1: \documentclass[aps,pre,amsmath,amssymb,twocolumn,showpacs,superscriptaddress]{revtex4}
2: %\documentclass[aps,pre,amsmath,amssymb,twocolumn,superscriptaddress]{revtex4} % PRE
3: %%\documentclass[aps,prl,amsmath,amssymb,twocolumn,showpacs,byrevtex,superscriptaddress]{revtex4}
4: %\documentclass[prl,twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
5: \usepackage{dcolumn}% align table columns on decimal point
6: %\usepackage{bbm}% blackboard bold math
7: %\documentstyle[twocolumn,prl,aps,epsf,graphicx]{revtex}
8: %\documentstyle[twocolumn,prl,aps]{revtex}
9: %\documentstyle{article}
10: \usepackage{graphicx}
11: %\usepackage{epstopdf}
12: \begin{document}
13: %\draft
14: %\twocolumn[
15: %\hsize\textwidth\columnwidth\hsize\csname @twocolumnfalse\endcsname
16: %\draft
17: \title{
18: On the relationship between phase transitions and topological
19: changes in one dimensional models
20: }
21:
22: \author{L.~Angelani}
23: \affiliation{
24: Dipartimento di Fisica,
25: Universit\`a di Roma {\em La Sapienza}, P.le A. Moro 2, 00185 Roma, Italy
26: }
27: \affiliation{
28: INFM - CRS SMC, Universit\`a di Roma {\em La Sapienza}, P.le A. Moro 2, 00185 Roma, Italy
29: }
30: \affiliation{
31: Istituto dei Sistemi Complessi, CNR-INFM, V. dei Taurini 19,
32: 00185 Roma, Italy
33: }
34: \author{G.~Ruocco} \affiliation{
35: Dipartimento di Fisica,
36: Universit\`a di Roma {\em La Sapienza}, P.le A. Moro 2, 00185 Roma, Italy
37: }
38: \affiliation{
39: INFM - CRS SOFT, Universit\`a di Roma {\em La Sapienza}, P.le A. Moro 2, 00185 Roma, Italy
40: }
41: \author{F.~Zamponi} \affiliation{
42: Dipartimento di Fisica,
43: Universit\`a di Roma {\em La Sapienza}, P.le A. Moro 2, 00185 Roma, Italy
44: }
45: \affiliation{
46: INFM - CRS SOFT, Universit\`a di Roma {\em La Sapienza}, P.le A. Moro 2, 00185 Roma, Italy
47: }
48:
49: \date{\today}
50: \begin{abstract}
51: We address the question of the quantitative relationship between
52: thermodynamic phase transitions and topological changes in the
53: potential energy manifold analyzing two classes of one
54: dimensional models, the Burkhardt solid-on-solid model and the
55: Peyrard-Bishop model for DNA thermal denaturation,
56: both in the confining and non-confining version.
57: These models, apparently, do not fit [M. Kastner, Phys.
58: Rev. Lett. {\bf 93}, 150601 (2004)] in the general idea that the
59: phase transition is signaled by a topological discontinuity.
60: We show that in both models the phase
61: transition energy $v_c$ is actually non-coincident with, and
62: always higher than, the energy $v_\theta$ at which a topological
63: change appears. However, applying a procedure already successfully
64: employed in other cases as the mean field $\phi^4$ model, i.~e.
65: introducing a map ${\cal{M}}\!: v\to v_s$ from levels of the
66: energy hypersurface $V$ to the level of the stationary points
67: "visited" at temperature $T$, we find that ${\cal{M}}
68: (v_c)$=$v_{\theta}$. This result enhances the relevance of the
69: underlying stationary points in determining the
70: thermodynamics of a system, and extends the validity of
71: the topological approach to the study of phase transition to
72: the elusive one-dimensional systems considered here.
73: \end{abstract}
74: \pacs{64.60.-i, 05.70.Fh, 02.40.-k}
75:
76:
77: \maketitle
78:
79: %%%%%%%%%%%%%%% TEXT %%%%%%%%%%%%%%%%
80: Phase transitions are a very well understood subject
81: in statistical mechanics. They have been characterized in many different
82: ways in the last century and many observables related to the phase
83: transition ({\it e.g.} critical exponents, correlation lengths, etc.) have
84: been computed and measured with very high accuracy \cite{Ma}.
85:
86: Recently, a new characterization of phase transitions has been proposed
87: by Pettini and coworkers \cite{cccp,fps,cpc,fra_pet}.
88: These authors conjectured that, for classical systems defined by a continuous
89: potential energy function $V(\{q_i\}_{i=1\ldots N})$,
90: a thermodynamic phase transition, occurring at a temperature $T_c$,
91: is the manifestation of a topological discontinuity, taking place at a
92: specific value $v(T_c)=N^{-1} \langle V(q) \rangle$ (where $\langle \cdot \rangle$ is the
93: statistical average at temperature $T$) of the potential energy function $V$,
94: or, more precisely, taking place on the hypersurface $\Sigma_v =\{
95: (q_1,\dots,q_N)\in{\Bbb R}^N \vert\ V(q_1, \dots ,q_N) = Nv\}$, at $v=v(T_c)$.
96: The most striking consequence of this hypothesis is that the signature of
97: a phase transition is present in the topology of the configuration
98: space independently on the statistical measure defined on it.
99:
100: The changes in the topology are identified through the Morse
101: theory \cite{morse}: according to this theory the topological
102: changes in a manifold like $\Sigma_v$ are related to the
103: presence of stationary points of $V$ (points for which $\nabla
104: V$=$0$) at energy $v$.
105: However, the precise meaning of the correlation between topological
106: changes and phase transitions in the general case is still a open question.
107: From one hand there
108: is a theorem of Franzosi and Pettini \cite{fra_pet}, asserting
109: that, for ``smooth, finite-range and confining microscopic
110: interaction potentials $V$ with continuously varying
111: coordinates,\dots, a topology change of the
112: $\{\Sigma_v\}_{v\in{\Bbb R}}$ at some $v_{\theta}$ is a {\it
113: necessary} condition for a phase transition to take place at the
114: corresponding energy \dots value'' \cite{fra_pet}. On the other
115: hand, there are different numerical studies of various models
116: \cite{fps,XY,ktrig,ktrig2,phi4,kastner,gri_mos,kastner2,teix}
117: (almost all with potentials $V$ {\it not} fulfilling the
118: hypotheses of the theorem) for which a variety of results has been
119: obtained: some are in agreement with the ``{\it topological
120: hypothesis}'' of Pettini and coworkers, others seem to indicate
121: its failure.
122:
123: It is important to underline that the theorem in
124: Ref.~\cite{fra_pet} establishes a {\em necessary} condition for a phase
125: transition to take place. The problem to find {\em sufficiency}
126: conditions is still an open problem, as pointed out also by the
127: proponents of the hypothesis.
128: This question has been addressed in two recent papers: in Ref.~\cite{kastner}
129: a one-dimensional model, the Burkhardt model with nonconfining
130: potential (see below) was investigated, finding that
131: a topological change is present without a phase transition at finite temperature;
132: in Ref.~\cite{teix} the mean field spherical ferromagnet was considered, and it was found
133: that the same topological changes happen either in absence and in presence of a magnetic
134: field, while in the latter case no phase transition occurs
135: (it is worth noting, however, that these two models do not fulfill the
136: hypotheses of the theorem in Ref.~\cite{fra_pet}).
137: Moreover, in the thermodynamic limit, it is likely that, for any finite
138: interval of energy $I=[v_1,v_2]$, there is always a stationary point
139: of $V(q)$ with energy $v \in I$; thus, in the thermodynamic limit
140: a topological change occurs with probability one in any finite interval
141: of energy.
142: Most of these topological changes are --obviously--
143: not related to phase transitions,
144: and indeed it seems that a topological change must be {\em strong enough}
145: to be related to a thermodynamic phase transition
146: (see Ref.~\cite{XY} for a detailed discussion of this point).
147: An indication coming from the analysis of the models cited above is
148: that the presence of a phase transition should be related to the
149: presence of a {\em singularity} in the Euler characteristic at
150: a given energy $v_\theta$.
151: Thus, basically, the idea of the ``{\em topological}''
152: approach is that the phase transitions are correlated to abrupt
153: changes in topological
154: quantities defined on the stationary points of $V$ (as, for example,
155: the Euler characteristic).
156:
157: Another open question is the equivalence between the energies
158: at which phase transition ($v_c$) and change in the topology
159: ($v_{\theta}$) take place. The original conjecture of Pettini
160: and coworkers asserts that the two energies ``{\em correspond}''
161: (let's call this the {\it strong topological hypothesis}).
162: To our knowledge there is only one system within the hypotheses
163: of the theorem, the two dimensional $\varphi ^4$ model \cite{fps},
164: for which
165: the equivalence has been numerically established. In other two
166: systems with long range interactions, thus out of the theorem
167: hypotheses (the mean field $XY$ model \cite{XY} and the
168: mean-field $k$-trigonometric model \cite{ktrig,ktrig2}) the equivalence has been
169: analytically proved. There are, conversely, analytical results for
170: a different model system (mean field $\varphi ^4$ model
171: \cite{phi4}) for which the correspondence does not hold: the
172: energy $v_c$ at which phase transition takes place is higher than
173: the energy $v_{\theta}$ of the topological singularity
174: (we stress here that also in this case the hypotheses of the theorem
175: in Ref. \cite{fra_pet} are not fulfilled).
176:
177:
178: The latter discussion is of particular importance, not only in the
179: context of phase transitions, but also in analogy with
180: glassy systems. These systems are characterized --at the mean field level--
181: by a dynamical transition taking place at a given temperature $T_{_{MCT}}$ (or
182: equivalently at energy $v_{_{MCT}}$) predicted by mode-coupling
183: theory, and a (hypothesized) true phase transition at a lower
184: temperature $T_K$ (Kauzmann temperature) or energy $v_K$. From
185: numerical simulations of Lennard-Jones like systems, one observes
186: that the dynamical transition is strictly related to the
187: properties of the saddles visited by the system \cite{sad_lj,sad_cav}.
188: The concept of ``visited saddles'' is quantitatively worked out defining a
189: pseudo-potential $W$=$|\nabla V|^2$ and minimizing it during the
190: dynamic evolution of the system, thus obtaining a map
191: ${\cal{M}}_q:q \rightarrow q_s$ associating to each equilibrium
192: configuration point $q=\{q_1,\dots,q_N\}$ a minimum $q_s$ of $W$.
193: When averaged over the dynamic trajectory one obtains an energy
194: map: ${\cal{M}}:v \rightarrow v_s$. Absolute minima of $W$ (having
195: $W(q_s)$=$0$) correspond to stationary points (saddles and minima)
196: of $V$. We note that the presence of local minima of $W$, with
197: $W(q_s)\neq 0$ but small (corresponding to inflection directions
198: in $V$ profile), does not affect the result \cite{jcp_comment}:
199: the order of visited
200: saddles (number of negative eigenvalues of the Hessian matrix of
201: $V$) extrapolates to zero at $T_{_{MCT}}$, and the energy of
202: saddles stays always below the instantaneous energy. Moreover the
203: true thermodynamic transition is achieved when the number of
204: visited stationary points of order zero (minima of $V$) grows less
205: than exponentially with the system size (in the glassy terminology
206: when the ``{\em configurational entropy}'' or ``{\em complexity}''
207: goes to zero). Solvable mean field spin-glass models ($p$-spins),
208: which manifest the same phenomenology of structural glasses,
209: corroborate these findings in an analytical way \cite{cavagna}.
210: Then, what emerges from glassy systems, is the great importance of
211: the {\em underlying} stationary points in the description of the
212: various transitions (dynamical and thermodynamical) taking place
213: in these systems. It is worth to note, however,
214: that the consistency of this picture beyond
215: mean field is still matter of debate \cite{VMF}, and that the definition
216: of the map ${\cal M}$ is not unique also at the mean field level,
217: different definitions giving similar but not quantitatively equal
218: results \cite{phi4}.
219:
220: One can argue, in line with the ``{\em
221: topological}'' approach to phase transitions, that also for
222: non-glassy systems the concept of {\em underlying} stationary
223: points continues to be useful. It is important to emphasize that,
224: in the study of the glass transition, is the discontinuity of the
225: average density number of underlying stationary points that marks the
226: dynamical transition at $T_{_{MCT}}$. Driven by this observation,
227: we recently proposed that the map ${\cal{M}}:v \rightarrow v_s$
228: has to be applied in order to spot the phase transition, i.e., if
229: a topological discontinuity exists at energy $v_{\theta}$, the
230: phase transition is expected at an energy $v_c$ such that ${\cal
231: M} (v_c)$=$v_{\theta}$. This has been proved to work
232: (at least approximately) in
233: those cases (e.g. the mean field $\varphi ^4$ model) where the
234: original ``{\em strong topological hypothesis}'' (i.e. coincidence
235: between $v_c$ and $v_{\theta}$) failed. It is worth to point out
236: that those cases where it has been proved that $v_c \equiv
237: v_\theta$ do not constitute counterexamples for the application of
238: the map ${\cal M}$, as in all these cases it turns out that $v_c$
239: is a fixed point for the map, i.~e. ${\cal M} (v_c)$=$v_{c}$. In
240: conclusion, for all the cases investigated so far \cite{XY,ktrig,ktrig2,phi4}, it
241: results that whenever a phase transition (including also "dynamic"
242: transitions as the glass transition in LJ liquids \cite{sad_lj,sad_cav}
243: and p-spin systems \cite{cavagna})
244: is present at a certain energy $v_c$,
245: this transition is signaled by a discontinuity in the topology,
246: specifically in the Euler characteristic or in the complexity, at
247: an energy $v_\theta$ such that ${\cal M} (v_c)$=$v_{\theta}$. At
248: variance with the original ({\it strong}) topological hypothesis,
249: where it was supposed that $v_c=v_\theta$, we will refer to the
250: latter conjecture as {\it weak topological hypothesis}.
251: Note that the {\it weak topological hypothesis},
252: at variance with the {\it strong topological hypothesis},
253: depends on the statistical measure, as the map
254: ${\cal M} : v \rightarrow v_s$ is defined through an {\it average}
255: over the dynamical trajectory (or, equivalently, over the statistical
256: measure). We will discuss this point in detail in the following.
257:
258:
259: Two recent papers \cite{kastner,gri_mos} addressed the question
260: concerning the relationship between phase transitions and topology
261: in one dimensional models. Kastner \cite{kastner} studied
262: two versions of a solid-on-solid model, one showing a phase
263: transition at finite temperature and the other not; he found that
264: both models exhibit the same topological change, thus concluded
265: towards an ``unattainability of a purely topological criterion for
266: the existence of a phase transition'' \cite{kastner}. Grinza and
267: Mossa \cite{gri_mos} considered the Peyrard-Bishop model
268: \cite{pey,pey1}, which exhibits both a phase transition and a change in
269: the topology, but in this case $v_c$ and $v_{\theta}$ are not
270: coincident \cite{nota}. These papers contributed to extend the
271: analyzed cases for the understanding of necessary and sufficient
272: conditions for the {\em topological hypothesis}. However, they
273: seem to reach contradictory results, one supporting and the other
274: falsifying the topological hypothesis, even if the investigated models
275: share many similarities.
276:
277: The aim of this work is to try to clarify this apparent inconsistency
278: with regard to what has been discussed above.
279: In particular, we reanalyze the
280: model investigated by Kastner and by Grinza and Mossa. As a result
281: of this study {\em i)} analyzing the Peyrard-Bishop Model (PB)
282: \cite{pey,pey1} we numerically show that the two energies, although
283: different $v_c\neq v_{\theta}$, satisfy the ``{\em weak
284: topological hypothesis}'', i.e. ${\cal M}(v_c)$=$v_{\theta}$. We also
285: investigate a slight modification of the PB model (allowing
286: non-confined motion of the variables), where we are able to study
287: the same quantities in absence of a thermodynamic phase transition
288: at finite temperature,
289: again finding results in agreement with the ``{\em
290: weak topological hypothesis}''.
291: {\em ii)} Analyzing the Burkhardt
292: model we introduce a further parameter defining the position of
293: the pinning potential, that can be moved from the origin, i.e.
294: fully confining potential, to infinity, fully non-confining
295: potential.
296: We found that the phase transition actually exists for
297: {\it all} the position of the pinning potential, and its critical
298: temperature goes continuously to infinity as the pinning potential
299: position goes to infinity.
300: Moreover, we found
301: that also the generalized Burkhardt model falls into the class of
302: systems that satisfy the ``{\em weak topological hypothesis}'',
303: i.e. ${\cal M}(v_c)$=$v_{\theta}$.
304: As the position of the pinning potential is moved toward
305: infinity, the energy $v_s(T)$ of the ``{\it underlying saddles}''
306: tends to reach the energy $v_\theta$ at higher temperature;
307: when the nonconfining limit is reached, $v_s(T) \leq v_\theta$ for
308: all $T$. The topological singularity is visited only for $T \rightarrow \infty$
309: and this is the reason why the phase transition is not observed,
310: i.e. $T_c \rightarrow \infty$.
311:
312: The present findings support the idea that also in the case of one
313: dimensional models, the relevant topological
314: quantity related to a phase transition is obtained from underlying
315: stationary points obtained from the map ${\cal M}$. As we already
316: noted, the choice of the map ${\cal M}$ is not univocal: we
317: have chosen the one obtained through $W$ (with some {\em ad hoc}
318: modifications), however different
319: choices are possible (we mention here, for example, the map
320: obtained using Euclidean distances in configuration space
321: \cite{phi4,cavagna}). The robustness of this conclusion with respect to
322: the possible different choices of the map ${\cal M}$ is still an
323: open question which goes beyond the scope of this paper.
324:
325:
326: \section{The models}
327: The one dimensional models we study are all defined by the
328: Hamiltonian ${\cal H} = \sum_{i=1}^{N} p_i^2/2m +
329: V(\{q\}_{i=1\dots N})$ ($m$ is the mass of each particle), where $V$
330: is the potential energy. We consider two different classes of
331: models. The first one, introduced by Burkhardt \cite{bur} as a
332: model for localization-delocalization transition of interfaces, is
333: defined by the potential energy $V^{(1)}$:
334: \begin{equation}
335: V^{(1)}(\{q\}_{i=1\dots N}) = \sum_{i=1}^{N} K |q_{i+1}-q_i| +
336: \sum_{i=1}^{N} V^{(1)}_{p}(q_i) \ , \label{v_bur}
337: \end{equation}
338: where $K$ measures the strength of the force between neighboring pairs,
339: $V^{(1)}_{p}(q)$ is the on-site pinning potential,
340: and periodic boundary conditions are assumed $q_{N+1} \equiv q_1$.
341: We chose for $V^{(1)}_{p}(q)$ the following form
342: \begin{equation}
343: V^{(1)}_{p}(q) = \left\{
344: \begin{array}{cl}
345: + \infty \ \ \ \ \ & \mbox{for $q \leq 0$}
346: \\
347: 0 \ \ \ \ \ & \mbox{for $0 < q < L$}
348: \\
349: -U_0 \ \ \ \ \ & \mbox{for $L \leq q \leq L + R$}
350: \\
351: 0 \ \ \ \ \ & \mbox{for $q > L + R$}
352: \end{array} \right.\label{gburk}
353: \end{equation}
354: that generalizes the original form in Ref. \cite{bur} introducing
355: a parameter $L$ that gives the position of the pinning potential
356: (a square well of depth $U_0$ and width $R$, see Figure \ref{fig_1})
357: from the edge of the system. The case with $L=0$ coincides with
358: the original Burkhardt confining model, while the non-confining
359: case is retrieved in the $L\rightarrow \infty$ limit.
360:
361: \begin{figure}[t]
362: %%\epsfysize= 10 truecm
363: %%\begin{center}
364: \includegraphics[width=.5\textwidth]{fig1.eps}
365: %%\end{center}
366: \caption{Sketch of the on-site pinning potential $V^{(1)}_{p}(q) $
367: for a given choice of the control parameter $L/R$. } \label{fig_1}
368: \end{figure}%%
369:
370: The models of the second class are defined by the the potential energy
371: $V^{(2)}$ and $V^{(3)}$ of the form:
372: \begin{equation}
373: V^{(2,3)}(\{q\}_{i=1\dots N}) = \sum_{i=1}^{N} \frac{K}{2}
374: (q_{i+1}-q_i)^2 + \sum_{i=1}^{N} V^{(2,3)}_{p}(q_i) \ .
375: \label{v_pb}
376: \end{equation}
377: We consider two different versions of this model, one defined by
378: the on-site Morse potential, introduced by Peyrard and Bishop
379: as a simple model for DNA thermal denaturation
380: \cite{pey,pey1} (PB model)
381: \begin{equation}
382: V^{(2)}_{p}(q) = U_o \{ (e^{-q/R} -1)^2 -1\}\ ; \label{model1}
383: \end{equation}
384: the other is a symmetric version of the former (SPB model)
385: \begin{equation}
386: V^{(3)}_{p}(q) = U_o \{ (e^{-|q|/R} -1)^2-1\} \
387: , \label{model2}
388: \end{equation}
389: a slight modification of PB model that allows for a non-confined motion
390: of the variables (see Fig. \ref{fig_2}).
391: We note that the introduction of the modulus in
392: the Eq. \ref{model2} does not introduce discontinuities up to the
393: second derivative of the potential. The quantities $U_o$ and $R$
394: determine respectively the energy and the length scales of the
395: on-site potential (in the following all quantities will be
396: reported in $U_o$ and $R$ units). We further chose $m$=1. In all
397: the three cases a parameter of the Hamiltonian is related to the
398: strength of the inter-particles interactions ($K$) and, for the
399: case of $V^{(1)}$ a second parameter is the position of the
400: pinning potential $L$.
401:
402: \begin{figure}[t]
403: %%\epsfysize= 10 truecm
404: %%\begin{center}
405: \includegraphics[width=.5\textwidth]{fig2.eps}
406: %%\end{center}
407: \caption{Plots of the on-site pinning potentials $V^{(2)}_{p}(q)$
408: (full line) and $V^{(3)}_{p}(q)$ (dashed line). } \label{fig_2}
409: \end{figure}%%
410:
411: Specifically, the relevant quantity defining the relative weight
412: of the on-site with respect to interparticles potentials is the
413: dimensionless ratio $\xi$=$K R/U_o$ or $\xi$=$K R^2/2U_o$ for the
414: potential models (1) or (2,3) respectively, while the position of
415: the pinning potential is given by $\zeta=L/R$
416: for the potential model (1).
417:
418: The generalized Burkhardt model has been treated analytically,
419: while the Peyrard-Bishop models are studied numerically. In the
420: latter cases we performed isothermal molecular dynamics
421: simulations using Nos\'e-Hoover thermostat at different
422: temperatures for systems with $N$=$500$ degrees of freedom with
423: periodic conditions $q_{N+1}$=$q_1$. We studied different values
424: of the control parameter $\xi$, as an example the results are
425: reported for $\xi$=$0.05$ and $0.5$, all the other $\xi$ values give
426: results in qualitative agreement with the two reported examples.
427:
428: %%%%
429: \section{Burkhardt model}
430:
431: \subsection{Thermodynamics}
432: The thermodynamics of the Burkhardt model is known since many years
433: \cite{bur} for both the $\zeta=0$ and $\zeta=\infty$ cases. The
434: method of solution is briefly outlined below.
435:
436: The determination of the thermodynamic of systems described by a
437: potential function of the form
438: \begin{equation}
439: V(\{q\}_{i=1\dots N}) = \sum_{i=1}^{N} K |q_{i+1}-q_i| +
440: \sum_{i=1}^{N} V_{p}(q_i) \ , \label{forma}
441: \end{equation}
442: i.~e. similar to the case (1) (Eq.~\ref{v_bur}), goes through the
443: exploitation of the transfer matrix technique. Indeed, the
444: configurational partition function ${\cal Z}$ is given by:
445: \begin{equation}
446: {\cal Z}_N = \int dq_1\dots dq_N \; e^{-\beta V(\{q\}_{i=1\dots N})} \ ,
447: \label{Z}
448: \end{equation}
449: that, defining the transfer "matrix"
450: \begin{equation}
451: {\cal T}(x,y) = e^{-\beta K \vert x-y\vert} \; \; e^{-\beta
452: [V_{p}(x)+V_{p}(y)]/2} , \label{T}
453: \end{equation}
454: can be written as:
455: \begin{equation}
456: {\cal Z}_N =\int dq_1\dots dq_N \; \prod_{i=1}^N {\cal T}(q_i,q_{i+1}) \ , \label{Z2}
457: \end{equation}
458: recalling that $q_{N+1}\equiv q_1$.
459: With this notation, the (configurational) free energy density
460: \begin{equation}
461: f=-\frac{1}{\beta N} \log ( {\cal Z}_N )
462: \end{equation}
463: in the thermodynamic limit is promptly written as
464: \begin{equation}
465: f=-\frac{1}{\beta} \log ( \max{ \{\bar \lambda\}} ) \label{fp}
466: \end{equation}
467: where $\bar \lambda$ is the set of eigenvalues of the transfer
468: matrix, i.~e. the eigenvalues of the integral equation
469: \begin{equation}
470: \int dy \; {\cal T}(x,y) \phi(y) = \lambda \phi(x).
471: \end{equation}
472: The latter equation, with the substitution
473: \begin{equation}
474: \psi(x) = e^{\beta V_{p}(x)/2} \phi(x),
475: \end{equation}
476: turns out to be
477: \begin{equation}
478: \int dy \; e^{-\beta K \vert x-y\vert} \; \; e^{-\beta V_{p}(y)}
479: \psi(y) = \lambda \psi(x). \label{ei}
480: \end{equation}
481: The next step is performed by noticing that the operator $[
482: -d^2/dx^2+\beta^2 K^2 ]$ applied to $\exp(-\beta K\vert x-y \vert)$
483: produces a delta-function:
484: \begin{equation}
485: \left [ -\frac{d^2}{dx^2}+\beta^2 K^2 \right ] e^{-\beta K \vert
486: x-y\vert} = 2\beta K \delta(x-y),
487: \end{equation}
488: thus by applying the previous operator to the integral
489: equation~\ref{ei}, it can be transformed in a Schroedinger like
490: differential equation:
491: \begin{equation}
492: \left [ -\frac{d^2}{dx^2}+\beta^2 K^2 -\frac{2\beta K}{\lambda}
493: e^{-\beta V_{p}(x) } \right ] \psi(x) =0.
494: \end{equation}
495: This equation must be solved with the conditions that {\it i)} the
496: "eigenfunction" $\psi(x)$ was normalizable, and, {\it ii)} the
497: boundary condition (implicit in Eq.~\ref{ei}) $\psi'(0)/
498: \psi(0)=\beta K$ was fulfilled. In summary, the calculation of the
499: thermodynamic of system defined by the potential energy of the
500: form in Eq.~\ref{v_bur} is reduced to the solution of a
501: Schroedinger-like differential equation and, in particular, to the
502: finding the largest eigenvalue of the original integral equation
503: \ref{ei}. In general, as the eigenvalues are continuous and smooth
504: function of the parameters (among which the temperature), no phase
505: transitions are expected unless the two largest among them cross
506: each other.
507:
508: \subsection{The $\zeta=0$ case.}
509:
510: Let us now apply the procedure to the potential function in
511: Eq.~\ref{v_bur} for the case $\zeta=0$.
512: We do not report the details of the calculation, as they are based on
513: standard techniques for solving Schroedinger equation in Quantum Mechanics
514: \cite{QM};
515: in summary the "eigenvalues" $\lambda$ are determined by the equation:
516: \begin{equation}
517: z(\lambda)=\beta K \label{zeta0}
518: \end{equation}
519: with
520: \begin{eqnarray}
521: z(\lambda)= \frac{f_1(P,Q) \sin{(Q R)}-f_2(P,Q))\cos{(Q R)}}
522: {f_3(P,Q))\sin{(Q R)}+ f_4(P,Q)\cos{(Q R)}}, \label{zeta02}
523: \end{eqnarray}
524: having defined
525: \begin{eqnarray}
526: f_1(P,Q)=Q^2
527: \\ \nonumber
528: f_2(P,Q)=P Q
529: \\ \nonumber
530: f_3(P,Q)=P
531: \\ \nonumber
532: f_4(P,Q)=Q \label{deff}
533: \end{eqnarray}
534: and
535: \begin{eqnarray}
536: Q(\lambda) = \sqrt{\frac{2\beta K}{\lambda} \; e^{\beta
537: U_o}-\beta^2 K^2}
538: \\ \nonumber
539: P(\lambda) =\sqrt{\beta^2 K^2-\frac{2\beta K}{\lambda} }.
540: \end{eqnarray}
541:
542: The only possibility for the function $z(\lambda)$ to be real
543: (condition required for Eq.~\ref{zeta0} to have solution) is that
544: $P$ was real (if $Q$ become imaginary, $z(\lambda)$ is still
545: real), thus it exists a solution to Eq.~\ref{zeta0} only if $P$ is
546: real. Therefore, when $P$ vanishes, the eigenvalues $\lambda$
547: disappear (more specifically, disappear the eigenvalues of the
548: discrete spectrum, and only those of the continuum spectrum
549: remain), and the (configurational) free energy is discontinuous.
550: For each temperature, the condition $P(\lambda)=0$ is fulfilled
551: for a "critical" $\lambda$, given by:
552: \begin{equation}
553: \lambda_c=\frac{2}{\beta K}.
554: \end{equation}
555: Thus, the equation for the largest eigenvalue at the "critical"
556: point is given by $z(\lambda_c)=\beta_c K$, or
557: \begin{equation}
558: \sqrt{e^{\beta_c U_o} -1 } \tan{\left [ \beta_c K R
559: \sqrt{e^{\beta_c U_o} -1 } \; \right ]}= 1,
560: \end{equation}
561: which gives us the required equation for the critical (inverse)
562: temperature $\beta_c$. This equation can be rearranged,
563: introducing the control parameter $\xi=KR/U_0$, as:
564: \begin{equation}
565: \xi = \frac{1}{\beta_c U_o }\frac{1}{\sqrt{e^{\beta_c U_o} -1 }}
566: \arctan{\left [ \frac{1}{\sqrt{e^{\beta_c U_o} -1 }} \; \right
567: ]}.\label{equ}
568: \end{equation}
569: The plot of the critical temperature (in reduced units $k_BT/U_o$)
570: as a function of $\xi$ is reported
571: as full line in Fig.~3.
572:
573: \begin{figure}[t]
574: %%\epsfysize= 10 truecm
575: %%\begin{center}
576: \includegraphics[width=.5\textwidth]{fig3.eps}
577: %%\end{center}
578: \caption{Full line: critical temperature (in reduced units $k_BT_c/U_o$) as a
579: function of the control parameter $\xi$ from Eq.~\ref{equ} for the
580: $\zeta$=$0$ case.
581: Dashed line: temperature $T_J$ (in reduced units) at which the ``underlying saddle''
582: jumps from minimum to saddle as a function of $\xi$ (see Sec. V B).}
583: \label{fig_3}
584: \end{figure}
585:
586: The phase transition is of the localization-delocalization type.
587: The particles, kept together by the $K\vert x-y \vert$ term of the
588: potential, for $T<T_c$ are pinned close to the square well, while,
589: for $T>T_c$ are delocalized in the $q$-axis.
590:
591: A simple calculation leads to the value of the critical energy
592: $v_c$ (the equilibrium energy $v(T)$ at the transition point
593: $v_c=v(T_c)$). From Eq.~\ref{fp}, we have
594:
595: \begin{equation}
596: v(T)=\frac{\partial (\beta f)}{\partial \beta} =
597: -\frac{\lambda'(\beta)}{\lambda(\beta)} \label{ff}
598: \end{equation}
599: where $\lambda(\beta)$ is the solution of Eq.~\ref{zeta0}. Close
600: to the critical point, $\lambda(\beta)=2/\beta K$, thus
601: $\lambda'(\beta)/\lambda(\beta)=1/\beta$ and
602: \begin{equation}
603: v_c= k_B T_c
604: \end{equation}
605: independently from the value of $\xi$.
606:
607: \subsection{The $\zeta \ne 0$ case.}
608:
609: The calculation for the case of generic $\zeta$ values is
610: quite similar to the previous one.
611: Also in this case, the eigenvalues
612: $\lambda$ are determined by an equation like Eq.~\ref{zeta0},
613: $z(\lambda)=\beta K$,
614: with $z(\lambda)$ again given by Eq.~\ref{zeta02}
615: and with the $f_n(P,Q)$ functions ($n$=$1,\dots,4$) given by
616: \begin{eqnarray}
617: f_1(P,Q)&=&P \left [ Q^2 \cosh(P R \zeta) - P^2 \sinh(P R \zeta)
618: \right] \nonumber
619: \\ \nonumber
620: f_2(P,Q)&=&P^2 Q \exp{(P R \zeta)}
621: \\ \nonumber
622: f_3(P,Q)&=& \left [ P^2 \cosh(P R \zeta) - Q^2 \sinh(P R \zeta)
623: \right]
624: \\
625: f_4(P,Q)&=&P Q \exp{(P R \zeta)}. \label{deff2}
626: \end{eqnarray}
627: Obviously, Eqs.~\ref{deff2} recover Eqs.~\ref{deff} in the $\zeta
628: \rightarrow 0$ limit. The same considerations on the reality of
629: $P(\lambda)$ reported above apply to Eq.~\ref{deff2}.
630: Thus the condition $P(\lambda)=0$ define the critical
631: value of the eigenvalue, $\lambda_c=2/\beta K$,and the equation
632: for the critical temperature ($z(\lambda_c)=\beta_c K$) becomes:
633: \begin{eqnarray}
634: && \sqrt{e^{\beta_c U_o} -1 } \sin{\left ( \beta_c K R
635: \sqrt{e^{\beta_c U_o} -1 } \; \right )} \times \\ && \times \big [
636: \cos{\left ( \beta_c K R \sqrt{e^{\beta_c U_o} -1 } \; \right )}-
637: \nonumber
638: \\ && - \beta_c K R \zeta \sqrt{e^{\beta_c U_o} -1 } \sin{\left (
639: \beta_c K R \sqrt{e^{\beta_c U_o} -1 } \; \right )} \big ]^{-1}=
640: 1. \nonumber
641: \end{eqnarray}
642: Similar to the $\zeta=0$ case, this equation can be rearranged,
643: introducing the control parameter $\xi$, as:
644: \begin{equation}
645: \xi = \frac{1}{\beta_c U_o }\frac{1}{\sqrt{e^{\beta_c U_o} -1 }}
646: \arctan{\left ( \frac{1}{\sqrt{e^{\beta_c U_o} -1 }} \;
647: \frac{1}{1+\beta_c U_o \xi \zeta } \;\right )}.\label{equu}
648: \end{equation}
649: At variance with Eq.~\ref{equ}, this equation cannot be cast in
650: the form $\xi=\xi(\beta_c)$, thus it must be solved numerically to
651: plot the critical temperature as a function of the control
652: parameter $\xi$. This plot is reported in Fig.~\ref{fig_4} for
653: different values of $\zeta$.
654:
655: \begin{figure}[t]
656: %%\epsfysize= 10 truecm
657: %%\begin{center}
658: \includegraphics[width=.5\textwidth]{fig4.eps}
659: %%\end{center}
660: \caption{Critical temperature (in reduced units $k_BT_c/U_o$) as a
661: function of the control parameter $\xi$ from Eq.~\ref{equu} for
662: the indicated $\zeta$ value: $\zeta=0$ (full line), $0.5$ (dashed
663: line), $1.5$ (dotted line), $3.5$ (dash-dotted line) and $7.5$
664: (dash-dot-dotted line). } \label{fig_4}
665: \end{figure}
666:
667: As can be observed in Fig.~\ref{fig_4}, on increasing $\zeta$,
668: i.~e. on displacing the position of the square well towards high
669: value of the coordinate, the critical temperature, for a given
670: $\xi$ value, increases, expanding the amplitude of the "cold"
671: (localized or pinned) phase. This can be better seen in
672: Fig.~\ref{fig_5}, where the $\zeta$ dependence of the critical
673: temperature is reported for some values of $\xi$. We conclude this
674: section noticing that the phase transition actually exists for all
675: the value of $\zeta$, and in the limit of $\zeta \rightarrow
676: \infty$, the critical temperature goes {\it without
677: discontinuities} to infinity.
678: Therefore, we are lead to conclude that the model investigated in Ref.~\cite{kastner} to
679: demonstrate the unattainability of a purely topological criterion for the
680: existence of a phase transition is a ``borderline'' model,
681: in which the phase transition can be thought to be present at
682: ``$T$ infinity'' (even though the precise meaning of this statement
683: is not well defined). Then, to the same
684: topology (as we will see in the next section, the topology of the
685: potential function in Eq.~\ref{gburk} does not depend on the value of
686: $\zeta$) always corresponds a phase transition.
687:
688: To discuss the question of the {\it coincidence} (or not) of the
689: critical energy with the topological discontinuity, we need to
690: calculate $v_c(\xi,\zeta)$. Following the same argument reported
691: for the case $\zeta=0$ we conclude that the critical potential
692: energy depends on $\xi$ and $\zeta$ only through $T_c$: $v_c=k_B T_c$.
693: As an example, in Fig.~\ref{fig_6} we report
694: the caloric curve $v(T)$ as a function of the inverse
695: temperature for different $\zeta$ value and for $\xi=1$. For all
696: the $\zeta$ values, on the low-$\beta$ side the curves end at the
697: points ($\beta_c$, $v_c$); these points are aligned along the
698: $v_c(\beta_c)$ line (thick dotted line) given by
699: $v_c(\beta_c)=1/\beta_c$.
700:
701: \begin{figure}[t]
702: %%\epsfysize= 10 truecm
703: %%\begin{center}
704: \includegraphics[width=.5\textwidth]{fig5.eps}
705: %%\end{center}
706: \caption{Critical temperature (in reduced units $k_BT_c/U_o$) as a
707: function of the control parameter $\zeta$ from Eq.~\ref{equu} for
708: the indicated $\xi$ value: $\xi=2$ (full line),$1$(dashed line),
709: $0.5$ (dotted line), $0.25$ (dash-dotted line). } \label{fig_5}
710: \end{figure}
711:
712: \begin{figure}[t]
713: %%\epsfysize= 10 truecm
714: %%\begin{center}
715: \includegraphics[width=.5\textwidth]{fig6.eps}
716: %%\end{center}
717: \caption{Inverse temperature dependence (in reduced units $\beta
718: U_o$) of the equilibrium potential energy for $\xi=1$ and for the
719: indicated $\zeta$ value: $\zeta=0$ (full line), 0.5 (dashed line),
720: 1.5 (dotted line), 3.5 (dash-dotted line) and 7.5 (dash-dot-dotted
721: line). The thick-dotted line represents the $\beta$ dependence of
722: the potential energy in the high temperature phase.} \label{fig_6}
723: \end{figure}
724:
725: \subsection{Topology}
726: The analysis of the topological properties of the
727: Burkhardt model is reported by Kastner in Ref. \cite{kastner}.
728: He analyzed only the two limiting cases of confining
729: and non-confining models, corresponding in our notation
730: to $\zeta$=$0$ and $\zeta$=$\infty$ respectively.
731: He found that a topology change is present in both cases,
732: even if not really equal in ``strength''.
733: The value of the potential energy at which the topological change
734: appears is $v_{\theta}$=$0$, irrespective of the considered model.
735: One can easy generalize the above analysis to the general case
736: with arbitrary $\zeta$, and conclude that the topological change is
737: always located at energy $v_{\theta}$=$0$.
738: It is worth noting that the energy at which topological change appears
739: is lower than the thermodynamic transition energy: $v_c > v_{\theta}$
740: (see Fig. \ref{fig_6}).
741: We will further discuss this issue in Sec. V, after having described
742: the thermodynamics and topology of the PB and SPB models.
743:
744: %%%%%
745: \section{Peyrard-Bishop model}
746:
747: \subsection{Thermodynamics}
748: The thermodynamics of the Peyrad-Bishop model
749: (defined in Eq.s \ref{v_pb}, \ref{model1}) can be studied using
750: transfer matrix techniques, as the Burkhardt model
751: described in the previous section.
752: However, in this case approximated methods have to be considered
753: in order to obtain a corresponding Schroedinger
754: like differential equation.
755: In the region $\xi \gg 1$ and
756: temperature window $U_o \ll k_B T \ll \xi U_o$
757: the classical statistical mechanics problem is mapped
758: to the quantum Morse oscillator problem \cite{pey2,theo}.
759: Similarly to the case of Burkhardt potential,
760: the presence of a second order phase transition for the Peyrad-Bishop model
761: is signaled by the bounded-unbounded transition of the
762: lower state in the corresponding quantum problem.
763: In the above range of $\xi$ and $T$, Peyrard and Bishop obtained an analytical
764: expression for the transition temperature
765: $k_B T_c / U_o$=$4\sqrt{\xi}$
766: and transition energy
767: $v_c / U_o$=$k_B T_c/2U_o$=$2\sqrt{\xi}$.
768: For generic $(\xi,T)$ values, only numerical results can be used to
769: infer the existence and location of a phase transition. In Fig.
770: \ref{fig_7} we report the temperature dependence of the potential
771: energy per particle $v=V/N$ (full symbols) of the PB model for
772: two different values of $\xi$: $0.05$ (upper panel) and $0.5$ (lower
773: panel). Also reported in the figure are the energy $v_c$ (dot-dashed
774: line) and temperature $T_c$ (full line) of the phase transition point:
775: $v_c/U_o \simeq 0.61$ and $k_B T_c/U_o \simeq 1.22$ for $\xi$=$0.05$,
776: $v_c/U_o \simeq1.59$ and $k_B T_c/U_o \simeq3.20$ for $\xi$=$0.5$.
777: Dashed lines are
778: the $T$-dependence of the potential energy in the high $T$ phase:
779: $v(T)$=$k_B T/2$.
780:
781:
782: \begin{figure}[t]
783: %%\epsfysize= 10 truecm
784: %%\begin{center}
785: \includegraphics[width=.5\textwidth]{fig7.eps}
786: %%\end{center}
787: \caption{
788: Temperature dependence (in reduced units $k_B T /U_o$) of the
789: equilibrium potential energy $v$ (full symbols)
790: and energy $v_s$ of underlying saddles (open symbols)
791: for the PB model defined by Eq. \ref{model1} with
792: $\xi$=$0.05$ (upper panel) and $\xi$=$0.5$ (lower panel).
793: Also indicated in the figure are the values of
794: topological change energy $v_{\theta}$ (horizontal dotted lines),
795: phase transition energy $v_c$ (horizontal dot-dashed lines)
796: and transition temperature $T_c$ (vertical full line)
797: for $\xi$=$0.05$ ($v_{\theta}/U_o$=$0$, $v_c/U_o\simeq0.61$ and
798: $k_BT_c/U_o \simeq 1.22$)
799: and for $\xi$=$0.5$ ($v_{\theta}/U_o$=$0$, $v_c/U_o \simeq1.59$ and
800: $k_BT_c/U_o \simeq 3.20$).
801: Dashed lines are the $T$-dependence of the potential energy in the
802: high $T$ phase: $v(T)$=$k_B T/2$.
803: }
804: \label{fig_7}
805: \end{figure}%%
806:
807: \subsection{Topology}
808: The topology of the Peyrard-Bishop model is studied in the paper
809: of Grinza and Mossa \cite{gri_mos}.
810: A topological change is found at the energy value
811: $v_{\theta}$=$0$, corresponding to a
812: topological change in the hypersurfaces $\Sigma_v$ varying $v$:
813: from a close hypersurface for $v < v_{\theta}$ to an open one for
814: $v \geq v_{\theta}$ \cite{gri_mos}.
815: In Fig. \ref{fig_7} the value of $v_{\theta}$ is
816: indicated by an horizontal dotted line.
817: We note that, also in this case, the topological discontinuity is lower in energy
818: than the thermodynamic one: $v_c > v_{\theta}$.
819:
820: %%%%%
821:
822: \section{Symmetric Peyrard-Bishop model}
823:
824:
825: \subsection{Thermodynamics}
826: The Symmetric Peyrard-Bishop model defined by Eq.s \ref{v_pb},
827: \ref{model2} does not exhibit phase transition at finite $T$.
828: This can be view from the fact that
829: there is always a bound state in the corresponding quantum problem,
830: in analogy with the non-confined Burkhardt model
831: \cite{bur,cuesta}.
832: In Fig. \ref{fig_8} the
833: same quantities as in the PB case are reported for the SPB model:
834: energy $v$ (full symbols) for $\xi$=$0.05$ (upper panel) and $\xi$=$0.5$
835: (lower panel). It is evident in this case the absence of a phase
836: transition at finite $T$ (in the $T$-range investigated).
837:
838: \begin{figure}[t]
839: %%\epsfysize= 10 truecm
840: %%\begin{center}
841: \includegraphics[width=.5\textwidth]{fig8.eps}
842: %%\end{center}
843: \caption{
844: Temperature dependence (in reduced units $k_B T /U_o$) of the
845: equilibrium potential energy $v$ (full symbols)
846: and energy $v_s$ of underlying saddles (open symbols)
847: for the SPB model defined by Eq. \ref{model2} with
848: $\xi$=$0.05$ (upper panel) and $\xi$=$0.5$ (lower panel).
849: Also indicated in the figure is the value of
850: the topological change energy $v_{\theta}/U_o$=$0$ (dotted line)
851: for both cases.
852: }
853: \label{fig_8}
854: \end{figure}%%
855:
856: \subsection{Topology}
857: Following a similar argument as in Ref. \cite{kastner}, one can
858: see that also in the SPB case one has a topological change at exactly the
859: same energy level as in the PB model $v_{\theta}$=$0$
860: (even if not identical in strength to the previous one).
861: We refer to the papers in
862: Ref.s \cite{kastner,gri_mos} for a more detailed discussion of the
863: topology.
864: In Fig. \ref{fig_8} the value of $v_{\theta}$ is indicated by an
865: horizontal dotted line.
866:
867:
868: %%%%%
869: \section{Underlying saddles}
870:
871: In this section we study the properties of the stationary points visited
872: by the systems. The concept of ``{\em underlying saddles}'' was first
873: introduced in the study of glassy disordered systems \cite{sad_lj,sad_cav,cavagna}
874: to better understand the topological counterpart of the dynamic transition
875: taking place in these systems.
876: Recently, it has been applied also in the analysis of models that exhibit
877: thermodynamic phase transitions, in order to emphasize the role of topological
878: changes at the
879: ``{\em underlying saddles}'' energy in driving the phase transition
880: \cite{ktrig,ktrig2,phi4}.
881:
882: Here we apply the same methodology to investigate
883: the one dimensional systems introduced before.
884: Let start with the models having a continuous potential energy function,
885: the PB and SPB models, which allow for the usual definition of stationary points.
886: At the end of the section we will extend the argument to the discontinuous case
887: of Burkhardt model.
888:
889: \subsection{Peyrard-Bishop and Symmetric Peyrard-Bishop models}
890:
891: There are only two stationary points in the potential energy
892: hypersurface of both models: a minimum located at
893: $q_1$=$q_2$=$\dots$=$q_N$=$0$
894: and a saddle (with degenerate Hessian matrix) at
895: $q_1$=$q_2$=$\dots$=$q_N$=$\infty$ \cite{gri_mos}.
896: In order to associate one of the two stationary points
897: to each instantaneous configuration of the system,
898: we used a similar trick as in the analysis of
899: glassy systems \cite{sad_lj} or mean-field models
900: \cite{ktrig,ktrig2,phi4}.
901: In the latter one minimized the pseudo-potential $W$=$|\nabla V|^2$
902: during the dynamic evolution at different temperatures,
903: so introducing a map from equilibrium energy levels to saddles energy levels:
904: ${\cal{M}}\!: v\to {\cal{M}}(v)\equiv v_s$.
905: Due to the peculiarity of the present models,
906: where the saddle point is ``infinitely'' far from each equilibrium configuration,
907: we decided to apply the $W$ minimization method
908: in a two steps procedure:
909: {\em i}) first we minimized the $W_{int}$ quantity defined using the interaction
910: potential part of $V$, $W_{int}$=$|\nabla V_{int}|^2$,
911: where $V_{int}$=$\sum_{i=1}^{N} \frac{K}{2} (q_{i+1}-q_i)^2$;
912: {\em ii}) then we minimized the $W_p$ defined using the on site
913: potential $W_{p}$=$|\nabla V_{p}^{(2,3)}|^2$.
914: This procedure ensures that the point reached is a true stationary point,
915: i.e. the minimum or the saddle.
916: Obviously, this is a quite arbitrary definition of basins of attraction
917: of stationary points. As said in the introduction, the robustness of the
918: results with respect to the possible choices of definition of a saddle
919: basin of attraction is still an open problem.
920:
921: In Fig. \ref{fig_7} the temperature dependence of the energy $v_s$
922: (open symbols) of underlying saddles
923: is shown for the case $\xi$=$0.05$ (upper panel)
924: and $\xi$=$0.5$ (lower panel) in the PB model.
925: The remarkable fact is that at $T_c$ (vertical full line in Fig. \ref{fig_7})
926: the identity $v_s$=$v_{\theta}$ holds.
927: The map ${\cal M}(v)$ is shown for PB model (open symbols)
928: in Fig. \ref{fig_9}
929: for the two cases $\xi$=$0.05$ (upper panel) and $\xi$=$0.5$ (lower
930: panel). One observe that, as before pointed out, one has
931: ${\cal M}(v_c)$=$v_{\theta}$ for both $\xi$ values.
932: The fact that $v_s(T)$ in Fig.~\ref{fig_7}, as well as ${\cal M}(v)$
933: in Fig.~\ref{fig_9}, has a ``smooth'' transition between its low $T$ (or $v$)
934: and high $T$ (high $v$) regions is most likely due to a finite size effect
935: ($N$=$500$ here) and both $v_s(T)$ and ${\cal M}(v)$ will probably tend towards
936: a step function in the thermodynamic limit.
937: The previous finding
938: indicates that the relevant quantity to consider when we are
939: looking for topological changes related to a phase transition is
940: the underlying stationary point energy, obtained trough a map from
941: the critical level $v_c$.
942: It is worth noting that the map ${\cal M}$ is constant
943: (${\cal M}(v)$=$v_{\theta}$) for a broad range of values,
944: also below $v_c$,
945: at variance with other cases where around the transition point
946: the properties of visited saddles change \cite{XY,ktrig,ktrig2,phi4}.
947: One can conjecture that the flatness of ${\cal M}(v)$
948: is a pathology of these one-dimensional models,
949: that have a number of stationary points that is not extensive in $N$
950: (actually there are only $2$ stationary points).
951:
952: \begin{figure}[t]
953: %%\epsfysize= 10 truecm
954: %%\begin{center}
955: \includegraphics[width=.5\textwidth]{fig9.eps}
956: %%\end{center}
957: \caption{
958: Map ${\cal M}\!: v \to v_s$ defined minimizing the pseudopotential $W$=$|\nabla V|^2$
959: in the PB and SPB models for $\xi$=$0.05$ (upper panel) and
960: $\xi$=$0.5$ (lower panel).
961: Also reported are the corresponding $v_{\theta}$ (dotted lines) and $v_c$ (dot-dashed lines)
962: for $\xi$=$0.05$ ($v_{\theta}$=$0$, $v_c\simeq0.61$) and $\xi$=$0.5$ ($v_{\theta}$=$0$, $v_c\simeq1.59$),
963: evidencing the identity ${\cal M}(v_c)$=$v_{\theta}$.
964: }
965: \label{fig_9}
966: \end{figure}%%
967:
968:
969: In Fig \ref{fig_8} we report the same quantities $v_s$ as before (open symbols),
970: now for the SPB model, with $\xi$=$0.05$ (upper panel) and $\xi$=$0.5$
971: (lower panel). In this case no phase transition is present, and indeed the
972: topological singularity is never visited, $v_s(T)<v_\theta$
973: for each finite temperature ($T<\infty$).
974:
975:
976: \subsection{Burkhardt model}
977:
978: To apply the analysis of the previous section also to the Burkhardt model, one has
979: to find a suitable definition of ``saddles'' and of ``basin of attraction of a saddle''
980: for a discontinuous potential.
981: One possibility is the following: we first minimize the interaction potential
982: $V_{int} = \sum_{i=1}^N K |q_{i+1}-q_i|$, which is equivalent
983: to put all the $q_i$ equal
984: to the center of mass coordinate $\bar q=N^{-1}\sum_i q_i$.
985: If $\bar q$ lies in the well of the potential, {\it i.e.} $\bar q \in [L,L+R]$, we
986: will associate the ``minimum'' to the initial configuration, otherwise we will
987: associate it to the ``saddle'' (we use this terminology by analogy with the PB model).
988: It is clear that the average energy of the ``underlying saddles'' is simply the
989: average of the on-site energy of the center of mass coordinate,
990: \begin{equation}
991: v_s(T)=\langle V^{(1)}_p(\bar q) \rangle_T \ .
992: \label{cm}
993: \end{equation}
994: In the thermodynamic limit the center of mass $\bar q$ is peaked around its mean value
995: and then we can substitute the right hand side of Eq. \ref{cm} with
996: $V^{(1)}_p(\langle \bar q \rangle)$, a quantity that can be explicitly computed.
997: To determine $\langle \bar q \rangle$ we can use the distribution probability
998: $|\phi(x)|^2$, where $\phi(x)$=$e^{-\beta V_{p}(x)/2} \psi(x)$ and $\psi(x)$ is the eigenfunction
999: of the transfer matrix operator corresponding to the maximum eigenvalue \cite{bur} (see Sec. II A).
1000: We note that the saddle energy $v_s(T)$ is a step function, equals to the minimum energy $-U_o$
1001: when $\langle \bar q \rangle$ lies inside the square well and equals to the saddle
1002: energy $0$ otherwise.
1003: The temperature $T_J$ at which the visited ``underlying saddle'' jumps from minimum to saddle
1004: is shown in Fig. \ref{fig_3} (dashed line) as a function of the
1005: parameter $\xi$ for the $\zeta$=$0$ case.
1006: It is worth noting that the temperature $T_J$ lies always below the thermodynamic transition
1007: temperature $T_c$ (in analogy with the PB model, see Fig. \ref{fig_7}).
1008: The same happens for all values of $\xi$.
1009: Therefore, also for the Burkhardt case, at the transition temperature $T_c$
1010: the ``underlying saddles'' lie at an energy equal to the topological discontinuity energy
1011: $v_{\theta}$, i.e. ${\cal M}(v_c)$=$v_{\theta}$.
1012:
1013:
1014:
1015:
1016:
1017: %%%%%
1018: \section{Conclusions} %{\em Conclusions.---}
1019:
1020: Studying two particular one dimensional models discussed in the
1021: recent literature \cite{kastner,gri_mos}
1022: (Burkhardt model in the confining and non-confining version,
1023: Peyrard-Bishop model and its non-confining counterpart),
1024: we have focused on the relationship between phase transitions and
1025: topological changes,
1026: recently proposed in the literature \cite{cccp,fps,cpc,fra_pet}.
1027: In these models, a topological
1028: singularity at a given energy value $v_\theta$(=0) is always found;
1029: however, {\it i)} in the confining version
1030: a phase transition is found but the critical energy is $v_c > v_\theta$
1031: \cite{gri_mos}; {\it ii)} in their non-confining version there is no phase transition
1032: at any finite temperature \cite{kastner}.
1033:
1034: These results generated confusion as {\it i)} was
1035: interpreted as a confirmation of the {\em strong topological hypothesis} of
1036: Pettini {\it et al.} \cite{nota}
1037: while {\it ii)} was considered as an evidence for the
1038: unattainability of a purely topological criterion
1039: for detecting phase transitions,
1040: although demonstrated only for the particular non-confining
1041: one dimensional models.
1042:
1043: Exploiting the concept of ``{\em underlying
1044: stationary points}'' defined through a generalization
1045: of the methods used in the glassy literature
1046: (minimization of the pseudopotential $W$=$|\nabla V|^2$),
1047: we have defined a map ${\cal M}:v\to
1048: v_s$ from energy level $v$ of $V$ to stationary points, with energy
1049: $v_s$.
1050: We have shown that: {\it i)} in the confining case, where the phase
1051: transition is present, one has ${\cal M}(v_c)=v_\theta$, in agreement
1052: with the {\em weak topological hypothesis};
1053: {\it ii)} in the non-confining case, where the phase transition is not
1054: present at finite temperature
1055: (as the transition temperature goes continuously to infinity when
1056: the confining wall is removed) the energy of the underlying saddles is
1057: always {\it below} the topological singularity, i.e. $v_s(T)<v_\theta \, , \, \forall T$;
1058: the singular point $v_\theta$ is indeed visited for $T\rightarrow \infty$,
1059: consistently with the observation that the critical temperature is ``infinite''
1060: in the non-confining case.
1061:
1062: The {\it weak topological hypothesis} appears as a possible
1063: framework to fit the results that recently appeared in the
1064: literature on all the different models investigated so far.
1065: Within this hypothesis three different scenarios are possible:
1066: \begin{enumerate}
1067: \item If there is no topological singularity $v_\theta$, a phase
1068: transition is not possible; this is consistent with the hypothesis
1069: of Pettini {\it et al.}: topological singularities are {\it necessary}
1070: conditions for a phase transition to take place.
1071: \item If there is a topological singularity at energy $v_\theta$,
1072: a phase transition is also
1073: present {\it if and only if} there exist a temperature $T_c$ such that
1074: $v_s(T_c) = v_\theta$
1075: (or equivalently an energy $v_c$ such that ${\cal M}(v_c)=v_\theta$).
1076: \end{enumerate}
1077: The above findings seem to indicate that, at least for the particular models investigated,
1078: a sufficiency criterion for the phase transition to take place requires
1079: the introduction of a statistical measure:
1080: thus, we believe that the statement of Kastner~\cite{kastner}
1081: concerning the unattainability of a purely topological criterion
1082: for detecting phase transitions is indeed correct,
1083: even though in Ref.~\cite{kastner} it has been derived using
1084: a ``borderline'' model (see Section II C).
1085:
1086: %%%%%
1087:
1088: Let us conclude with two remarks:
1089: {\it i)} as already stated, the definition of the map ${\cal M}$ is
1090: not unique, different definitions giving (slightly) different results.
1091: Thus, the {\it weak topological hypothesis} contains in its formulation
1092: an ambiguity and must be regarded only as a {\it practical} tool, at
1093: least at this stage of comprehension;
1094: {\it ii)} nevertheless, we hope that this approach can be of interest
1095: for the numerical investigation of systems of ``mesoscopic'' size
1096: (e.g. proteins and large molecules),
1097: i.e. such that the number of degrees of freedom is not large enough
1098: to allow to detect the presence of a phase transition using
1099: standard techniques.
1100:
1101:
1102:
1103: We thank M.~Pettini and M.~Kastner for helpful comments and suggestions.
1104:
1105: \begin{thebibliography}{99}
1106:
1107: \bibitem{Ma}
1108: See {\it e.g.} S.~K.~Ma, {\it Modern Theory of Critical Phenomena}
1109: (W. A. Benjamin, New York, 1976).
1110:
1111: \bibitem{cccp}
1112: L.~Caiani, L.~Casetti, C.~Clementi, and M.~Pettini,
1113: Phys. Rev. Lett. {\bf 79}, 4361 (1997).
1114:
1115: \bibitem{fps}
1116: R.~Franzosi, M.~Pettini, and M.~Spinelli,
1117: Phys. Rev. Lett. {\bf 84}, 2774 (2000).
1118:
1119: \bibitem{cpc}
1120: L.~Casetti, M.~Pettini, and E.G.D.~Cohen,
1121: Phys. Rep. {\bf 337}, 237 (2000).
1122:
1123: \bibitem{fra_pet}
1124: R.~Franzosi and M.~Pettini,
1125: Phys. Rev. Lett. {\bf 92}, 060601 (2004).
1126:
1127: \bibitem{morse}
1128: M.W.~Hirsch, {\em Differential topology},
1129: Springer, New York (1976).
1130:
1131:
1132: \bibitem{XY}
1133: L.~Casetti, M.~Pettini, and E.G.D.~Cohen,
1134: J. Stat. Phys. {\bf 111}, 1091 (2003).
1135:
1136: \bibitem{ktrig}
1137: L.~Angelani, L.~Casetti, M.~Pettini, G.~Ruocco, and F.~Zamponi,
1138: Europhys. Lett. {\bf 62}, 775 (2003).
1139:
1140: \bibitem{ktrig2}
1141: F.~Zamponi, L.~Angelani, L.F.~Cugliandolo, J.~Kurchan, and G.~Ruocco,
1142: J. Phys. A: Math. Gen. {\bf 36}, 8565 (2003).
1143:
1144: \bibitem{phi4}
1145: A.~Andronico, L.~Angelani, G.~Ruocco, and F.~Zamponi,
1146: Phys.~Rev.~E {\bf 70}, 041101 (2004);
1147: D.~A.~Garanin, R.~Schilling, and A.~Scala,
1148: Phys.~Rev.~E {\bf 70}, 036125 (2004).
1149:
1150: \bibitem{kastner}
1151: M.~Kastner,
1152: Phys. Rev. Lett. {\bf 93}, 150601 (2004).
1153:
1154: \bibitem{gri_mos}
1155: P.~Grinza and A.~Mossa,
1156: Phys. Rev. Lett. {\bf 92}, 158102 (2004).
1157:
1158: \bibitem{kastner2}
1159: M.~Kastner,
1160: preprint cond-mat/0412199.
1161:
1162: \bibitem{teix}
1163: A.C.R.~ Teixeira and D.A.~Stariolo
1164: Phys. Rev. E {\bf 70}, 016113 (2004).
1165:
1166: \bibitem{sad_lj}
1167: L.~Angelani {\em et al.},
1168: Phys. Rev. Lett. {\bf 85}, 5356 (2000);
1169: J. Chem. Phys. {\bf 116}, 10297 (2002);
1170: J. Chem. Phys. {\bf 119}, 2120 (2003);
1171: J. Chem. Phys. {\bf 121}, 7533 (2004).
1172:
1173: \bibitem{sad_cav}
1174: K.~Broderix, K.K.~Bhattacharya, A.~Cavagna, A.~Zippelius, and I.~Giardina,
1175: Phys. Rev. Lett. {\bf 85}, 5360 (2000).
1176:
1177: \bibitem{jcp_comment}
1178: J.P.K.~Doye and D.J.~Wales,
1179: J. Chem. Phys. {\bf 118}, 5263 (2003);
1180: L.~Angelani, R.Di~Leonardo, G.~Ruocco, A.~Scala, and F.~Sciortino,
1181: J. Chem. Phys. {\bf 118}, 5265 (2003).
1182:
1183: \bibitem{cavagna}
1184: A.~Cavagna, I.~Giardina,, and G.~Parisi,
1185: Phys. Rev. B {\bf 57}, 11251 (1998);
1186: J. Phys. A {\bf 34}, 5317 (2001).
1187:
1188: \bibitem{VMF}
1189: See e.g. J.~P.~Bouchaud and G.~Biroli,
1190: cond-mat/0406317 for a recent review.
1191:
1192:
1193: \bibitem{pey}
1194: M.~Peyrard and A.R.~Bishop,
1195: Phys. Rev. Lett. {\bf 62}, 2755 (1989).
1196:
1197: \bibitem{pey1}
1198: M.~Peyrard,
1199: Nonlinearity {\bf 17}, R1 (2004).
1200:
1201: \bibitem{nota}
1202: The authors give support to the ``{\em topological hypothesis}'',
1203: without evidencing the difference between the two energies
1204: $v_c \neq v_{\theta}$.
1205:
1206: \bibitem{bur}
1207: T.W.~Burkhardt,
1208: J. Phys. A: Math. Gen. {\bf 14}, L63 (1981).
1209:
1210: \bibitem{QM}
1211: L.D.~Landau and E.M.~ Lifshitz, {\it Quantum Mechanics}
1212: (Pergamon Press, New York, 1977).
1213:
1214: \bibitem{pey2}
1215: T.~Dauxois, N.~Theodorakopoulos, and M.~Peyrard,
1216: J. Stat. Phys. {\bf 107}, 869 (2002).
1217:
1218: \bibitem{theo}
1219: N.~Theodorakopoulos,
1220: Phys. Rev. E {\bf 68}, 026109 (2003).
1221:
1222: \bibitem{cuesta}
1223: J.A.~Cuesta and A.~Sanchez,
1224: J. Stat. Phys. {\bf 115}, 869 (2004).
1225:
1226:
1227: \end{thebibliography}
1228:
1229: \end{document}
1230: