arxiv:nucl-th/ v2 26 Feb 2002

Similar documents
Metastability in the Hamiltonian Mean Field model and Kuramoto model

Non-Gaussian equilibrium in a long-range Hamiltonian system

Slow dynamics in systems with long-range interactions

Dynamics and Statistical Mechanics of (Hamiltonian) many body systems with long-range interactions. A. Rapisarda

Quasi-stationary-states in Hamiltonian mean-field dynamics

Algebraic Correlation Function and Anomalous Diffusion in the HMF model

Challenges of finite systems stat. phys.

Thermodynamics of nuclei in thermal contact

arxiv: v2 [cond-mat.stat-mech] 31 Dec 2014

Recent progress in the study of long-range interactions

High order corrections to density and temperature of fermions and bosons from quantum fluctuations and the CoMD-α Model

arxiv:cond-mat/ v2 [cond-mat.stat-mech] 8 Sep 1999

Invaded cluster dynamics for frustrated models

1 Notes on the Statistical Mechanics of Systems with Long-Range Interactions

Nuclear Science Research Group

Global Optimization, Generalized Canonical Ensembles, and Universal Ensemble Equivalence

Microcanonical scaling in small systems arxiv:cond-mat/ v1 [cond-mat.stat-mech] 3 Jun 2004

Environmental Atmospheric Turbulence at Florence Airport

Density and temperature of fermions and bosons from quantum fluctuations

arxiv: v1 [cond-mat.stat-mech] 19 Feb 2016

Making thermodynamic functions of nanosystems intensive.

Gouy-Stodola Theorem as a variational. principle for open systems.

MD Thermodynamics. Lecture 12 3/26/18. Harvard SEAS AP 275 Atomistic Modeling of Materials Boris Kozinsky

1. Thermodynamics 1.1. A macroscopic view of matter

arxiv:cond-mat/ v1 8 Jan 2004

Electrical Transport in Nanoscale Systems

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 24 Jul 2001

arxiv:nucl-th/ v1 12 Jun 2000

(# = %(& )(* +,(- Closed system, well-defined energy (or e.g. E± E/2): Microcanonical ensemble

arxiv: v1 [nlin.cd] 22 Feb 2011

arxiv: v1 [cond-mat.stat-mech] 7 Nov 2017

Dissipative nuclear dynamics

From time series to superstatistics

arxiv: v4 [cond-mat.stat-mech] 9 Jun 2017

arxiv:cond-mat/ v1 [cond-mat.other] 4 Aug 2004

arxiv:cond-mat/ v1 29 Dec 1996

Microscopic Deterministic Dynamics and Persistence Exponent arxiv:cond-mat/ v1 [cond-mat.stat-mech] 22 Sep 1999

Property of Tsallis entropy and principle of entropy increase

CURRICULUM VITAE. Affiliation: National Institute of Physics and Nuclear Engineering, Bucharest, POB-MG 6, Romania

Statistical mechanics in the extended Gaussian ensemble

NPTEL

Istituto Nazionale Fisica Nucleare

Isoscaling, isobaric yield ratio and the symmetry energy: interpretation of the results with SMM

caloric curve of King models

Critical like behavior in the lattice gas model

PHYSICAL REVIEW LETTERS

Energy conservation and the prevalence of power distributions

Basics of Statistical Mechanics

arxiv:cond-mat/ v1 [cond-mat.mtrl-sci] 6 Jun 2001

Physica A. Thermodynamics and dynamics of systems with long-range interactions

Bimodal pattern in the fragmentation of Au quasi-projectiles

arxiv:nucl-ex/ v1 13 Oct 2003

Perturbation theory for the dynamics of mean-field systems

Generalized Entropy Composition with Different q Indices: A Trial

arxiv:astro-ph/ v1 15 Mar 2002

Power law distribution of Rényi entropy for equilibrium systems having nonadditive energy

On the Asymptotic Convergence. of the Transient and Steady State Fluctuation Theorems. Gary Ayton and Denis J. Evans. Research School Of Chemistry

PHYSICAL REVIEW LETTERS

arxiv:nucl-ex/ v1 27 May 2002

arxiv: v1 [nucl-th] 14 Mar 2008

How the Nonextensivity Parameter Affects Energy Fluctuations. (Received 10 October 2013, Accepted 7 February 2014)

PHYS3113, 3d year Statistical Mechanics Tutorial problems. Tutorial 1, Microcanonical, Canonical and Grand Canonical Distributions

arxiv: v1 [cond-mat.stat-mech] 6 Mar 2008

Phase transitions in the Potts spin-glass model

PHY 6500 Thermal and Statistical Physics - Fall 2017

Power law distribution of Rényi entropy for equilibrium systems having nonadditive energy

Isospin and symmetry energy effects on nuclear fragment distributions in liquid-gas type phase transition region

Particle-number projection in finite-temperature mean-field approximations to level densities

Boundary Dissipation in a Driven Hard Disk System

Javier Junquera. Statistical mechanics

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 25 Jul 2003

Bimodality and Scaling-Signs of phase transition in Nuclear Multifragmentation? B.Grabez Institute of Physics Zemun

Basics of Statistical Mechanics

arxiv:cond-mat/ v1 10 Aug 2002

PROTON-PROTON FEMTOSCOPY AND ACCESS TO DYNAMICAL SOURCES AT INTERMEDIATE ENERGIES

Renormalization Group: non perturbative aspects and applications in statistical and solid state physics.

SMR/ International Workshop on QCD at Cosmic Energies III. 28 May - 1 June, Lecture Notes. E. Zabrodin University of Oslo Oslo, Norway

The liquid-gas phase transition and the caloric curve of nuclear matter

Statistical Mechanics

PHASE TRANSITIONS. A. Budzanowski

Lecture 5: Temperature, Adiabatic Processes

arxiv: v2 [cond-mat.stat-mech] 27 Jun 2016

UNIVERSITY OF OSLO FACULTY OF MATHEMATICS AND NATURAL SCIENCES

Two-dimensional dissipative maps at chaos threshold:sensitivity to initial conditions and relaxation dynamics

arxiv: v1 [astro-ph.sr] 30 Mar 2019

Nonextensivity of hadronic systems

Rensselaer Polytechnic Institute, 110 8th Street, Troy, NY , USA. Florida State University, Tallahassee, Florida , USA

Ensemble equivalence for non-extensive thermostatistics

Nuclear Science Research Group

arxiv: v3 [nlin.cd] 10 Nov 2016

arxiv:nucl-th/ v2 28 Nov 2000

Lee Yang zeros and the Ising model on the Sierpinski gasket

arxiv: v1 [cond-mat.stat-mech] 21 Oct 2007

Quantum Chaos as a Practical Tool in Many-Body Physics

Spin Peierls Effect in Spin Polarization of Fractional Quantum Hall States. Surface Science (2) P.1040-P.1046

Quasi-Stationary Simulation: the Subcritical Contact Process

arxiv: v1 [cond-mat.stat-mech] 3 Apr 2007

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 8 Oct 1996

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 21 Jan 2004

Based on work in progress in collaboration with: F. Scardina, S. Plumari and V. Greco

Transcription:

NEGAIVE SPECIFIC HEA IN OU-OF-EQUILIBRIUM NONEXENSIVE SYSEMS A. Rapisarda and V. Latora Dipartimento di Fisica e Astronomia, Università di Catania and INFN Sezione di Catania, Corso Italia 57, I-95129 Catania, Italy alk presented at the International Workshop on Multifragmentation INFN-LNS, Catania 28 th November - 1 st December 2001 Corresponding author, e-mail: andrea.rapisarda@ct.infn.it (February 8, 2008) arxiv:nucl-th/0202075v2 26 Feb 2002 We discuss the occurrence of negative specific heat in a nonextensive system which has an equilibrium second-order phase transition. he specific heat is negative only in a transient regime before equilibration, in correspondence to long-lasting metastable states. For these states standard equilibrium Bolzmann-Gibbs thermodynamics does not apply and the system shows non-gaussian velocity distributions, anomalous diffusion and correlation in phase space. Similar results have recently been found also in several other nonextensive systems, supporting the general validity of this scenario. hese models seem also to support the conjecture that a nonexstensive statistical formalism, like the one proposed by sallis, should be applied in such cases. he theoretical scenario is not completely clear yet, but there are already many strong theoretical indications suggesting that, it can be wrong to consider the observation of an experimental negative specific heat as an unique and unambiguous signature of a standard equilibrium first-order phase transition. I. INRODUCION In the last years there has been an increasing interest in phase transitions occurring in systems of finite size. Nuclear multifragmentation phase transition is only one of the most interesting examples [1]. here has also been a lively debate in the literature on how to detect unambiguously the occurrence of a phase transition in small systems. An interesting proposal has been the measurement of negative specific heat [1,2]. Recently several experiments have found this signal not only in nuclei [3 ], but also in atomic clusters [4]. he point we want to stress here is that in standard Bolzmann-Gibbs (BG) thermodynamics, which is based on extensive systems, i.e. systems for which energy and the entropy are proportional to N, the specific heat can never be negative [5]. However several theoretical pioneering investigations have demonstrated that such a property is violated, in the microcanonical ensemble, for nonextensive systems [1,6,7], i.e. for long-range interactions (Coulomb and gravitational forces), but also for finite-size systems with short-range interaction. From these studies, the application of BG statistics seems to remain valid for some nonextensive systems, if the microcanonical ensemble is used, although an inequivalence of ensembles remains also in the thermodynamic limit [6]. A different approach to tackle nonextensive systems has been proposed by sallis in 1988 [8]. sallis generalized statistics, which contains the BG formalism as a particular case, has found since then an increasing and successful amount of applications in several fields. he main validity of this approach has been found for physical situations where out-of-equilibrium phenomena, long-range corre-lations, long-term memory effects and anomalous fluctuations are observed [9]. Recently the appearance of a negative specific heat has been numerically observed in correspondence to a metastable long-lasting regime, where sallis formalism and not the BG one seems to apply [10-14]. Hot nuclear compound systems, formed in high energy heavy-ion reactions, are nonextensive systems, therefore the above mentioned results suggest that one should be very careful in applying standard equilibrium BG thermodynamics in such a fast dynamical phenomenon like multi-fragmentation. Even a small deviation from equilibrium could prevent the application of the BG formalism. In this short contribution, we briefly discuss some recent numerical simulations of a nonextensive system, the Hamiltonian Mean Field (HMF) model, which illustrate the occurrence of a negative specific only in a transient out-of-equilibrium regime [10]. his model is not a peculiar exceptional example and its anomalous behavior should be a strong warning against a simple and straightforward extrapolation of an equilibrium phase transition signature, in finite nonextensive systems, when measuring a negative specific heat. he paper is organized as follows: we introduce the model in section II, numerical simulations are presented in section III and conclusion are drawn in section IV. 1

II. HE MODEL HAMILONIAN Our arguments are based on the following system of N fully-coupled classical rotators, whose Hamiltonian is given by [15] H = K + V = N i=1 p i 2 2 + 1 2N N [1 cos(θ i θ j )], (1) where θ i is the the angle and p i is the corresponding conjugate momentum. he canonical analytical solution of the model predicts a second-order phase transition, whose order parameter is the total magnetization M, given by the averaged sum over the spin vectors m i = (cosθ i, sinθ i ), i.e. i,j=1 M = 1 N N m i. (2) i=1 For small energies the absolute value of the magnetization is M 1, while for energies greater than the critical value c = 0.5 (U c = E c /N = 0.75), we have M 0. he equilibrium caloric curve has been derived in ref [15] and is given by the expression for the energy density U = E/N = 2 + 1 2 ( 1 M 2 ), (3) being the temperature. So it is very interesting to compare the exact canonical solution (3) with numerical microcanonical simulations. Moreover, due to the long-range nature of the interaction, this system is nonextensive. hus the model is of extreme interest also from a fundamental point of view, for statistical mechanics, whose standard Bolzmann-Gibbs text-books formulation is based on the extensive and thus additive properties of energy and entropy. For these reasons the model has been intensively and extensively studied in the last years [10-11,15,16]. It has also been slightly modified to understand the influence of the range of the interaction [17]. Finite-size effects, chaotic dynamics and superdiffusion have been investigated in detail for the HMF model [10-11,15,16]. But an important point which has also been particularly studied is the relaxation to equilibrium: in fact, when out-of-equilibrium initial conditions are considered, the model presents a very slow and anomalous dynamical behavior, in a energy range below the critical point. In the following we will focus our attention mainly to this out-of-equilibrium regime before relaxation, discussing its eventual relevance for nuclear multifragmentation experiments. III. NUMERICAL SIMULAIONS We present in this section some numerical simulations which show the interesting transient dynamical behavior of the HMF model. In fig.1a we plot the equilibrium caloric curve (3) (full line) in correspondence of two numerical results, for N=10000 and 100000 (open circles and squares), before complete equilibration, corresponding to a time t=1200. he time step used was 0.2. he points which mostly disagree with the equilibrium curve are in the energy range 0.5 U 0.75, below the critical energy density, indicated as a dashed vertical line. In this region, a complete equilibration is generally obtained only after 10 6-10 7 time-steps, according to the size of the system, see ref.[16] for more technical details about the integration scheme. In order to study the dynamics of this slow relaxation, we fix a particular energy density, i.e. U=0.69, and we plot in fig.1b the quantity 2 < K > /N as a function of time, < K > being the average kinetic energy. he simulations display a plateau for a long transient time which does not correspond to the equilibrium value eq = 76, also reported as a dashed red line. he system is trapped in a quasi-stationary state (QSS), whose whose lifetime increases with N [10]. he quantity 2 < K > /N coincides with the temperature if a stationary situation exists, thus we can refer to the plateau values as the N-dependent temperatures of the quasi-stationary states (QSS). he relaxation is reached, as the plot shows, only after a long time, which increases linearly with N. Also the QSS temperature depends on the size and converge to the infinite size value = 0.38 as a power-law[10]. 2

0.6 a) Boltzmann-Gibbs equilibrium N=100000 time=1200 N=10000 time=1200 2<K>/N 0.2 0.55 0.50 5 0 0 0 0.2 0.6 0.8 U b) U=0.69 WBIC eq =76 N=500 N=1000 N=10000 N=100000 0.35 inf =0.38 10 1 10 2 10 3 10 4 10 5 10 6 time-t 0 FIG. 1. (a) We show the temperature, calculated by taking the average kinetic energy per particle 2 < K > /N, and the energy density U = 0.69 for HMF systems of different sizes N=10000 (open circles) and N=100000 (squares). For comparison the equilibrium caloric curve is also shown as full line. he numerical simulations were initialized considering water bag initial conditions. he points are taken after a short time, t=1200, see text. he dashed line indicates the critical point. (b) We show the temperature time evolution for different N values. he initial part has been subtracted (t 0 = 100). We report also the equilibrium temperature eq (upper red dashed line) and the temperature to which the QSS tend for infinite size (bottom black dashed line). 3

2<K>/N Initial evolution U=0.69 0.6 N=500 N=1000 N=5000 N=10000 0.5 eq =76 0 20 40 60 80 100 time FIG. 2. Initial time evolution of the quantity 2 < K > /N for the HMF model. he numerical simulations correspond to different N sizes at U = 0.69. In fig.2 we show the initial time evolution of 2 < K > /N to display the fast collapse of the initial condition to the metastable regime. We start our simulations by considering the so-called water-bag initial conditions, i.e. putting all the angles at zero and distributing the total energy uniformly over the momenta. he figure shows that there is a rapid evolution, which is not size-dependent, towards the QSS state. In ref. [10], we have also checked that these simulations are not affected by the numerical accuracy of the integration scheme used. his is certainly true in the range explored, with a relative error 10 7 < E/E < 10 3, and demonstrates the robustness of these metastable QSS against small perturbations. At this point, it is interesting to calculate the specific heat in correspondence of this slow relaxation in the energy region below the critical point. he specific heat can be calculated from the fluctuations of kinetic energy, by using the microcanonical formulas derived in the 60 s by Lebowitz, Percus and Verlet (LPV formula) [18], i.e. C V = 1 2 [ 1 2 ( ) ] 2 1 Σ (4) being the microcanonical temperature and Σ the kinetic energy microcanonical fluctuation. A second alternative is another formula which was derived more recently by Pearson, Haliciouglu and iller (PH formula) [19]. he latter should be more precise, because it takes into account finite-size effects and is given by C V = [ 2 < K >< K > 1 +N(1 < K >< K > 1 ) ] 1. (5) 4

100 50 LPV formula PH formula Equilibrium C v =2.34 10 8 6 4 N=500 U=0.69 C V 2 0 100000 200000 0-50 0 200000 400000 time FIG. 3. Specific heat calculations for the HMF model as a function of time in the case N=500 and U=0.69. he two different formulas discussed in the text, LPV formula (4) and PH formula (5), are compared in the plot and in the inset magnification. he figure shows a very similar time evolution and a nice convergence for long times to the correct positive equilibrium value, indicated by the dashed line. We report in fig.3 the specific heat calculated according to the eqs. (4) and (5) for N=500 and U=0.69. he figure shows a very similar time evolution for both formulas: after some oscillations, in which the specific heat is negative, both numerical simulations converge towards the same correct equilibrium positive value [16], also indicated as a dashed line. Let us now focus our attention on the transient regime, where the specific heat is negative. We have found that, in the transient QSS regime, the system does not show a Gaussian velocity probability distribution, while for longer integration times, when the system finally relaxes 5

Pdf 1 0.1 (a) U=0.69 N=500 Pdf 1 0.1 (b) 0.01 0.01 0.001-3 -2-1 0 1 2 3 p 0.001-3 -2-1 0 1 2 3 p 0.5 (c) 0.5 (d) 0.5 0.6 0.7 0.8 U 0.5 0.6 0.7 0.8 U time QSS regime BG regime FIG. 4. N=500. Caloric curves and velocity pdfs for the HMF model in the QSS regime and in the BG equilibrium regime for to the equilibrium temperature, the velocity probability distribution (pdf) is perfectly Gaussian [10]. his result is nicely illustrated in fig.4. here we plot the numerical simulations for the caloric curve and the velocity pdfs both in the transient case, yellow circles in panels (c) and (a), and in the equilibrium case, white circles in panels (d) and (b). A back-bending of the HMF caloric curve in the transient regime clearly coincides with a non-gaussian shape of the velocity pdf. In particular pdf tails are missing and slow decaying velocity correlations are present. One can also show that the velocity pdfs are frozen in this anomalous distribution for all the duration of the metastable regime. his result can be easily understood since the force, which each spin feels, is almost zero in the QSS regime. In fact the force on the single spin is given by the formula F i = M x sinθ i + M y cosθ i, where M x and M y are the components of the magnetization vector. hus being M 2 = 0 in the QSS regime, it follows that also F i = 0. he system is attracted by the QSS state and then remains in a frozen state. Why it then relaxes to the BG equilibrium? he answer is simple. If the system size is finite, the magnetization is not exactly zero, a small noise (O 1 N ) always exists and this is responsible for the final relaxation to the BG equilibrium regime. he bigger the size, the smaller is the noise and thus the longer is the lifetime of the QSS state as numerically found [10]. his fact implies the interesting result that, if one inverts the order of the limits, i.e. takes first the infinite size limit instead of the infinite time limit, the noise is perfectly zero and the system remains trapped in the QSS state for ever. In [10] we have shown that the 6

formalism proposed by sallis [8,9] seems to explain the shape of the velocity pdfs. 4 CONCLUSIONS Summarizing,we have shown an example of a nonextensive system where a negative specific heat is found in correspondence to quasi-stationary states (QSS) and non-gaussian velocity pdfs. his result which has been recently confirmed also in other long-range interacting models such as self-gravitating systems [7,12] and modified Lennard- Jones potentials [13] is due to the nonextensivity nature of the system into exam [14]. his fact is a serious warning for a straightforward claiming of a standard equilibrium first-order phase transition in nuclear fragmenting systems. Although some sort of equilibration seems to be reached in multifragmentation, it is not certain whether this corresponds to a complete relaxation. We stress the fact that the temperature deviation from its equilibrium value, we have discussed so far, is only of the order of 10%! In this respect more detailed investigations should be done in order to further clarify the general theoretical scenario for nonextensive systems. But also more precise experimental data are very welcome and could be extremely useful to understand deeper the intriguing nature of phase transitions in finite systems, a field surely at the frontier of modern statistical mechanics. REFERENCES [1] D.H.E. Gross, Microcanonical thermodynamics: phase transitions in small systems, Lecture Notes in Physics, Springer-Verlag Heidelberg 2001 and refs therein. See also the proceedings of this workshop. [2] P. Chomaz and F. Gulminelli, Nucl. Phys. A 647 (1999) 153 [3] M.D Agostino et al, Phys. Lett. B A473 (2000) 219. [4] M. Schmidt et al. Phys. Rev. Lett. 86 (2001) 1191. [5] K. Huang, Statistical mechanics, Wiley (1987). [6] J. Barrè, D. Mukamel, S. Ruffo, Phys. Rev. Lett. 87 (2001) 030601. [7] A. orcini, M. Antoni, Phys. Rev. E 89 (1999) 2746. [8] C. sallis, J.Stat. Phys. 52 (1988) 479. [9] For an updated review of this generalized statistics see the proceedings of the conference NEX2001 published in Physica A 305 (2002). An updated reference list is also available at http://tsallis.cat.cbpf.br/biblio.htm [10]V.Latora and A. Rapisarda, Nucl. Phys. A 681 (2001) 406c; V. Latora, A. Rapisarda and C. sallis, Phys. Rev. E 64 (2001) 056134 and Physica A 305 (2002) 129. [11] C.sallis, B.J.C. Cabral, A. Rapisarda and V. Latora, [cond-mat/0112266] submitted to Phys. Rev. Lett. [12] Sota et al, Phys. Rev. E 64 (2001) 05613; A. aruya, M. Sakagami, Physica A (2002) in press [cond-mat/0107494]. [13] E.P. Borges and C. sallis, Physica A 305 (2002) 148. [14] A. Campa, A. Giansanti, D. Moroni, Physica A 305 (2002) 137. [15] M. Antoni and S. Ruffo Phys. Rev. E 52 (1995) 2361. [16] V. Latora, A. Rapisarda and S. Ruffo, Phys. Rev. Lett. 80 (1998) 698, Physica D 131 (1999) 38 and Phys. Rev. Lett. 83 (1999) 2104. [17] C. Anteneodo and C. sallis, Phys. Rev. Lett. 80 (1998)5313; A. Campa, A. Giansanti, D. Moroni and C.sallis, Phys. Lett. A 286 (2001) 251. [18] J.L. Lebowitz, J.K. Percus and L. Verlet, Phys. Rev. A 153 (1967) 250. [19] E.M. Pearson,. Halicioglu and W.A. iller Phys. Rev. A 32 (1985) 3030. 7