information: at the transition temperature,
1035: the correlations among the different degrees of freedom are destroyed,
1036: confirming the
1037: strong chaoticity of the dynamics.
1038:
1039: The same high temperature estimates of $k_0(T)$ and $\sigma^2_{k}(T)$
1040: have been performed for the $3d$ system.
1041: In Fig. \ref{keflutt.3dfig},
1042: the numerical determination of $\sigma^2_{k}(T)$
1043: shows the appearance of a very pronounced peak at the phase transition point
1044: which is not predicted by the analytic estimate,
1045: whereas the average Ricci curvature $k_0(T)$ is in agreement
1046: with the analytic values of the high temperature estimate, computed by spin
1047: decoupling, above the critical temperature, as in the $2d$ model.
1048:
1049: \medskip
1050: \subsection{Geometric observables and Lyapunov exponents}
1051: \medskip
1052: We have seen that the largest
1053: Lyapunov exponent is sensitive to the phase transition and at the same time
1054: we know that
1055: it is also related to the average curvature properties of the ``mechanical
1056: manifolds''.
1057: Thus, the geometric observables $k_0(T)$ and $\sigma^2_k(T)$
1058: above considered can be used to
1059: estimate the Lyapunov exponents, as well as to detect the phase transition.
1060:
1061: In principle, by means of Eq.(\ref{lambda}), one can evaluate
1062: the largest Lyapunov exponent without any need of dynamics,
1063: but simply using
1064: global geometric quantities of the manifold
1065: associated to the physical system.
1066: For $2d$ and $3d$ XY models, fully analytic computations are possible only
1067: in the limiting cases of high and low temperatures. Microcanonical averages
1068: of $k_0$ and $\sigma^2_k$ at arbitrary $T$ have been numerically computed
1069: through time averages. We can call this hybrid method semi-analytic.
1070:
1071: In Fig. \ref{prev.Lyap.2d}, the results of the semi-analytic prediction of the
1072: Lyapunov exponents for the $2d$ model are plotted vs. temperature and
1073: compared with the numerical outcomes of the tangent dynamics.
1074: As one can see, the prediction formulated on the basis of Eq.(\ref{lambda})
1075: underestimates the numerical values given by the tangent
1076: dynamics.
1077: The semi-analytic prediction can be improved
1078: by observing that the replacement of the sectional curvature fluctuation
1079: $\delta K^{(2)}$ in Eq.(\ref{sectional}) with a fraction
1080: of the Ricci curvature [which underlies the derivation of Eq.(\ref{eqpsi})]
1081: underestimates the frequency of occurrence of negative sectional curvatures,
1082: which was already the case of the $1d$ XY model \cite{CCP}.
1083: The correction procedure can be implemented
1084: by evaluating the probability $P(T)$ of obtaining a negative value of the
1085: sectional curvature along a generic trajectory and then by operating the
1086: substitution
1087: \beq
1088: K_{R}(T)\rightarrow \frac{K_{R}(T)}{1+P(T) \alpha}~.
1089: \label{kappafrac}
1090: \eeq
1091: The parameter $\alpha$ is a free parameter to be empirically estimated.
1092: Its value ranges from $100$ to $200$, without appreciable differences in the
1093: final result.
1094: It resumes
1095: the non trivial information about the more pronounced tendency
1096: of the trajectories towards negative
1097: sectional curvatures with respect to the predictions of the geometric model
1098: describing the chaoticity of the dynamics.
1099:
1100: The probability $P(T)$ is estimated through the occurrence along
1101: a trajectory of negative values
1102: of the sum of the coefficients that appear in the definition
1103: of $K^{(2)}$ [Eqs.(\ref{curvsezXY}) and (\ref{curvsez3d})]
1104: \beq
1105: P(T)\sim \frac{\int_{-\pi}^{\pi}
1106: \Theta (-\cos(q_{k+1,l}-q_{k,l})-\cos(q_{k,l+1}-q_{k,l}))
1107: \exp[-\beta V({\bf q})] \prod_{k,l=1}^{N}dq_{k,l}}
1108: {\int_{-\pi}^{\pi}\exp[-\beta V({\bf
1109: q})] \prod_{k,l=1}^{N}dq_{k,l}}~~,
1110: \label{2d.pro.cos.neg}
1111: \eeq
1112: averaged over all the sites
1113: $\forall k,l\in (1,\ldots,N)$; $\Theta$ is the step function.
1114:
1115: Alternatively, owing to the already remarked relation between vorticity
1116: and sectional curvature $K^{(2)}$,
1117: $P(T)$ can be replaced by the average density of
1118: vortices
1119: \beq
1120: K_{R}(T)\rightarrow \frac{K_{R}(T)}{1+\overline{\alpha} {\cal V}(T)}~,
1121: \label{correction}
1122: \eeq
1123: where $\overline{\alpha}$ a free parameter.
1124: Actually,
1125: in the $2d$ model, the two corrections, one given by
1126: Eq.(\ref{kappafrac}) with $P(T)$ of Eq. (\ref{2d.pro.cos.neg}), the other
1127: given by Eq.(\ref{correction}) with the vorticity function in place of $P(T)$,
1128: convey the same information.
1129: The semi-analytic predictions of $\lambda_1(T)$ with correction
1130: are reported in Fig. \ref{prev.Lyap.2d}.
1131:
1132: In the limits of high and low temperatures,
1133: $\lambda_1(T)$ can be given the analytic forms
1134: $\lambda_{1}(T)\sim T^{-1/6}$ at high
1135: temperature, and
1136: $\lambda_{1}(T)\sim T^{2}$ at low temperature.
1137: In the former case, the high temperature approximation (\ref{partition})
1138: is used, and in the latter case the quasi-harmonic oscillators approximation
1139: is done. The deviation of $\lambda_1(T)$ from the quasi-harmonic scaling,
1140: starting at $T \simeq0.6$ and already observed to correspond to the
1141: appearance of vortices, finds here a simple explanation through the geometry
1142: of dynamics: vortices are associated with negative sectional curvatures,
1143: enhancing chaos.
1144:
1145: By increasing the spatial dimension of the system, it becomes more and more
1146: difficult to accurately estimate the probability of obtaining negative
1147: sectional curvatures.
1148: The assumption that the occurrence of negative values
1149: of the cosine of the difference between the directions of two nearby spins
1150: is nearly equal to $P(T)$, is less effective in the $3d$
1151: model than in the $2d$ one.
1152: Again, the vorticity function can be assumed as an estimate of $P(T)$
1153: [Eq. (\ref{correction})].
1154: The quality of the results has a weak dependence upon
1155: the parameter $\alpha$.
1156: The correction remains good, with $\alpha$ belonging to a broad interval
1157: of values ($100 \div 200$).
1158: In the limits of high and low temperatures, the model predicts correctly
1159: the same scaling laws of the $2d$ system.
1160:
1161: In Fig. \ref{prev.Lyap.3d} the semi-analytic
1162: predictions for the Lyapunov exponents,
1163: with and without correction, are
1164: plotted vs. temperature together with the numerical results
1165: of the tangent dynamics.
1166: It is noticeable that the prediction of Eq. (\ref{lambda}) is able to give
1167: the correct asymptotic behavior of the Lyapunov exponents
1168: also at low temperatures, the most difficult part
1169: to obtain by means of dynamical simulations.
1170:
1171: \medskip
1172: \subsection{A topological hypothesis}
1173: \medskip
1174:
1175: We have seen in Fig. \ref{keflutt.3dfig} that a sharp peak of the
1176: Ricci-curvature fluctuations
1177: $\sigma_\kappa^2(T)$ is found for the $3d$ model in correspondence of the
1178: second order phase transition, whereas, for the $2d$ model,
1179: $\sigma_\kappa^2(T)$ appears regular and in agreement with the theoretically
1180: predicted smooth pattern.
1181: On the basis of heuristic arguments, in Refs.\cite{CCCP,CCCPPG} we suggested
1182: that the peak of $\sigma_\kappa^2$ observed for the $3d$ XY model, as well as
1183: for $2d$ and $3d$ scalar and vector lattice $\varphi^4$ models, might originate
1184: in some change of the {\it topology} of the mechanical manifolds. In fact,
1185: in abstract mathematical models, consisting of families of surfaces undergoing
1186: a topology change -- i.e. a loss of diffeomorphicity among them --
1187: at some critical value of
1188: a parameter labelling the members of the family, we have actually observed the
1189: appearance of cusps of $\sigma_K^2$ at the
1190: transition point between two subfamilies of surfaces of different topology,
1191: $K$ being the Gauss curvature.
1192:
1193: Actually, for the mean-field XY model, where both $\sigma_\kappa^2(T)$ and
1194: $\lambda_1(T)$ have theoretically been shown to loose analyticity at the phase
1195: transition point, a direct evidence of a ``special'' change of the topology
1196: of equipotential hypersurfaces of configuration space has been given
1197: \cite{cegdcp}. Other indirect and direct evidences of the actual involvement
1198: of topology in the deep origin of phase transitions have been recently
1199: given
1200: \cite{fps1,fps2} for the lattice $\varphi^4$ model.
1201:
1202: In the following Section we consider the extension
1203: of this
1204: topological point of view about phase transitions
1205: from equipotential hypersurfaces of configuration space
1206: to constant energy hypersurfaces of phase space.
1207:
1208: \medskip
1209: \section{Phase space geometry and thermodynamics.}
1210: \medskip
1211:
1212: In the preceding Section we have used some elements of intrinsic differential
1213: geometry of submanifolds of configuration space to describe the average
1214: degree of dynamical instability (measured by the largest Lyapunov exponent).
1215: In the present Section we are interested in the relationship
1216: between the extrinsic geometry of the constant energy
1217: hypersurfaces $\Sigma_E$ and thermodynamics.
1218:
1219: Hereafter, phase space is considered as an even-dimensional subset $\Gamma$
1220: of ${\Bbb R}^{2N}$
1221: and the hypersurfaces $\Sigma_E=\{(p_1,\dots ,p_N,q_1,\dots ,q_N)\in{\Bbb R}
1222: \vert H(p_1,\dots ,p_N,q_1,\dots ,q_N)=E\}$ are manifolds that can be equipped
1223: with the standard Riemannian metric induced from ${\Bbb R}^{2N}$. If, for
1224: example, a surface is parametrically defined through the equations
1225: $x^i=x^i(z^1,\dots ,z^k)$, $i=1,\dots ,2N$, then the metric $g_{ij}$
1226: {\it induced} on the surface is given by $g_{ij}(z^1,\dots ,z^k)=
1227: \sum_{n=1}^{2N}
1228: \frac{\partial x^n}{\partial z^i} \frac{\partial x^n}{\partial z^j}$.
1229: The geodesic flow associated with the metric induced on $\Sigma_E$ from
1230: ${\Bbb R}^{2N}$ has nothing to do with the
1231: Hamiltonian flow that belongs to $\Sigma_E$. Nevertheless, it exists an
1232: intrinsic
1233: Riemannian metric $g_S$ of phase space $\Gamma$ such that the geodesic
1234: flow of $g_S$, restricted to $\Sigma_E$, coincides with the Hamiltonian flow
1235: ($g_S$ is the so called Sasaki lift to the tangent bundle
1236: of configuration space of the Jacobi
1237: metric $g_J$ that we mentioned in a preceding Section).
1238:
1239: The link between extrinsic geometry of the $\Sigma_E$ and thermodynamics
1240: is estabilished through the microcanonical definition of entropy
1241: \beq
1242: S = k_B \log \int_{\Sigma_E} \frac{d\sigma}{\|{\nabla H}\|} ~,
1243: \label{entropy}
1244: \eeq
1245: where $d\sigma = \sqrt{det(g)} dx_1...dx_{2N-1}$ is the invariant volume
1246: element of $\Sigma_E \subset{\Bbb R}^{2N}$, ${g}$ is the metric induced from
1247: ${\Bbb R}^{2N}$ and $x_1...x_{2N-1}$ are the coordinates on $\Sigma_E$.
1248:
1249: Let us briefly recall some necessary definitions and concepts that are needed
1250: in the study of hypersurfaces of euclidean spaces.
1251:
1252: A standard way to investigate the geometry of an hypersurface $\Sigma^m$ is
1253: to study the way in which it curves around in ${\Bbb R}^{m+1}$: this is
1254: measured by the way the normal direction changes as we move from point to
1255: point on the surface. The rate of change
1256: of the normal direction ${\bf N}$ at a point $x\in\Sigma$ is described by
1257: the {\it shape operator} $L_x({\bf v}) = - \nabla_{\bf v}{\bf N}
1258: = - (\nabla N_1\cdot{\bf v},\dots ,\nabla N_{m+1}
1259: \cdot{\bf v})$, where
1260: ${\bf v}$ is a tangent vector at $x$ and $\nabla_{\bf v}$ is the directional
1261: derivative of the unit normal ${\bf N}$.
1262: As $L_x$ is an operator of the tangent space at $x$ into
1263: itself, there are $m$ independent eigenvalues \cite{thorpe} $\kappa_1(x),
1264: \dots,\kappa_m(x)$, which are called the principal curvatures of $\Sigma$ at
1265: $x$. Their product is the {\it Gauss-Kronecker curvature}:
1266: $K_G(x)=\prod_{i=1}^m\kappa_i(x)={\rm det}(L_x)$, and
1267: their sum is the so-called {\it mean curvature}:
1268: $M_1(x)=\frac{1}{m}\sum_{i=1}^m\kappa_i(x)$.
1269: The quadratic form $L_x({\bf v})\cdot{\bf v}$, associated with the shape
1270: operator at a point $x$, is called the second fundamental form of $\Sigma$ at
1271: $x$.
1272:
1273: It can be
1274: shown \cite{doCarmo} that the mean curvature of the energy hypersurfaces
1275: is given by
1276: \begin{equation}
1277: M_1(x)= -\frac{1}{2N-1}\nabla\cdot\left(\frac{\nabla H(x)}{\Vert\nabla H(x)
1278: \Vert}\right)~~,
1279: \label{emme1}
1280: \end{equation}
1281: where $\nabla H(x)/\Vert\nabla H(x)\Vert$ is the unit normal to
1282: $\Sigma_E$ at a given point $x=(p_1,\dots ,p_N,q_1,\dots ,q_N)$, and
1283: $\nabla =(\partial/\partial p_1,\dots,\partial/\partial q_N)$,
1284: whence the explicit expression
1285: \begin{eqnarray}
1286: (2N-1)\, M_1 = &-&\frac{1}{\Vert\nabla H\Vert}\left[ N + \sum_{\bf i}
1287: \left(\frac{\partial^2V}{\partial q_{\bf i}^2}\right)\right]\nonumber\\
1288: & +& \frac{1}{\Vert\nabla H\Vert^3}
1289: \left[\sum_{\bf i}p_{\bf i}^2 + \sum_{\bf i,j}\left(
1290: \frac{\partial^2V}{\partial q_{\bf i}\partial q_{\bf j}}\right)\left(
1291: \frac{\partial V}{\partial q_{\bf i}}\right)\left(\frac{\partial V}
1292: {\partial q_{\bf j}}\right)\right]~,
1293: \label{M1}
1294: \end{eqnarray}
1295: where ${\bf i,j}$ are multi-indices according to the number of spatial
1296: dimensions.
1297:
1298: The link between geometry and physics stems from the microcanonical definition
1299: of the temperature
1300: \begin{equation}
1301: \frac{1}{T}=\frac{\partial S}{\partial E}=\frac{1}{\Omega_{\nu}(E)}\frac
1302: {d\Omega_{\nu}(E)}{dE}~,
1303: \label{temperature}
1304: \end{equation}
1305: where we used Eq.(\ref{entropy}) with $k_B=1$, $\nu =2N-1$, and
1306: $\Omega_{\nu}(E)=\int_{\Sigma_E}d\sigma /\Vert\nabla H\Vert$.
1307: >From the formula \cite{laurence}
1308: \begin{equation}
1309: {d^k \over dE^k} \left( \int_{\Sigma_E} \alpha~d\sigma \right) (E')
1310: = \int_{\Sigma_{E'}} A^k(\alpha)\, d\sigma~~,
1311: \end{equation}
1312: where $\alpha$ is an integrable function and $A$ is the operator
1313: $ A(\alpha)= {\nabla \over \Vert \nabla H\Vert}
1314: \cdot \left( \alpha \cdot
1315: {\nabla H \over \Vert \nabla H\Vert} \right),
1316: $
1317: it is possible to work out the result
1318: \begin{equation}
1319: \frac{1}{T}=
1320: \frac{1}{\Omega_{\nu}(E)}\frac{d\Omega_{\nu}(E)}{dE} = \frac{1}{\Omega_{\nu}}
1321: \int_{\Sigma_E} \frac{d\sigma}{\Vert\nabla
1322: H\Vert} \left[ 2 \frac{M_1^\star}{\Vert\nabla H\Vert} - \frac{\triangle H}
1323: {\Vert\nabla H\Vert^2} \right]\simeq \frac{1}{\Omega_{\nu}}
1324: \int_{\Sigma_E} \frac{d\sigma}
1325: {\Vert\nabla H\Vert}\, \frac{M_1^\star}{\Vert\nabla H\Vert} ~~ ,
1326: \label{ballerotte}
1327: \end{equation}
1328: where $M_1^\star = \nabla (\nabla H/ \Vert\nabla H\Vert)$ is directly
1329: proportional to the mean curvature (\ref{emme1}). In the last term of
1330: Eq.(\ref{ballerotte}) we have neglected a contribution which vanishes
1331: as ${\cal O}(1/N)$. Eq. (\ref{ballerotte}) provides the
1332: fundamental link between extrinsic geometry and thermodynamics \cite{nota1}.
1333: In fact, the microcanonical average of ${M_1^\star}/{\Vert\nabla H\Vert}$,
1334: which is a quantity
1335: tightly related with the mean curvature of $\Sigma_E$, gives the inverse
1336: of the temperature, whence other important thermodynamic observables can be
1337: derived. For example, the constant volume specific heat
1338: \begin{equation}
1339: \frac{1}{C_V}=\frac{\partial T(E)}{\partial E}~,
1340: \label{cv}
1341: \end{equation}
1342: using Eq.(\ref{temperature}), yields
1343: \begin{equation}
1344: C_V=-\left(\frac{\partial S}{\partial E}\right)^2\,\left(\frac{\partial^2S}
1345: {\partial E^2}\right)^{-1}~~,
1346: \label{cv1}
1347: \end{equation}
1348: becoming at large $N$
1349: \begin{equation}
1350: C_V=-\left<\frac{{M_1^\star}}{{\Vert\nabla H\Vert}}\right>_{mc}^2 \left[
1351: \frac{1}{\Omega_{\nu}} \frac{d}{dE}
1352: \int_{\Sigma_E} \frac{d\sigma}
1353: {\Vert\nabla H\Vert}\,\left( \frac{M_1^\star}{\Vert\nabla H\Vert}+R(E)\right)
1354: - \left<\frac{{M_1^\star}}{{\Vert\nabla H\Vert}}\right>_{mc}^2\right]^{-1}~,
1355: \label{cv2}
1356: \end{equation}
1357: where the subscript $mc$ stands for microcanonical average, and
1358: $R(E)$ stands for the quantities of order ${\cal O}(1/N)$
1359: neglected in the last term of Eq.(\ref{temperature})
1360: (a-priori, its derivative can be non negligible and has to be taken into
1361: account).
1362: Eq. (\ref{cv2}) highlights a more elaborated link between geometry and
1363: thermodynamics: the specific heat depends upon the microcanonical average of
1364: ${M_1^\star}/{\Vert\nabla H\Vert}$ and upon the energy variation rate of the
1365: surface integral of this quantity.
1366:
1367: Remarkably, the relationship between curvature
1368: properties of the constant energy surfaces $\Sigma_E$ and thermodynamic
1369: observables given by Eqs.(\ref{temperature}) and (\ref{cv2}) can be extended
1370: to embrace also a deeper and very interesting relationship between
1371: thermodynamics and {\it topology} of the constant
1372: energy surfaces. Such a relationship can be discovered through a
1373: reasoning which, though approximate, is highly non-trivial, for it makes use
1374: of a deep theorem due to Chern and Lashof \cite{ChernLashof}.
1375: As $\Vert\nabla H\Vert =\{\sum_i p_i^2 + [\nabla_iV(q)]^2\}^{1/2}$ is a
1376: positive quantity increasing with the energy, we can write
1377: \begin{equation}
1378: \frac{1}{T}=
1379: \frac{1}{\Omega_{\nu}}\frac{d\Omega_{\nu}}{dE} \simeq\frac{1}{\Omega_{\nu}}
1380: \int_{\Sigma_E} \frac{d\sigma}
1381: {\Vert\nabla H\Vert}\, \frac{M_1^\star}{\Vert\nabla H\Vert}
1382: = D(E)\frac{1}{\Omega_{\nu}} \int_{\Sigma_E} d\sigma \, M_1~,
1383: \label{deformaz}
1384: \end{equation}
1385: where we have introduced the factor function $D(E)$ in order to extract the
1386: total mean curvature $\int_{\Sigma_E} d\sigma \, M_1$; $D(E)$ has been
1387: numerically found to be smooth and very close to $\langle {1}/{\Vert
1388: \nabla H\Vert^2} \rangle_{mc}$ (see Section \ref{5a} and Fig.
1389: \ref{D(E)}). Then,
1390: recalling the expression of a multinomial expansion
1391: \begin{equation}
1392: (x_1 + \cdots + x_{\nu})^{\nu} = \sum_{_{\{n_i\},\sum n_k =\nu}}
1393: \frac{\nu !}{n_1!\cdots n_{\nu}!} \cdot x_1^{n_1} \cdots x_{\nu}^{n_{\nu}}~~ ,
1394: \label{somma1}
1395: \end{equation}
1396: and identifying the $x_i$ with the principal curvatures $k_i$, one obtains
1397: \begin{equation}
1398: M_1^{\nu} = {\nu}! \prod_{i=1}^{\nu} k_i + R = {\nu}! K + R~~~,
1399: \end{equation}
1400: where $K=\prod_i k_i$ is the Gauss-Kronecker curvature,
1401: and $R$ is the sum (\ref{somma1}) without the term with the
1402: largest coefficient ($n_k=1, \ \forall k$).
1403: Using $\nu ! \simeq \nu^{\nu} e^{-\nu} \sqrt{2\pi\nu}$,
1404: \begin{equation}
1405: M_1^{\nu} \simeq \nu^{\nu} e^{-\nu} \sqrt{4\pi N} K + R
1406: \label{pallino}
1407: \end{equation}
1408: is obtained.
1409: The above mentioned theorem of Chern and Lashof
1410: states that
1411: \begin{equation}
1412: \int_{\Sigma_E}
1413: \vert K\vert\,d\sigma\geq Vol[{\Bbb S}_1^{\nu}]\sum_{i=0}^{\nu}
1414: b_i(\Sigma_E )~,
1415: \label{chern1}
1416: \end{equation}
1417: i.e. the total absolute Gauss-Kronecker curvature of a hypersurface is
1418: related with the sum of all its Betti numbers $b_i(\Sigma_E)$. The Betti
1419: numbers are {\it diffeomorphism invariants} of fundamental topological
1420: meaning \cite{Betti}, therefore their sum is also a topologic invariant.
1421: ${\Bbb S}_1^\nu$ is a hypersphere of unit radius. Combining Eqs.
1422: (\ref{pallino}) and (\ref{chern1}) and integrating on $\Sigma_E$, we obtain
1423: \begin{equation}
1424: \int_{\Sigma_E}\vert M_1^{\nu}\vert\, d \sigma \simeq
1425: \nu^{\nu} e^{-\nu} \sqrt{2\pi\nu}
1426: \int_{\Sigma_E}\vert K\vert d \sigma + \int_{\Sigma_E}\vert R\vert d \sigma
1427: \ge {\cal A}\,\sum_{i=0}^{\nu} b_i(\Sigma_E ) + {\cal R}(E)~~,
1428: \label{tria}
1429: \end{equation}
1430: with the shorthands
1431: ${\cal A}=\nu^{\nu} e^{-\nu} Vol(S_1^{\nu})$ and ${\cal R}=\int_{\Sigma_E}
1432: \vert R\vert d \sigma$.
1433:
1434: Now, with the aid of the inequality $\int\Vert f\Vert^{1/n}d\mu\geq\Vert\int f
1435: d\mu\Vert^{1/n}$, we can write
1436: \begin{equation}
1437: \int_{\Sigma_E}\vert M_1\vert\, d \sigma = \int_{\Sigma_E}\vert M_1^{\nu}\vert
1438: ^{1/\nu}\, d \sigma \geq
1439: \left\vert\int_{\Sigma_E}M_1^{\nu}\,d\sigma\right\vert^{1/\nu}~~.
1440: \label{chern2}
1441: \end{equation}
1442: If $M_1\geq 0$ everywhere on $\Sigma_E$, then
1443: $\left\vert\int_{\Sigma_E}M_1^{\nu}\,d\sigma\right\vert^{1/\nu} =
1444: \left(\int_{\Sigma_E}\vert M_1^{\nu}\vert\,d\sigma\right)^{1/\nu}$, whence, in
1445: the hypothesis that $M_1\geq 0$ is largely prevailing \cite{bassi-indici},
1446: $\left\vert\int_{\Sigma_E}M_1^{\nu}\,d\sigma\right\vert^{1/\nu} \sim
1447: \left(\int_{\Sigma_E}\vert M_1^{\nu}\vert\,d\sigma\right)^{1/\nu}$. Under the
1448: same assumption, $\int_{\Sigma_E}M_1d\sigma\sim\int_{\Sigma_E}\vert M_1\vert
1449: d\sigma$ and therefore
1450: \begin{equation}
1451: \int_{\Sigma_E} M_1\, d \sigma \sim \int_{\Sigma_E}\vert M_1^{\nu}\vert
1452: ^{1/\nu}\, d \sigma \geq
1453: \left\vert\int_{\Sigma_E}M_1^{\nu}\,d\sigma\right\vert^{1/\nu} \sim
1454: \left(\int_{\Sigma_E}\vert M_1^{\nu}\vert\,d\sigma\right)^{1/\nu}
1455: \geq \left[ {\cal A}\sum_{i=0}^{\nu} b_i(\Sigma_E ) + {\cal R}(E)
1456: \right]^{1/\nu}~~.
1457: \label{chern3}
1458: \end{equation}
1459: Finally,
1460: \begin{eqnarray}
1461: \frac{1}{T(E)} =
1462: \frac{1}{\Omega_{\nu}}\frac{d\Omega_{\nu}}{dE}& \simeq &
1463: \left<\frac{{M_1^\star}}{{\Vert\nabla H\Vert}}\right>_{mc} =\,
1464: \frac{1}{\Omega_{\nu}} \int_{\Sigma_E} \frac{d\sigma}
1465: {\Vert\nabla H\Vert}\, \frac{M_1^\star}{\Vert\nabla H\Vert}
1466: = D(E)\frac{1}{\Omega_{\nu}} \int_{\Sigma_E} d\sigma \, M_1\nonumber \\
1467: &\geq &
1468: \frac{D(E)}{\Omega_{\nu}}\left[ {\cal A}
1469: \sum_{i=0}^{\nu} b_i(\Sigma_E ) + {\cal R}(E) \right]^{1/\nu}~~.
1470: \label{thermtop}
1471: \end{eqnarray}
1472: Equation (\ref{thermtop}) has the remarkable property of relating
1473: the microcanonical definition of
1474: temperature of Eq.(\ref{deformaz}) with a {\it topologic invariant} of
1475: $\Sigma_E$.
1476: The Betti numbers can be exponentially large with $N$ [for example,
1477: in the case of $N$-tori ${\Bbb T}^N$, they are $b_k={N\choose k}$], so that
1478: the quantity $(\sum b_k)^{1/N}$ can converge, at arbitrarily large $N$, to a
1479: non-trivial limit (i.e. different from one).
1480: Thus, even though the energy dependence of ${\cal R}$ is
1481: unknown, the energy variation of $\sum b_i(\Sigma_E)$ must be mirrored -- at
1482: any arbitrary $N$ -- by the energy variation of the temperature.
1483: By considering Eq.(\ref{cv2}) in the light of Eq.(\ref{thermtop}), we can
1484: expect that
1485: some suitably abrupt and major change in the topology of the $\Sigma_E$ can
1486: reflect into the appearance of a peak of the specific heat,
1487: as a consequence of the associated
1488: energy dependence of $\sum b_k(\Sigma_E)$ and of its derivative with respect
1489: to $E$. In other words, we see that a link must exist between thermodynamical
1490: phase transitions and suitable topology changes of the constant energy
1491: submanifolds of the phase space of microscopic variables.
1492: The arguments given above, though in a still rough formulation, provide a
1493: first attempt to
1494: make a connection between the topological aspects of the {\it microcanonical}
1495: description of phase transitions and the already proposed {\it topological
1496: hypothesis} about topology changes in configuration space and phase transitions
1497: \cite{CCCP,CCCPPG,cegdcp,fps1,fps2}.
1498:
1499: Direct support to the {\it topological
1500: hypothesis} has been given by the analytic study of a mean-field XY model
1501: \cite{cegdcp} and by the numerical computation of the Euler characteristic
1502: $\chi$ of the equipotential hypersurfaces $\Sigma_v$ of the configuration
1503: space in a $2d$ lattice $\varphi^4$ model \cite{fps2}.
1504: The Euler characteristic is the alternate sum of all the Betti numbers of
1505: a manifold, so it is another topological invariant, but it identically
1506: vanishes for odd dimensional manifolds, like the $\Sigma_E$.
1507: In Ref.\cite{fps2},
1508: $\chi (\Sigma_v)$ neatly reveals the symmetry-breaking phase transition
1509: through a sudden change of
1510: its variation rate with the potential energy density $v$. A sudden
1511: ``second order'' variation of the topology of the $\Sigma_v$ appears in both
1512: Refs.\cite{cegdcp,fps2} as the requisite for the appearance of a phase
1513: transition. These results strenghten the arguments given in the present
1514: Section about the role of the topology of the constant energy hypersurfaces.
1515: In fact, the larger is $N$, the smaller are the relative fluctuations
1516: $\langle\delta^2V\rangle^{1/2}/\langle V\rangle$ and $\langle\delta^2K\rangle
1517: ^{1/2}/\langle K\rangle$ of the potential and kinetic energies respectively.
1518: At very large $N$, the product manifold
1519: $\Sigma^{N-1}_v\times{\Bbb S}_t^{N-1}$, with $v\equiv\langle V\rangle$ and
1520: $t\equiv\langle K\rangle$, $v+t=E$, is a good model manifold to represent
1521: the part of $\Sigma_E$ that is overwhelmingly sampled by the dynamics and that
1522: therefore constitutes the effective support of the microcanonical measure
1523: on $\Sigma_E$. The kinetic energy submanifolds ${\Bbb S}_t^{N-1}=\{(p_1,\dots
1524: ,p_N)\in{\Bbb R}^N\vert\sum_{i=1}^N\frac{1}{2}p_i^2=t\}$ are hyperspheres.
1525:
1526: In other words, at very large $N$ the microcanonical measure
1527: mathematically extends over a whole energy surface but, as far as physics
1528: is concerned, a non-negligible contribution to the microcanonical measure is
1529: in practice given only by a small subset of an energy surface. This subset can
1530: be reasonably modeled by the product manifold $\Sigma^{N-1}_v\times{\Bbb S}
1531: _t^{N-1}$, because the total kinetic and total potential energies - having
1532: arbitrarily small fluctuations, provided that $N$ is large enough - can be
1533: considered almost constant. Thus, since ${\Bbb S}_t^{N-1}$ at any $t$ is always
1534: an hypersphere, a change in the topology of $\Sigma^{N-1}_v$ directly entails
1535: a change of the topology of $\Sigma^{N-1}_v\times {\Bbb S}_t^{N-1}$, that is
1536: of the effective model-manifold for the subset of $\Sigma_E$ where the dynamics
1537: mainly ``lives'' at a given energy $E$.
1538:
1539: At small $N$, the model with a single product manifold is no longer good and
1540: should be replaced by the non-countable union $\bigcup_{v\in{\cal I}\subset
1541: {\Bbb R}}\Sigma^{N-1}_v\times{\Bbb S}_{E-v}^{N-1}$, with $v$ assuming all the
1542: possible values in a real interval ${\cal I}$. From this fact the smoothing
1543: of the energy dependence of thermodynamic variables follows.
1544: Nevertheless, the geometric and
1545: topologic signals of the phase transition can remain much sharper than the
1546: thermodynamic signals also at small $N$ $(< 100)$, as it is witnessed by the
1547: $2d$ lattice $\varphi^4$ model \cite{fps1,fps2}.
1548:
1549: Finally, let us comment about the relationship between intrinsic geometry,
1550: in terms of
1551: which we discussed the geometrization of the dynamics, and extrinsic
1552: geometry, dealt with in the present Section.
1553:
1554: The most direct and intriguing link is estabilished by the
1555: expression for microcanonical averages of generic observables of the kind
1556: $A(q)$, with $q=(q_1,\dots,q_N)$,
1557: \begin{equation}
1558: \langle A\,\rangle_{mc} = \frac{1}{\Omega_{2N}(E)} \int_{H(p,q)\leq E}
1559: d^Np\, d^Nq\, A(q)
1560: =\frac{1}{Vol(M_E)}\int_{V(q)\leq E} d^Nq\, [E - V(q)]^{N/2}\, A(q)~,
1561: \label{medie1}
1562: \end{equation}
1563: where $M_E=\{q\in{\Bbb R}^N\vert V(q)\leq E\}$. Eq. (\ref{medie1}) is
1564: obtained by means of a Laplace-transform method \cite{Pearson}; it is
1565: remarkable that $[E-V(q)]^N\equiv det(g_J)$, where $g_J$ is the Jacobi metric
1566: whose geodesic flow coincides with newtonian dynamics (see Section $V$),
1567: therefore $d^Nq\, [E - V(q)]^{N/2}\equiv d^N q\, \sqrt{det (g_J)}$
1568: is the invariant Riemannian volume element of $(M_E,g_J)$. Thus,
1569: \begin{equation}
1570: \frac{1}{Vol(M_E)}\int_{V(q)\leq E} d^Nq\, [E - V(q)]^{N/2}\, A(q) \equiv\;
1571: \frac{1}{Vol(M_E)}\int_{M_E} d^Nq\, \sqrt{det (g_J)}\, A(q)~,
1572: \label{medie2}
1573: \end{equation}
1574: which means that the microcanonical averages $\langle A(q)\,\rangle_{mc}$
1575: can be expressed as Riemannian integrals on the mechanical manifold
1576: $(M_E,g_J)$.
1577:
1578: In particular, this also applies to the microcanonical definition of entropy
1579: \begin{equation}
1580: S = k_B \log \int_{H(p,q)\leq E} d^N p\, d^N q\, =
1581: k_B \log \int_0^E dE^\prime \int_{\Sigma_{E^\prime}} \frac{d\sigma}
1582: {\Vert{\nabla H}\Vert}~,
1583: \label{entropy1}
1584: \end{equation}
1585: which is alternative to that given in Eq.(\ref{entropy}), though equivalent
1586: to it in the large $N$ limit. We have
1587: \begin{eqnarray}
1588: S &=& k_B \log \left[ \frac{1}{C \Gamma (N/2+1)} \int_{V(q)\leq E}
1589: d^Nq \;[E - V(q)]^{N/2}\right] \nonumber \\
1590: &\equiv & k_B \log \int_{M_E} d^N q\, \sqrt{det (g_J)}+\, const\,.~~,
1591: \label{entropy2}
1592: \end{eqnarray}
1593: where the last term gives the entropy as the logarithm of the
1594: Riemannian volume of the manifold.
1595:
1596: The topology changes of the surfaces $\Sigma_v^{N-1}$, that are to be
1597: associated with phase transitions, will deeply affect also the geometry of
1598: the mechanical manifolds $(M_E, g_J)$ and $(M\times{\Bbb R}^2, g_E)$
1599: and, consequently, they will affect the average instability properties of
1600: their geodesic flows. In fact, Eq.(\ref{lambda}) links some curvature
1601: averages of these manifolds with the numeric value of the largest Lyapunov
1602: exponent. Loosely speaking, major topology changes of $\Sigma_v^{N-1}$ will
1603: affect microcanonical averages of geometric quantities
1604: computed through Eq.(\ref{medie1}), likewise entropy, computed
1605: through Eq.(\ref{entropy2}).
1606:
1607: Thus, the peculiar temperature patterns displayed by the largest Lyapunov
1608: exponent at a second-order phase transition point -- in the present paper
1609: reported for the $3d$ $XY$ model, in Ref.\cite{CCCPPG} reported for
1610: lattice $\varphi^4$ models -- appear as reasonable consequences of the deep
1611: variations of the topology of the equipotential hypersurfaces of configuration
1612: space.
1613:
1614: We notice that topology seems to provide a common ground to the roots of
1615: microscopic dynamics and of thermodynamics and, notably, it can account for
1616: major qualitative changes simultaneously occurring in both dynamics and
1617: thermodynamics when a phase transition is present.
1618:
1619: \medskip
1620: \subsection{Some preliminary numerical computations}
1621: \label{5a}
1622: \medskip
1623: Let us briefly report on some preliminary numerical computations concerning
1624: the extrinsic geometry of the hypersurfaces $\Sigma_E$ in the case of the $3d$
1625: XY model.
1626:
1627: The first point about extrinsic geometry that we numerically addressed was to
1628: check whether the inverse of the temperature, that appears in
1629: Eq.(\ref{deformaz}), can be reasonably factorized into the product of a smooth
1630: ``deformation factor'' $D(E)$ and of the total mean curvature
1631: $\int_{\Sigma_E}M_1d\sigma$.
1632: To this purpose, the two independently computed quantities
1633: $\langle 1/\Vert\nabla H\Vert^2\rangle_{mc}$ and $D(E)=[\int_{\Sigma_E}
1634: (d\sigma /\Vert\nabla H\Vert )(M_1^\star/\Vert\nabla H\Vert)]\,/\,
1635: [\int_{\Sigma_E}d\sigma M_1]$ are compared in Fig. \ref{D(E)},
1636: showing that actually $\int_{\Sigma_E}(d\sigma /\Vert\nabla H\Vert )
1637: (M_1^\star/\Vert\nabla H\Vert) \simeq
1638: \langle 1/\Vert\nabla H\Vert^2\rangle_{mc}\int_{\Sigma_E}d\sigma M_1$.
1639: In other words, $D(E)\simeq\langle 1/\Vert\nabla H\Vert^2\rangle_{mc}$
1640: and no ``singular'' feature in its energy pattern
1641: seems to exist, what suggests that $\int_{\Sigma_E} d\sigma \ M_1$ has to
1642: convey all the information
1643: relevant to the detection of the phase transition.
1644: There is no reason to think that the validity of the
1645: factorization given in Eq.(\ref{deformaz}) is limited to the special case
1646: of the XY model.
1647:
1648: The other point that we tackled concerns an indirect quantification of how
1649: a phase space trajectory curves around and knots on the $\Sigma_E$ to which
1650: it belongs. We can expect that the way in which an hypersurface
1651: $\Sigma_E$ is ``filled'' by a phase space trajectory living on it will be
1652: affected by the geometry and the topology of the $\Sigma_E$.
1653: In particular, we computed the normalized autocorrelation function of the time
1654: series $M_1[x(t)]$ of the mean curvature at the points of $\Sigma_E$ visited
1655: by the phase space trajectory,
1656: that is, the quantity
1657: \begin{equation}
1658: \Gamma (\tau )=\langle \delta M_1(t+\tau )\delta M_1(t)\rangle_t~~,
1659: \label{autocor}
1660: \end{equation}
1661: where $\delta M_1(t)=M_1(t)-\langle M_1(t^\prime )\rangle_{t^\prime}$ is the
1662: fluctuation with respect to the average (the ``process'' $M_1(t)$ is supposed
1663: stationary). Our aim was to highlight the extrinsic
1664: geometric-dynamical counterpart of a symmetry-breaking phase transition.
1665:
1666: The practical computation of $\Gamma (\tau )$ proceeds by working out the
1667: Fourier power spectrum $\vert \tilde M_1(\omega )\vert^2$ of $M_1[x(t)]$,
1668: obtained by averaging $15$ spectra computed by an FFT algorithm with a mesh
1669: of $2^{15}$ points and a sampling time $\Delta t=0.1$.
1670: Some typical results for $\Gamma (\tau )$, obtained at different temperatures,
1671: are reported in Fig.\ref{Gamma}. The patterns $\Gamma (\tau )$ display a
1672: first regime of very fast decay, which is not surprising because of the
1673: chaoticity of the trajectories at
1674: any energy, followed by a longer tail of slower
1675: decay. An autocorrelation time $\tau_{corr}$ can be defined
1676: through the first intercept of $\Gamma (\tau )$ with an almost-zero level
1677: ($\Gamma =0.01$).
1678: In Fig.\ref{tau} we report the values of $\tau_{corr}$ so defined vs.
1679: temperature.
1680: In correspondence of the phase transition (whose critical temperature is
1681: marked by a vertical dotted line), $\tau_{corr}$ changes its temperature
1682: dependence: by lowering the temperature,
1683: below the transition $\tau_{corr}(T)$ rapidly increases,
1684: whereas it mildly
1685: decreases above the transition. Below $T\simeq 0.9$, where the vortices
1686: disappear, the autocorrelation functions of $M_1$ look quite different and
1687: it seems no longer possible to coherently define a correlation time.
1688: This result has an intuitive meaning and confirms that the phase transition
1689: corresponds to a change in the microscopic dynamics, as already signaled by
1690: the largest Lyapunov exponent; however, notice that the correlation times
1691: $\tau_{corr}(T)$ are much longer than the inverse values of the corresponding
1692: $\lambda_1(T)$. Qualitatively, $\lambda_1(T)$ and $\tau_{corr}^{-1}(T)$
1693: look similar, however the two functions are not simply related.
1694: \medskip
1695: \section{Discussion and perspectives}
1696: \medskip
1697: The microscopic Hamiltonian dynamics of the classical Heisenberg XY model in
1698: two and three spatial dimensions has been numerically investigated.
1699: This has been possible after the addition to the Heisenberg potentials of
1700: a standard (quadratic) kinetic energy term.
1701: Special emphasis has been given to the study of the dynamical counterpart
1702: of phase transitions, detected through the time averages of conventional
1703: thermodynamic observables, and to the new mathematical concepts that are
1704: brought about by Hamiltonian dynamics.
1705:
1706: The motivations of the present study are given in the Introduction.
1707: Let us now summarize what are the outcomes of our investigations and comment
1708: about their meaning. There are three main topics, tightly related one to
1709: the other:
1710: \begin{itemize}
1711: \item{} the phenomenological description of phase transitions through the
1712: natural, microscopic dynamics in place of the usual Monte Carlo stochastic
1713: dynamics;
1714: \smallskip
1715: \item{} the investigation, in presence of phase transitions, of certain
1716: aspects of the (intrinsic) geometry of the mechanical manifolds where the
1717: natural dynamics is represented as a geodesic flow;
1718: \smallskip
1719: \item{} the discussion of the relationship between the (extrinsic)
1720: geometry of constant energy hypersurfaces of phase space and thermodynamics.
1721: \smallskip
1722: \end{itemize}
1723: About the first point, we have found that microscopic Hamiltonian dynamics very
1724: clearly evidences the presence of a second order phase transition through the
1725: time averages of conventional thermodynamic observables. Moreover, the
1726: familiar
1727: sharpening effects, at increasing $N$, of the specific heat peak and
1728: of the order parameter bifurcation are observed. The evolution of the
1729: order parameter with respect to the physical time (instead of the
1730: fictitious Monte Carlo time) is also accessible, showing the appearance of
1731: Goldstone modes and that, in presence of a second order phase transition,
1732: there is a clear tendency to the freezing of transverse fluctuations
1733: of the order parameter when $N$ is increased.
1734: The "freezing" is observed together with a reduction of the
1735: longitudinal fluctuations, i.e. the rotation of the magnetization
1736: vector slows down, preparing the breaking of the $O(2)$ symmetry at
1737: $N\rightarrow\infty$. At variance, when a Kosterlitz-Thouless transition is
1738: present, at increasing $N$ the magnetization vector has a faster rotation
1739: and a smaller norm, preparing the absence of symmetry-breaking in the
1740: $N\rightarrow\infty$ limit as expected.
1741:
1742: Remarkably, to detect phase transitions, microscopic Hamiltonian dynamics
1743: provides us with additional observables of purely dynamical nature, i.e.
1744: without statistical counterpart: Lyapunov exponents. Similarly to what we and
1745: other authors already reported for other models (see Introduction), also
1746: in the case of the $3d$ XY model a peculiar
1747: temperature pattern of the largest Lyapunov exponent shows up in presence of
1748: the second order phase transition, signaled by a ``cuspy'' point.
1749: By comparing the patterns $\lambda_1(T)$ given by Hamiltonian dynamics and by
1750: a suitably defined random dynamics respectively, we suggest that the
1751: transition between thermodynamically ordered and disordered phases has its
1752: microscopic dynamical counterpart in a transition between weak and strong
1753: chaos. Though $a-posteriori$ physically reasonable, this result is far from
1754: obvious, because the largest Lyapunov exponent measures the average
1755: {\it local instability} of the dynamics, which $a-priori$ has little to do
1756: with a {\it collective}, and therefore global, phenomenon such as a phase
1757: transition.
1758: The effort to understand the reason of such a sensitivity of $\lambda_1$ to
1759: a second order phase transition and to other kinds of transitions, as
1760: mentioned in the Introduction, is far reaching.
1761:
1762: Here we arrive to the second
1763: point listed above. In the framework of a Riemannian geometrization of
1764: Hamiltonian dynamics, the largest Lyapunov exponent is related to the
1765: curvature properties of suitable submanifolds of configuration space whose
1766: geodesics coincide with the natural motions. In the mathematical light of this
1767: geometrization of the dynamics, and after the numerical evidence of a sharp
1768: peak of curvature fluctuations at the phase transition point, the peculiar
1769: pattern of $\lambda_1(T)$ is due to some major change occurring to the geometry
1770: of mechanical manifolds at the phase transition. Elsewhere, we have conjectured
1771: that indeed some major change in the {\it topology} of configuration space
1772: submanifolds should be the very source of the mentioned major change of
1773: geometry.
1774:
1775: Thus, we have made a first attempt to provide an analytic argument
1776: supporting this topological hypothesis (third point of the above list).
1777: This is based on the appearance of a non trivial relationship between the
1778: geometry of constant energy hypersurfaces of phase space with their topology
1779: and with the microcanonical definition of thermodynamics. Even still in a
1780: preliminary formulation, our reasoning already seems to indicate the
1781: topology of energy hypersurfaces as the best candidate to explain the deep
1782: origin of the dynamical signature of phase transitions detected through
1783: $\lambda_1(T)$.
1784:
1785: The circumstance, mentioned in the preceding Section, of the persistence
1786: at small $N$ of geometric and topologic signals of the phase transition
1787: that are much sharper than the thermodynamic signals is of
1788: prospective interest for the study of phase transition phenomena in finite,
1789: small systems, a topic of growing interest thanks to the modern developments
1790: - mainly experimental - about the physics of nuclear, atomic and molecular
1791: clusters, of conformational phase transitions in homopolymers and proteins,
1792: of mesoscopic systems, of soft-matter systems of biological interest.
1793: In fact, some unambiguous information for small systems - even about the
1794: existence itself of a phase transition - could be better obtained by means
1795: of concepts and mathematical tools outlined here and in the quoted papers.
1796: Here we also join the very interesting line of thought of Gross and
1797: collaborators \cite{Gross,Gross3} about the microcanonical description of phase
1798: transitions in finite systems.
1799:
1800: Let us conclude with a speculative comment about another possible direction of
1801: investigation related with this signature of phase transitions through Lyapunov
1802: exponents.
1803: In a field-theoretic framework, based on a path-integral
1804: formulation of classical mechanics \cite{reuter,gozzi1,gozzi2}, Lyapunov
1805: exponents are defined through the expectation values of suitable operators.
1806: In the field-theoretic framework, ergodicity breaking appears related to a
1807: supersymmetry breaking \cite{reuter}, and Lyapunov exponents are related to
1808: mathematical objects that have many analogies with topological concepts
1809: \cite{gozzi2}.
1810:
1811: The new mathematical concepts and methods, that the Hamiltonian
1812: dynamical approach brings about, could hopefully be useful also in the study
1813: of more ``exotic'' transition phenomena than those tackled in the present
1814: work. Besides the above mentioned soft-matter systems, this could be the case
1815: of transition phenomena occurring in amorphous and disordered materials.
1816:
1817: \medskip
1818: \section{acknowledgments}
1819: \medskip
1820: We warmly thank L. Casetti, E.G.D. Cohen,R. Franzosi and L. Spinelli for
1821: many helpful discussions.
1822: During the last year C.C. has been supported by the
1823: NSF (Grant \# 96-03839) and by the La Jolla Interfaces in Science
1824: program (sponsored by the Burroughs Wellcome Fund).
1825: This work has been partially supported by I.N.F.M., under the PAIS
1826: {\it Equilibrium and non-equilibrium dynamics in condensed matter systems},
1827: which is hereby gratefully acknowledged.
1828:
1829: \medskip
1830: \section{Appendix A}
1831: \medskip
1832:
1833: Let us briefly explain how a random markovian dynamics is constructed on a
1834: given constant energy hypersurface of phase space. The goal is to
1835: compare the energy dependence of the largest Lyapunov exponent computed
1836: for the Hamiltonian flow and for a suitable random walk respectively.
1837: One has to devise an algorithm to generate a random walk on a given energy
1838: hypersurface such that, once the time interval $\Delta t$ separating two
1839: successive steps is assigned, the average increments of the coordinates
1840: are equal to the average increments of the same coordinates for the
1841: differentiable dynamics integrated with a time step $\Delta t$.
1842: In other words, the random walk has to roughly mimick the differentiable
1843: dynamics with the exception of its possible time-correlations.
1844:
1845: One starts with a random initial configuration of the coordinates
1846: $q_{i},~ i=1,2,\ldots,N$, uniformly distributed in the interval
1847: $[0,2\pi]$, and with a random gaussian-distributed choice of the
1848: coordinates $p_{i}$.
1849: The random pseudo-trajectory is generated according to the simple scheme
1850: \begin{eqnarray}
1851: (q_{i})_{(k+1)\Delta t} &\mapsto& (q_{i})_{k \Delta t} +
1852: \alpha _{q} G_{i,k} \Delta t \nonumber\\
1853: (p_{i})_{(k+1)\Delta t} &\mapsto& (p_{i})_{k \Delta t} +
1854: \alpha _{p} G_{i,k} \Delta t~~,
1855: \label{step.micro}
1856: \end{eqnarray}
1857: where $\Delta t$ is the time interval associated to one step $k \mapsto k+1$
1858: in the markovian chain, $G_{i,k}$ are gaussian distributed random numbers with
1859: zero expectation value and unit variance; the parameters
1860: $\alpha_{q}$ and $\alpha_{p}$ are the variances of the processes
1861: $(q_i)_k$ and $(p_i)_k$.
1862: These variances are functions of the energy per degree of freedom
1863: $\varepsilon$. They have to be set equal to the
1864: numerically computed average increments of the coordinates obtained
1865: along the differentiable trajectories integrated with the same time step
1866: $\Delta t$, that is
1867: \begin{eqnarray}
1868: \alpha_{q}(\epsilon)&=&\left\langle\left[{\frac{1}{N}\sum_{i=1}^{N}
1869: \frac{(q_{i}(t+\Delta t)
1870: -q_{i}(t))^{2}}{\Delta t}}\right]^{1/2}\right\rangle_t \sim
1871: \left\langle\left[ {\frac{1}{N}\sum_{i=1}^{N}p_{i}^{2}}\right]^{1/2}
1872: \right\rangle_t\sim \sqrt{T}\nonumber\\
1873: \alpha_{p}(\epsilon)&=&\left\langle\left[{\frac{1}{N}\sum_{i=1}^{N}
1874: \frac{(p_{i}(t+\Delta t)-p_{i}(t))^{2}}{\Delta t}}\right]^{1/2}\right
1875: \rangle_t\sim
1876: \left\langle\left[{\frac{1}{N}\sum_{i=1}^{N} \dot{p}_{i}^{2}}\right]^{1/2}
1877: \right\rangle_t~~,
1878: \end{eqnarray}
1879: where $T$ is the temperature.
1880: Then, in order to make minimum the energy fluctuations around any given value
1881: of the total energy, a criterium to accept or reject a new step along the
1882: markovian chain has to be assigned.
1883: A similar problem has been considered by Creutz, who developed
1884: a Monte Carlo microcanonical algorithm \cite{Creutz}, where a
1885: "Maxwellian demon" gives a part of its energy to the system to let it move to
1886: a new configuration, or gains energy from the system,
1887: if the new proposed configuration produces an energy lowering. If the demon
1888: does not have enough energy to allow an energy increasing update of the
1889: coordinates, no coordinate change is performed. In this way, the total
1890: energy remains almost constant with only small fluctuations.
1891: As usual in Monte Carlo simulations, it is appropriate to fix the parameters
1892: so as the acceptance rate of the proposed updates of the configurations is
1893: in the range $30\%$ -- $60\%$.
1894:
1895: A reliability check of the so defined random walk, and of the
1896: adequacy of the phase space sampling through the number of steps
1897: adopted in each run, is obtained by computing the averages of typical
1898: thermodynamic observables of known temperature dependences.
1899:
1900: An improvement to the above described ``demon'' algorithm has been
1901: obtained through a simple reprojection on $\Sigma_E$ of the
1902: updated configurations \cite{Pettini};
1903: the coordinates generated by means of (\ref {step.micro})
1904: are corrected with the formulae
1905: \beq
1906: q_{i}(k \Delta t) \mapsto q_{i}(k \Delta t) +
1907: \left[\frac{(\frac{\partial H}{\partial q_{i}}) \Delta E}
1908: {\sum_{i=1}^N (p_{j}^2+
1909: (\frac{\partial H}{\partial q_{j}})^2 )}\right]_{x_{R}({k \Delta t})} \
1910: \eeq
1911: \[
1912: p_{i}({k \Delta t}) \mapsto p_{i}({k \Delta t}) -
1913: \left[\frac{p_{i} \Delta E}{\sum_{j=1}^N (p_{j}^2+
1914: (\frac{\partial H}{\partial q_{j}})^2 )}\right]_{x_{R}({k \Delta t})}~~, \
1915: \]
1916: where $\Delta E$ is the difference between the energy of the new configuration
1917: and the reference energy, and $x_{R}({k \Delta t})$ denotes the random
1918: phase space trajectory.
1919: At each assigned energy, the computation of the largest Lyapunov exponent
1920: $\lambda_{1}^R$ of this random trajectory is obtained by means of the
1921: standard definition
1922: \begin{equation}
1923: \lambda_{1}^R = \lim_{n \rightarrow \infty} \frac{1}{n \Delta t}\sum_{k=1}^n
1924: \log \frac{\| \zeta ((k+1) \Delta t)\|}{\| \zeta (k \Delta t)\|}~~,
1925: \label{bgs}
1926: \end{equation}
1927: where $\zeta (t)\equiv (\xi (t),\dot\xi (t))$ is given by the discretized
1928: version of the tangent dynamics
1929: \begin{equation}
1930: \frac{\xi_i ((k+1)\Delta t)-2 \xi_i(k\Delta t)+\xi_i ((k-1)\Delta t)}
1931: {\Delta {t}^2}+
1932: \left(\frac{\partial^2 V}{\partial q_{i} \partial q_{j}}\right)_
1933: {x_{R}(k \Delta t)}
1934: \xi_j (k \Delta t) = 0~~.
1935: \label{tandynR}
1936: \end{equation}
1937:
1938: For wide variations of the parameters ($\Delta t$ and
1939: acceptance rate),
1940: the resulting values of $\lambda_{1}^R$ are in very good agreement.
1941: Moreover, the algorithm is sufficiently stable
1942: and the final value of $\lambda_{1}^R$ is independent of
1943: the choice of the initial condition.
1944:
1945: A more refined algorithm could be implemented by constructing a random
1946: markovian process $q(t_k)\equiv [q_1(t_k),\dots ,q_N(t_k)]$ performing an
1947: importance sampling of the measure $d\mu =[E-V(q)]^{N/2-1}\,dq$ in
1948: configuration space. In fact, similarly to what is reported in
1949: Eq.(\ref{medie1}), one has \cite{Pearson} $\int_{H(p,q)= E} d^Np\, d^Nq\, =
1950: \, const\,\int_{V(q)\leq E} d^Nq\, [E - V(q)]^{N/2-1}$. A random process
1951: obtained by sampling such a measure -- with the additional property
1952: of a relation between the average increment and the physical time step
1953: $\Delta t$ as discussed above --
1954: would enter into Eq.(\ref{tandynR}) to yield $\lambda_1^R$.
1955: However, this would result in much heavier numerical computations (with some
1956: additional technical difficulty at large $N$) which was not worth in view of
1957: the principal aims of the present work.
1958:
1959:
1960: \begin{thebibliography}{999}
1961:
1962: \bibitem[a]{moni} Electronic address: mcs@arcetri.astro.it
1963:
1964: \bibitem[b]{cecilia} Electronic address: cclementi@ucsd.edu
1965:
1966: \bibitem[c]{marco} Also at INFN, Sezione di Firenze, Italy.
1967: Electronic address: pettini@arcetri.astro.it
1968:
1969: \bibitem{CCCP} L. Caiani, L. Casetti, C. Clementi and M. Pettini,
1970: Phys. Rev. Lett. {\bf 79}, 4361 (1997).
1971:
1972: \bibitem{Gallavotti} G. Gallavotti, {\it Meccanica Statistica},
1973: (Quaderni C.N.R., Roma, 1995).
1974:
1975: \bibitem{Thirring} P. Hertel, and W. Thirring, Ann. Phys. (NY) {\bf 63},
1976: 520 (1971).
1977:
1978: \bibitem{Lyndenbell} R.M. Lynden-Bell, in {\it Gravitational Dynamics},
1979: O. Lahav, E. Terlevich and R.J. Terlevich, eds.,
1980: Cambridge Contemporary Astrophysics, (Cambridge
1981: Univ. Press, 1996)
1982:
1983: \bibitem{Gross} D.H.E. Gross, Phys. Rep. {\bf 279}, 119 (1997); D.H.E. Gross
1984: and M.E. Madjet, {\it Microcanonical vs. canonical thermodynamics},
1985: archived in cond-mat/9611192
1986: see also the references quoted in these papers.
1987:
1988: \bibitem{Gross2} D.H.E. Gross, A. Ecker and X.Z. Zhang, Ann. Physik {\bf 5},
1989: 446 (1996).
1990:
1991: \bibitem{Gross1} A. H\"uller, Z. Physik B{\bf 95}, 63 (1994); M. Promberger,
1992: M. Kostner, and A. H\"uller, {\it Magnetic properties of finite systems:
1993: microcanonical finite-size scaling}, archived in cond-mat/9904265.
1994:
1995: \bibitem{Butera} P. Butera and G. Caravati, Phys. Rev. A {\bf 36}, 962 (1987).
1996:
1997: \bibitem{Leoncini} X. Leoncini, A. Verga, and S. Ruffo, Phys. Rev. E {\bf 57},
1998: 6377 (1998).
1999:
2000: \bibitem{Rapisarda} A. Bonasera, V. Latora, A. Rapisarda, Phys. Rev. Lett.
2001: {\bf 75}, 3434 (1995).
2002:
2003: \bibitem{Ruffo} M. Antoni and S. Ruffo, Phys. Rev E {\bf 52}, 2361 (1995).
2004:
2005: \bibitem{Antoni} M. Antoni and A. Torcini, Phys. Rev. E {\bf 57}, R6233 (1998).
2006:
2007: \bibitem{CCP1} L. Caiani, L. Casetti and M. Pettini, J. Phys. A:
2008: Math. Gen., {\bf 31}, 3357 (1998).
2009:
2010: \bibitem{CCCPPG} L. Caiani, L. Casetti, C. Clementi, G. Pettini,
2011: M. Pettini and R. Gatto, Phys. Rev E {\bf 57}, 3886 (1998).
2012:
2013: \bibitem{polimeri} C. Clementi, {\it Dynamics of homopolymeric chain models},
2014: Master Thesis, SISSA/ISAS, Trieste, (1996).
2015:
2016: \bibitem{Dellago} Ch. Dellago, H. A. Posch, W. G. Hoover, Phys. Rev. E
2017: {\bf 53}, 1485 (1996); Ch. Dellago, H. A. Posch, Physica A {\bf 230},
2018: 364 (1996); Ch.\ Dellago and H.\ A.\ Posch, Physica A {\bf 237}, 95 (1997);
2019: Ch.\ Dellago and H.\ A.\ Posch, Physica A {\bf 240}, 68 (1997) .
2020:
2021: \bibitem{Mehra} V. Mehra, R. Ramaswamy, preprint
2022: {\tt chao-dyn/9706011}.
2023:
2024: \bibitem{Berry} S. K. Nayak, P. Iena, K. D. Ball and R. S. Berry,
2025: J. Chem. Phys. {\bf 108}, 234 (1998).
2026:
2027: \bibitem{Firpo} M.-C. Firpo, Phys. Rev. E {\bf 57}, 6599 (1998).
2028:
2029: \bibitem{Ruffo_prl} V. Latora, A. Rapisarda, and S. Ruffo, Phys. Rev. Lett.
2030: {\bf 80}, 692 (1998).
2031:
2032: \bibitem{Ruffo_talk} V. Latora, A. Rapisarda, and S. Ruffo, Physica D (1998),
2033: in press.
2034:
2035: \bibitem{TobChes} J. Tobochnik, and G. V. Chester, Phys. Rev. B{\bf 20},
2036: 3761 (1979).
2037:
2038: \bibitem{Gupta} R. Gupta, and C.F. Baillie, Phys. Rev. B{\bf 45}, 2883 (1992).
2039:
2040: \bibitem{Lapo} L. Casetti, Phys. Scr. {\bf 51}, 29 (1995).
2041:
2042: \bibitem{Pearson} E. M. Pearson, T. Halicioglu, and W. A. Tiller, Phys. Rev.
2043: A{\bf 32}, 3030 (1985).
2044:
2045: \bibitem{Palmer} A thorough discussion about ergodicity breaking, non
2046: interchangeability of $t\rightarrow\infty$ and $N\rightarrow\infty$ limits
2047: can be found in: R.G. Palmer, Adv. Phys. {\bf 31}, 669 (1982).
2048:
2049: \bibitem{LPV} J. Lebowitz, J. Percus, and L. Verlet, Phys. Rev.
2050: {\bf 153}, 250 (1967).
2051:
2052: \bibitem{CCP} L. Casetti, C. Clementi and M. Pettini, Phys. Rev. E
2053: {\bf 54}, 5969 (1996).
2054:
2055: \bibitem{Pettini} M.Pettini, Phys. Rev. E {\bf 47}, 828 (1993).
2056:
2057: \bibitem{Eisenhart} L. P. Eisenhart, Ann. of Math. {\bf 30}, 591 (1939).
2058:
2059: \bibitem{doCarmo} M. P. Do Carmo, {\it Riemannian Geometry} (Birkh\"{a}user,
2060: Boston, 1992).
2061:
2062: \bibitem{CerrutiPettini} M.Cerruti-Sola and M. Pettini, Phys. Rev.
2063: E{\bf 53}, 179 (1996); M.Cerruti-Sola, R. Franzosi and
2064: M. Pettini, Phys. Rev. E{\bf 56}, 4872 (1997).
2065:
2066: \bibitem{cegdcp} L. Casetti, E.G.D. Cohen, and M. Pettini, Phys. Rev. Lett.
2067: {\bf 82}, 4160 (1999).
2068:
2069: \bibitem{fps1} R. Franzosi, L. Casetti, L. Spinelli, and M. Pettini,
2070: Phys. Rev. E{\bf 60}, 5009 (1999).
2071:
2072: \bibitem{fps2} R. Franzosi, M. Pettini, and L. Spinelli, {\it Topology and
2073: Phase transitions: a paradigmatic evidence}, Phys. Rev. Lett.,
2074: (1999) submitted; archived in cond-mat/9911235.
2075:
2076: \bibitem{thorpe} J.A. Thorpe, {\it Elementary Topics in Differential Geometry},
2077: (Springer, New York, 1979).
2078:
2079: \bibitem{laurence} P. Laurence, Zeit. Angew. Math. Phys.{\bf 40}, 258 (1989).
2080:
2081: \bibitem{nota1} A similar formula is obtained in: H.H. Rugh, J.Phys.A: Math.
2082: Gen. {\bf 31}, 7761 (1998), though without using the result of
2083: Ref. \protect\cite{laurence}, and in: C. Giardin\`a, R. Livi, J. Stat. Phys.
2084: {\bf 98}, 1027 (1998); however in these papers the relation between temperature
2085: and mean curvature was not established.
2086:
2087: \bibitem{ChernLashof} S. Chern, and R.K. Lashof, Michigan Math. J. {\bf 5},
2088: 5 (1958).
2089:
2090: \bibitem{Betti} M. Nakahara, {\it Geometry, Topology and Physics},
2091: (Adam Hilger, Bristol, 1989).
2092:
2093: \bibitem{bassi-indici} The mathematical details to justify such an assumption
2094: are discussed in: R. Franzosi, {\it Geometrical and Topological Aspects
2095: in the study of Phase Transitions}, PhD thesis, Department
2096: of Physics, University of Florence, (1999).
2097:
2098: \bibitem{Gross3} D.H.E. Gross and E. Votyakov, {\it Phase transitions in
2099: finite systems = topological peculiarities of the microcanonical entropy
2100: surface}, archived in cond-mat/9904073;
2101: D.H.E. Gross and E. Votyakov, {\it Phase transitions in ``small'' systems},
2102: archived in cond-mat/9911257; D.H.E. Gross, {\it Phase
2103: Transitions without thermodynamic limit}, archived in cond-mat/9805391.
2104:
2105: \bibitem{reuter} E. Gozzi and M. Reuter, Phys. Lett. B {\bf 233}, 383 (1989).
2106:
2107: \bibitem{gozzi1} E. Gozzi, M. Reuter and W. D. Thacker, Chaos, Solitons \&
2108: Fractals {\bf 2}, 441 (1992).
2109:
2110: \bibitem{gozzi2} E. Gozzi and M. Reuter, Chaos, Solitons \&
2111: Fractals {\bf 4}, 1117 (1994).
2112:
2113: \bibitem{Creutz} M. Creutz, Phys. Rev. Lett. {\bf 50}, 1411 (1983).
2114:
2115:
2116: \end{thebibliography}
2117: %---------------------------------------------------------------------------
2118: \newpage
2119:
2120: \begin{figure}
2121: \centerline{\includegraphics[width=0.80\linewidth]{Fig01.ps}}
2122: \caption{ The magnetization vector ${\bf M}(t)$ computed along a trajectory
2123: for the $2d$ XY model at different temperatures on a lattice
2124: of $N= 10 \times 10$. Each point represents a vector ${\bf M}(t)$ at
2125: a certain time $t$. }
2126: \label{figura.spin2d.10e10}
2127: \end{figure}
2128: \clearpage
2129: %-------------------------------------------------------
2130:
2131: \begin{figure}
2132: \centerline{\includegraphics[width=0.80\linewidth]{Fig02.ps}}
2133: \caption{ The magnetization vector ${\bf M}(t)$ at the temperature $T=0.74$,
2134: corresponding to the specific energy $\epsilon = 0.8$ and computed in
2135: a time interval $\Delta t = 10^5$, with a random initial configuration,
2136: on lattices of $N = 10 \times 10$ (external points)
2137: and of $N = 200 \times 200$
2138: (internal points). }
2139: \label{fig.spin2d.t0.74}
2140: \end{figure}
2141: \clearpage
2142: %-------------------------------------------------------
2143:
2144: \begin{figure}
2145: \centerline{\includegraphics[width=0.80\linewidth]{Fig03.ps}}
2146: \caption{ The magnetization vector ${\bf M}(t)$ at the temperature $T=1$,
2147: corresponding to the energy $\epsilon = 1.2$, computed in
2148: a time interval $\Delta t = 10^5$, with a random initial configuration on
2149: lattices of {\it a)} $N = 10 \times 10$, {\it b)} $N = 50 \times 50$,
2150: {\it c)} $N = 100 \times 100$ and {\it d)} $N = 200 \times 200$ sites,
2151: respectively. }
2152: \label{fig.spin2d.t1}
2153: \end{figure}
2154: \clearpage
2155: %-------------------------------------------------------
2156:
2157: \begin{figure}
2158: \centerline{\includegraphics[width=0.80\linewidth]{Fig04.ps}}
2159: \caption{ Specific heat at constant volume
2160: computed by means of Eq. (\ref{cvmicro})
2161: on a lattice of $N= 10 \times 10$ (open circles) and of $N= 15\times 15$
2162: (full triangles). Starlike squares refer to specific heat values
2163: computed by means of Eq. (\ref{specheat}) on a lattice of $N= 10 \times 10$
2164: . }
2165: \label{calspec_2d}
2166: \end{figure}
2167: \clearpage
2168: %-------------------------------------------------------
2169:
2170: \begin{figure}
2171: \centerline{\includegraphics[width=0.80\linewidth]{Fig05.ps}}
2172: \caption{ Vorticity function (plotted in {\it a)} linear scale
2173: and {\it b)} logarithmic scale) computed at different temperatures
2174: for lattices of $N=10\times 10$ (open circles) and $N=40\times 40$
2175: (full circles). The results of the Monte Carlo
2176: simulations for a lattice of $N=60\times 60$ (crosses)
2177: are from \protect\cite{TobChes}.
2178: The dashed line represents the power
2179: law ${\cal V}(t) \sim T^{10}$. }
2180: \label{fig.vort2d}
2181: \end{figure}
2182: \clearpage
2183: %-------------------------------------------------------
2184:
2185: \begin{figure}
2186: \centerline{\includegraphics[width=0.80\linewidth]{Fig06.ps}}
2187: \caption{ The largest Lyapunov exponents computed on different lattice
2188: sizes:
2189: $N = 10 \times 10$ (starred squares), $N= 20 \times 20$ (open triangles),
2190: $N= 40 \times 40$ (open stars), $N= 50 \times 50$ (open squares) and
2191: $N = 100 \times 100$ (open circles). In the inset, symbols have the same
2192: meaning. }
2193: \label{xy2d.lyap.num.fig}
2194: \end{figure}
2195: \clearpage
2196: %-------------------------------------------------------
2197:
2198: \begin{figure}
2199: \centerline{\includegraphics[width=0.80\linewidth]{Fig07.ps}}
2200: \caption{ The magnetization vector ${\bf M}(t)$, computed at the temperature
2201: $T = 1.7$, on lattices of different sizes. By increasing the lattice
2202: dimensions,
2203: the longitudinal fluctuations decrease. The time interval $\Delta t = 3.5
2204: \times 10^4 - 8 \times 10^4$ is the same for the four simulations. }
2205: \label{mag3d.e2}
2206: \end{figure}
2207: \clearpage
2208: %-------------------------------------------------------
2209:
2210: \begin{figure}
2211: \centerline{\includegraphics[width=0.80\linewidth]{Fig08.ps}}
2212: \caption{ The magnetization vector ${\bf M}(t)$ computed at different
2213: temperatures on a lattice of $N = 10 \times 10 \times 10$ spins. }
2214: \label{fig.1.spin3d.9}
2215: \end{figure}
2216: \clearpage
2217: %-------------------------------------------------------
2218:
2219: \begin{figure}
2220: \centerline{\includegraphics[width=0.50\linewidth]{Fig09a.ps}}
2221: \medskip
2222: \centerline{\includegraphics[width=0.50\linewidth]{Fig09b.ps}}
2223: \caption{ The magnetization vector ${\bf M}(t)$ computed at
2224: the temperature $T = 2.22$ (slightly higher than the critical value)
2225: on lattices of {\it a)} $N = 10 \times 10 \times 10$
2226: and {\it b)}$N = 15 \times 15 \times 15$, respectively.
2227: The time interval $\Delta t = 0.5 \times 10^4 - 1.5 \times 10^4$ is the same
2228: for both simulations. }
2229: \label{3d.fig.spin.altat}
2230: \end{figure}
2231: \clearpage
2232: %-------------------------------------------------------
2233:
2234: \begin{figure}
2235: \centerline{\includegraphics[width=0.80\linewidth]{Fig10.ps}}
2236: \caption{ The dynamical order parameter,
2237: defined as the average of the modulus $| {\bf M}(t) |$ along a trajectory,
2238: computed on lattices of
2239: $N = 10 \times 10 \times 10$ (full circles)
2240: and $N = 15 \times 15 \times 15$ (open circles). }
2241: \label{parord.3d.fig}
2242: \end{figure}
2243: \clearpage
2244: %-------------------------------------------------------
2245:
2246: \begin{figure}
2247: \centerline{\includegraphics[width=0.80\linewidth]{Fig11.ps}}
2248: \caption{ Specific heat at constant volume for the $3d$ model, computed
2249: by means of Eq. (\ref{cvmicro}) on
2250: lattices of $N= 8 \times 8 \times 8$ (open triangles),
2251: $N = 10 \times 10 \times 10$ (open circles), $N = 12 \times 12 \times 12$
2252: (open stars) and $N= 15 \times 15 \times 15$ (open squares) .
2253: Full circles refer to specific heat values computed by means of Eq. (\ref
2254: {specheat}) on a lattice of $N= 10 \times 10 \times 10$.
2255: The dashed line points out the critical temperature $T_c \simeq 2.17$ at which
2256: the phase transition occurs. }
2257: \label{calspec.3d.fig}
2258: \end{figure}
2259: \clearpage
2260: %-------------------------------------------------------
2261:
2262: \begin{figure}
2263: \centerline{\includegraphics[width=0.80\linewidth]{Fig12.ps}}
2264: \caption{ Vorticity function at different temperatures along a dynamical
2265: trajectory on a lattice of $N = 10 \times 10 \times 10$ sites. }
2266: \label{vort.fig3d}
2267: \end{figure}
2268: \clearpage
2269: %-------------------------------------------------------
2270:
2271: \begin{figure}
2272: \centerline{\includegraphics[width=0.80\linewidth]{Fig13.ps}}
2273: \caption{ The largest Lyapunov exponents computed at different temperatures
2274: for the $3d$ model.
2275: Numerical results are for lattices of $N= 10 \times 10 \times 10$ (open
2276: circles) and $N= 15 \times 15 \times 15$ (open stars).
2277: In the inset, symbols have the same meaning. The dashed line points out the
2278: temperature $T_c \simeq 2.17$ of the phase transition. The solid line puts in
2279: evidence the departure of $\lambda_1(T)$ from the linear growth. }
2280: \label{lyap.3d.fig}
2281: \end{figure}
2282: \clearpage
2283: %-------------------------------------------------------
2284:
2285: \begin{figure}
2286: \centerline{\includegraphics[width=0.80\linewidth]{Fig14.ps}}
2287: \caption{ The largest Lyapunov exponents computed by means of the random
2288: dynamics algorithm (full circles) are plotted
2289: in comparison with those computed by means of the standard dynamics
2290: (open stars) for a lattice of $N= 10 \times 10 \times 10$. }
2291: \label{randyn.3d.fig}
2292: \end{figure}
2293: \clearpage
2294: %-------------------------------------------------------
2295:
2296: \begin{figure}
2297: \centerline{\includegraphics[width=0.80\linewidth,angle=90]{Fig15.ps}}
2298: \caption{ Time average of Ricci curvature (open circles) and its r.m.s.
2299: fluctuations
2300: (full circles) at different temperatures computed for a lattice
2301: of $N = 40 \times 40$ sites. Solid lines are the analytic estimates obtained
2302: from a high temperature expansion. }
2303: \label{fig.Ricci2d}
2304: \end{figure}
2305: \clearpage
2306: %-------------------------------------------------------
2307:
2308: \begin{figure}
2309: \centerline{\includegraphics[width=0.80\linewidth,angle=90]{Fig16.ps}}
2310: \caption{ Time average of Ricci curvature (open triangles) and its r.m.s.
2311: fluctuations
2312: (full triangles) computed at different
2313: temperatures for a lattice of $N = 10 \times 10 \times 10$.
2314: Open circles and full diamonds refer to a lattice size
2315: of $N= 15 \times 15 \times 15$. Solid lines are the analytic estimates
2316: in the limit of high temperatures. The dashed line points out the temperature
2317: $T_c \simeq 2.17$ of the phase transition. }
2318: \label{keflutt.3dfig}
2319: \end{figure}
2320: \clearpage
2321: %-------------------------------------------------------
2322:
2323: \begin{figure}
2324: \centerline{\includegraphics[width=0.80\linewidth]{Fig17.ps}}
2325: \caption{ Analytic Lyapunov exponents computed for the $2d$ model by means of
2326: Eq.(\protect\ref{lambda}) without correction (dots) and incorporating
2327: the correction that accounts for
2328: the probability of obtaining negative sectional curvatures
2329: (full squares) for a lattice size of
2330: $N = 40 \times 40$
2331: are plotted in comparison with the numerical values
2332: of Fig. \ref{xy2d.lyap.num.fig}.
2333: The dashed lines are the asymptotic behaviors at high and low
2334: temperatures in the thermodynamic limit. }
2335: \label{prev.Lyap.2d}
2336: \end{figure}
2337: \clearpage
2338: %-------------------------------------------------------
2339:
2340: \begin{figure}
2341: \centerline{\includegraphics[width=0.80\linewidth]{Fig18.ps}}
2342: \caption{ Analytic Lyapunov exponents computed for the $3d$ model by means of
2343: Eq.(\protect\ref{lambda}) without correction (dots) and
2344: incorporating
2345: the correction that accounts for
2346: the probability of obtaining negative sectional curvatures (full circles)
2347: are plotted in comparison with the numerical values of
2348: Fig. \ref{lyap.3d.fig}.
2349: The dashed lines are the
2350: asymptotic behaviors at high and low temperatures in the thermodynamic
2351: limit. }
2352: \label{prev.Lyap.3d}
2353: \end{figure}
2354: \clearpage
2355: %-------------------------------------------------------
2356:
2357: \begin{figure}
2358: \centerline{\includegraphics[width=0.80\linewidth]{Fig19.ps}}
2359: \caption{ The deformation factor $D(E)=[\int_{\Sigma_E}(d\sigma /\Vert
2360: \nabla H\Vert )(M_1^\star/\Vert\nabla H\Vert)]\,/\,
2361: [\int_{\Sigma_E}d\sigma M_1]$ of Eq. (\ref{deformaz}) (open circles)
2362: is plotted vs. energy density $E/N$ and compared to the quantity
2363: $\langle 1/\Vert\nabla H\Vert^2\rangle$
2364: (open triangles). $N=10 \times 10\times 10$. }
2365: \label{D(E)}
2366: \end{figure}
2367: \clearpage
2368: %-------------------------------------------------------
2369:
2370: \begin{figure}
2371: \centerline{\includegraphics[width=0.80\linewidth]{Fig20.ps}}
2372: \caption{ The normalized autocorrelation functions $\Gamma(\tau)$ are plotted
2373: vs. time $\tau$ for a lattice of $N=10 \times 10\times 10$ and for
2374: four different
2375: values of the temperature (from top to bottom: $T=0.49, 1.28, 1.75, 2.16$). }
2376: \label{Gamma}
2377: \end{figure}
2378: \clearpage
2379: %-------------------------------------------------------
2380:
2381: \begin{figure}
2382: \centerline{\includegraphics[width=0.80\linewidth]{Fig21.ps}}
2383: \caption{ Autocorrelation times $\tau_{corr}$ are plotted vs. temperature $T$.
2384: The vertical dashed line points out the temperature $T_c \simeq2.17$ at
2385: which the phase transition occurs. }
2386: \label{tau}
2387: \end{figure}
2388: \clearpage
2389: %-------------------------------------------------------
2390: \end{document}
2391:
2392: