Spectral Statistics of Instantaneous Normal Modes in Liquids and Random Matrices

Similar documents
arxiv:cond-mat/ v1 [cond-mat.stat-mech] 20 Apr 2004

On the Speed of Quantum Computers with Finite Size Clocks

Effect of Unfolding on the Spectral Statistics of Adjacency Matrices of Complex Networks

Introduction to Theory of Mesoscopic Systems

(Anti-)Stable Points and the Dynamics of Extended Systems

Size Effect of Diagonal Random Matrices

Random matrix analysis of complex networks

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 13 Mar 2003

ORIGINS. E.P. Wigner, Conference on Neutron Physics by Time of Flight, November 1956

Renormalization Group Analysis of the Small-World Network Model

Spectral Fluctuations in A=32 Nuclei Using the Framework of the Nuclear Shell Model

Anderson Localization Looking Forward

Misleading signatures of quantum chaos

How Different is the Dynamics of a Lennard-Jones Binary Fluid from One-Component Lennard-Jones Fluid? 1

Population Annealing: An Effective Monte Carlo Method for Rough Free Energy Landscapes

Eigenvectors under a generic perturbation: non-perturbative results from the random matrix approach

Crossover to potential energy landscape dominated dynamics in a model glass-forming liquid

1 Intro to RMT (Gene)

Spectral properties of random geometric graphs

arxiv:physics/ v1 [physics.med-ph] 8 Jan 2003

Universality for random matrices and log-gases

Causal Effects for Prediction and Deliberative Decision Making of Embodied Systems

Some Polyomino Tilings of the Plane

Experimental and theoretical aspects of quantum chaos

arxiv: v2 [cond-mat.dis-nn] 9 Feb 2011

Statistical Mechanics of Jamming

Poisson statistics in the high temperature QCD Dirac spectrum

Edited by H. Eugene Stanley, Boston University, Boston, MA, and approved January 9, 2009 (received for review November 6, 2008)

arxiv: v1 [hep-lat] 30 Oct 2014

Breit-Wigner to Gaussian transition in strength functions

Random Matrix: From Wigner to Quantum Chaos

The Bootstrap is Inconsistent with Probability Theory

Potentially useful reading Sakurai and Napolitano, sections 1.5 (Rotation), Schumacher & Westmoreland chapter 2

Spacing distributions for rhombus billiards

On Irreversible Radiation Processes

arxiv: v1 [cond-mat.soft] 20 Jun 2008

Non-equilibrium Dynamics of One-dimensional Many-body Quantum Systems. Jonathan Karp

From the mesoscopic to microscopic scale in random matrix theory

Spectra of Large Random Stochastic Matrices & Relaxation in Complex Systems

Length Scales Related to Alpha and Beta Relaxation in Glass Forming Liquids

Theory of Mesoscopic Systems

Modeling the Free Energy Landscape for Janus Particle Self-Assembly in the Gas Phase. Andy Long Kridsanaphong Limtragool

Intrinsic Quantum Computation

Sweep from Superfluid to Mottphase in the Bose-Hubbard model p.1/14

LEVEL REPULSION IN INTEGRABLE SYSTEMS

Relaxation processes and entropic traps in the Backgammon model

Recent results in quantum chaos and its applications to nuclei and particles

TOP EIGENVALUE OF CAUCHY RANDOM MATRICES

Solving the Schrödinger equation for the Sherrington Kirkpatrick model in a transverse field

TitleQuantum Chaos in Generic Systems.

Eigenvalue Density for Random Covariance Matrices from a 2 2 Partitioned GOE

Emergence of chaotic scattering in ultracold lanthanides.

Pre-yield non-affine fluctuations and a hidden critical point in strained crystals

Physics on Random Graphs

A Multi-Level Lorentzian Analysis of the Basic Structures of the Daily DJIA

Spectral statistics of random geometric graphs arxiv: v2 [physics.soc-ph] 6 Jun 2017

A Microscopically-Based, Global Landscape Perspective on Slow Dynamics in. Condensed-Matter

Thermodynamic and structural aspects of the potential energy surface of simulated water

Universality of local spectral statistics of random matrices

Melting line of the Lennard-Jones system, infinite size, and full potential

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

12. Spectral diffusion

Can imaginary instantaneous normal mode frequencies predict barriers to self-diffusion?

Large Petermann factor in chaotic cavities with many scattering channels

Nonextensive Aspects of Self- Organized Scale-Free Gas-Like Networks

Slightly off-equilibrium dynamics

On Information and Sufficiency

Is Quantum Mechanics Chaotic? Steven Anlage

Lectures about Python, useful both for beginners and experts, can be found at (

Dynamics of Lennard-Jones clusters: A characterization of the activation-relaxation technique

COMPLEX HERMITE POLYNOMIALS: FROM THE SEMI-CIRCULAR LAW TO THE CIRCULAR LAW

Large deviations of the top eigenvalue of random matrices and applications in statistical physics

arxiv: v1 [cond-mat.quant-gas] 12 Jul 2010

Trapped ion spin-boson quantum simulators: Non-equilibrium physics and thermalization. Diego Porras

Advanced Workshop on Anderson Localization, Nonlinearity and Turbulence: a Cross-Fertilization. 23 August - 3 September, 2010

The Kawasaki Identity and the Fluctuation Theorem

arxiv: v1 [math-ph] 19 Oct 2018

Robustness of edge states in graphene quantum dots

Equilibrium Study of Spin-Glass Models: Monte Carlo Simulation and Multi-variate Analysis. Koji Hukushima

Quantum chaos analysis of the ideal interchange spectrum in a stellarator

A coarse-grained Schrödinger cat

Discrete Simulation of Power Law Noise

Multicanonical parallel tempering

Numerical estimates of critical exponents of the Anderson transition. Keith Slevin (Osaka University) Tomi Ohtsuki (Sophia University)

Quantum Chaos and Nonunitary Dynamics

Numerical study of localization in antidot lattices

Multiple time step Monte Carlo

Chapter 29. Quantum Chaos

Analysis of the simulation

Exploring the energy landscape

arxiv: v2 [physics.data-an] 18 Nov 2007

Spectral Analysis of Directed Complex Networks. Tetsuro Murai

LOCAL MOMENTS NEAR THE METAL-INSULATOR TRANSITION

Spectra of Sparse Random Matrices

arxiv:cond-mat/ v1 [cond-mat.dis-nn] 31 Oct 2001

Multi-Ensemble Markov Models and TRAM. Fabian Paul 21-Feb-2018

Growth oscillations. LElTER TO THE EDITOR. Zheming Cheng and Robert Savit

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

Shaking a Box of Sand

Aging and energy landscapes: application to liquids and glasses

Transcription:

Spectral Statistics of Instantaneous Normal Modes in Liquids and Random Matrices Srikanth Sastry Nivedita Deo Silvio Franz SFI WORKING PAPER: 2000-09-053 SFI Working Papers contain accounts of scientific work of the author(s) and do not necessarily represent the views of the Santa Fe Institute. We accept papers intended for publication in peer-reviewed journals or proceedings volumes, but not papers that have already appeared in print. Except for papers by our external faculty, papers must be based on work done at SFI, inspired by an invited visit to or collaboration at SFI, or funded by an SFI grant. NOTICE: This working paper is included by permission of the contributing author(s) as a means to ensure timely distribution of the scholarly and technical work on a non-commercial basis. Copyright and all rights therein are maintained by the author(s). It is understood that all persons copying this information will adhere to the terms and constraints invoked by each author's copyright. These works may be reposted only with the explicit permission of the copyright holder. www.santafe.edu SANTA FE INSTITUTE

Spectral Statistics of Instantaneous Normal Modes in Liquids and Random Matrices Srikanth Sastry, 1 Nivedita Deo, 2 and Silvio Franz 3 1 Jawaharlal Nehru Center for Advanced Scientific Research Bangalore 560064, INDIA; E-mail: sastry@jncasr.ac.in 2 Santa Fe Institute 1399 Hyde Park Road, Santa Fe, NM 87501, USA; E-mail: ndeo@jncasr.ac.in and ndeo@santafe.edu 3 The Adbus Salam International Centre for Theoretical Physics Treieste, ITALY; E-mail: franz@ictp.trieste.it Abstract We study the statistical properties of eigenvalues of the Hessian matrix H (matrix of second derivatives of the potential energy) for a classical atomic liquid, and compare these properties with predictions for random matrix models (RMM). The eigenvalue spectra (the Instantaneous Normal Mode or INM spectra) are evaluated numerically for configurations generated by molecular dynamics simulations. We find that distribution of spacings between nearest-neighbor eigenvalues, s, obeys quite well the Wigner prediction s exp( s 2 ), with the agreement being better for higher densities at fixed temperature. The deviations display a correlation with the number of localized eigenstates (normal modes) in the liquid; there are fewer localized states at higher densities which we quantify by calculating the participation ratios of the normal modes. We confirm this observation by calculating the spacing distribution for parts of the INM spectra with high participation ratios, obtaining greater conformity with the Wigner form. We also calculate the spectral rigidity and find a substantial dependence on the density of the liquid. 1 Introduction The local topography of the potential energy surface, as characterized by the ensemble averaged spectrum (INM spectrum) of eigenvalues of the second- 1

derivative matrix (Hessian) of the potential energy function, have been studied in recent years as an approach to the analysis of dynamics in liquids [1, 2]. Broadly, the study of the INM spectra has been directed at analyzing shorttime dynamics as in studying solvation [3], and at elucidating information about pathways to long-time relaxation in the form of potential energy barriers etc. [1, 2]. Considerable effort has also been dedicated to developing analytical theories for calculating the INM spectra within an equilibrium description [4, 5, 6, 7, 8]. The approach in much of these attempts has been to formulate the problem of calculating the INM spectrum as an exercise in random matrix theory. If one treats the individual elements of the Hessian matrix as independent, and distributed according to the appropriate Boltzmann weight, then the Hessian may be viewed as a real, symmetric random matrix with a known distribution of matrix elements. Two properties, however, distinguish the Hessian from the standard corresponding case treated in random matrix theory: (i) The diagonal entries of the Hessian are related to the off-diagonal entries by the property H αβ ii = j i H αβ ij, where i, j label the particles and α, β the spatial coordinates x, y, z. (ii) For liquids with short-ranged interaction potentials, the Hessian matrix is sparse, with the fraction of non-zero entries, p, depending on the system size as p 1/N (the size of the matrix is N N). In view of the above considerations, it is of interest to inquire to what extent the INM spectrum displays universal features identified in random matrix theory. In this paper, we address this question by obtaining INM spectra numerically for a model atomic liqiud that has been studied in the context of slow dynamics in supercooled liquids [9]. The statistics we consider are the spacing statistics between nearest-neighbor eigenvalues and the spectral rigidity, which we explain below. 2 Eigenvalues of the Hessian The model liquid we study is a binary mixture composed of 80% of particles of type A and 20% of type B, interacting via the Lennard-Jones potential, with Lennard-Jones parameters ɛ AA =1.0, ɛ AB =1.5, ɛ BB =0.5, σ AA =1.0, σ AB =0.8, and σ BB =0.88, and a ratio of masses m B /m A = 1. Further details may be found in [9]. Molecular dynamics simulations of the liquid are performed at ten reduced densities ρ from ρ =0.65 to 1.40, at reduced 2

temperatuer T =1.0. Two hundred sample configurations in each case are chosen from the equilibrated trajectory for the INM analysis. For each of these configurations, the Hessian is calculated and diagonalized numerically to obtain the eigenvalues λ i as well as the eigenvectors e i. The eigenvectors are used to calculate the localization properties of the normal modes, via the participation ratio P i = [ N 3N α=1 ] 1 (e α i.e α i ) 2. (2.1) The participation ratio thus defined is small (order of 1/N ) for localized modes and large (order of 1) for extended modes. By constructing the histogram of eigenvalues λ of the Hessian for all configurations considered, we obtain the INM Density of States (DOS) or spectrum [10]. Figure 1 shows the INM spectrum D(λ) verses λ for different densities. Note that the DOS is very different from the Wigner semi-circle distribution, obtained in RMM. We know from the literature [12] that the correlation and spacing functions are universal in certain regions no matter what the DOS is within the RMM. Thus we use the corresponding statistics of spacings between eigenvalues, Wigner-Dyson statistics, as a standard of reference. 3 Unfolding The Spectrum The statistical analysis of the numerical data proceeds by first using an unfolding procedure. The numerical calculation yields the eigenvalues of the Hessian which is ordered and forms the sample spectrum {λ 1,λ 2,...,λ n }.In order to analyze the spacing statistics, one must transform the eigenvalues λ i in such a way that the transformed eigenvalues ζ i are uniformly distributed. That is, the spectral density ρ(ζ) = 1. This procedure is referred to as unfolding the spectrum [13, 11]. In general such a transformation for a spectral density function ρ(λ) is most easily accomplished through its cumulative distribution C(λ) λ ρ(λ )dλ, (3.2) by defining ζ(λ) =C(λ). For a discrete spectrum such as the ones we consider here, the corresponding procedure would be to extract from the sample 3

0.008 10 2 10 3 0.006 10 4 10 5 D(λ) 0.004 0.002 10 6 10 7 1e+03 1e+03 3e+03 5e+03 7e+03 ρ=0.65 ρ=1.00 ρ=1.35 0 500 500 1500 2500 3500 4500 5500 λ Figure 1: INM Density of States shown for three densities. Inset shows the DOS on a logarithmic scale. 4

1 Smooth Part of Cumulative Spectrum "Staircase" Cumulative Spectrum 0.98 1 ζ 0.8 ζ i 0.96 0.6 0.4 0.2 0 500 1500 3500 0.94 1500 λ 2000 2500 3000 i λ Figure 2: A portion of the staircase cumulative spectrum from a single configuration is shown, along with the estimate of the smooth part of the cumulative distribution, obtained from combining eigenvalues from 200 configurations, for density ρ =1.0 and temperature T =1.0. The arrows indicate the mapping of any given eigenvalue λ i to the unfolded eigenvalue ζ i. The inset shows the cumulative spectra for the full range. spectrum the smooth part of the staircase cumulative distribution [11]. In this work, the procedure we adopt to estimate the smooth part of the cumulative distribution is to evaluate the cumulative distribution for the union of all eigenvalues λ i obtained for 200 independent configurations at each density and temperature value. Such a smooth cumulative spectrum along with the staircase cumulative spectrum for a single configuration, is shown in Fig. 2, and the unfolding procedure is indicated. 5

4 Spacing Distribution The spacing distribution P (s) for the random matrix models is defined as the probability of finding the next nearest-neighbor eigenvalue of the spectrum to be at a distance s i.e., s = λ i+1 λ i where is the mean-level spacing. In the present case where we use unfolded eigenvalues, s = ζ i+1 ζ i. Then P (s) =As β e Bs2 where β = 1 for the orthogonal random matrix model, which is the Wigner surmise. This spacing distribution arises in many diverse real quantum systems which show an underlying classically chaotic behavior e.g., quantum billiards, quantum dots, nuclear spectra, disordered mesoscopic systems, etc. The system we analyze, on the other hand, is a classical liquid, with a disordered microscopic structure. The spacing distributions obtained are shown in Fig. 3 for three densities of the liquid. We find that to a first approximation, all the cases considered display the universal Wigner surmise to a large extent, with the agreement with the universal behavior improving with increasing density. We note here that in Ref. [8] a sparse random matrix is used to calculate the INM DOS analytically for a one-dimensional system which is very different from the usual Wigner semicircle, and displays qualitative features very similar to the DOS we calculate numerically here. It would be very interesting to see whether the correlators of the sparse random matrix proposed in ref. [8] captures the above behaviour for different densities of the liquid. 5 Spectral Rigidity We next study the spectral rigidity, which measures the fluctuations of the number of eigenvalues in a window of given size as a function of the size of the window (or equivalently, the average number of eigenvalues expected in the window). The number fluctuations are plotted as a function of the mean number of eigenvalues in Fig. 4 for the same three densities as in the preceding figures. For the case of the Poisson spectrum and the harmonic oscillator the number fluctuations are δn 2 N and langleδn 2 = constant. While for the Gaussian Random matrix ensembles δn 2 ln N for N 1. The number fluctuations are of the form δn 2 N γ where γ(ρ) depends on the different densities. This situation is reminescent of that found for the mobility edge in the quantum hall effect and in the Anderson metal-insulator transition. A detailed analysis of this awaits future work. 6

600 1000 100 400 10 P(s) 1 200 ρ=0.65 ρ=1.00 ρ=1.35 Fit to ρ=0.65 Fit to ρ=1.00 Fit to ρ=1.35 1e 04 1e 03 1e 02 0 0.0e+00 2.0e 03 4.0e 03 s Figure 3: The level spacing distribution P (s) for densities ρ =0.65, 1.0, 1.35. 7

60 ρ=0.65 ρ=1.00 ρ=1.35 <N 2 > <N> 2 40 20 0 0 50 100 150 <N> Figure 4: Spectral rigidity: Lines are fits to the form N 2 N 2 N γ, and the values of γ are 1.036, 0.85, 0.765 respectively for ρ =0.65, 1.00, 1.35. 8

0.8 0.6 Participation Ratio 0.4 0.2 ρ=0.65 ρ=1.00 ρ=1.35 0 0 0.2 0.4 0.6 0.8 1 Unfolded Eigenvalues 6 Participation Ratio Figure 5: Participation ratio. In order to get some insight into the reason for the systematically better agreement with increasing density of the spacing distribution, we consider the localization properties of the normal modes in the liquid. In the standard random matrix case for orthogonal matrices, the eigenvectors are all extended, while from previous numerical and analytical studies we know that a fraction of the INM eigenmodes are localized. As described earlier, we use the participation ratio to quantify the localization of modes, averaging over modes corresponding to eigenvalues in each histogram bin, for the unfolded eigenvalues. In Fig. 5, the participation ratio is plotted as a function of the unfolded eigenvalues for T =1.00, for values of the density ρ =0.65, 1.00 and 1.35. One notes that for the highest density ρ = 1.35 the participation ratios are highest overall, while for the lower densities the participation ratios are quite small for a substantial fraction exit Next, we calculate the spacing distribution for ρ =1.35 for unfolded 9

1000 100 P(s) 10 1 ρ=1.35 ρ=1.35, Participation Ratio > 0.75 Fit: A s exp( s 2 /B) 1.00e 04 1.00e 03 1.00e 02 s Figure 6: The level spacing distribution shown for ρ =1.35 for the full eigenvalue spectrum (open rhombs), for the range of eigenvalues with high participation ratio ( 0.75) (filled rhombs), along with the RMM prediction. eigenvalues between 0.1 and 0.6 for which the participation ratio is high (0.75) and relatively unchanging (Fig. 5). The resulting spacing distribution is shown in Fig. 6, along with the spacing distribution for the entire eigenvalue spectrum and the expectation based on the RMM result. The data shown clearly demonstrate that the spacing distribution is practically identical to the standard RMM result, confirming the speculation that the increasing fraction of localized states at lower densities are responsible for the deviations at these densities from the standard RMM result. Indeed, this observation has been used, in a different context to locate the mobility edge in disordered systems [19]. 10

7 Conclusions We have presented the spacing statistics and spectral rigidity for numerically calculated INM spectra. The spacing statistics is seen to conform better with increasing density (at fixed temerature), with the predictions for random matrix models. We demonstrate that the source of deviations from RMM predictions is related to the presence of localized instantaneous normal modes in the liquid, whose number is greater for lower density. As the two features distinguishing the INM spectra from the standard random matrix case are the nature of the diagonal elements and the sparseness of the Hessian matrix, further understanding of the non-universality of the INM spectra are to be sought in the manner in which these aspects affect the INM spectral statistics. Acknowledgments We would like to thank S. Jain, V. E. Kravtsov, A. Cavagna, I. Giardina, P. J. Garrahan and P. Carpena for very useful discussions during the course of this work. References [1] T. Keyes, J. Chem. Phys. A 101, 2921, (1997) and references therein. [2] F. Sciortino and P. Tartaglia, Phys. Rev. Lett. 78, 2385 (1997). [3] R. M. Stratt and M. Cho, J. Chem. Phys. 100, 6700, (1994). [4] B.-C. Xu and R. M. Stratt, J. Chem. Phys. 92, 1923 (1990). [5] T. M. Wu and R. F. Loring, J. Chem. Phys. 97, 8568, (1992). [6] Y. Wan and R. M. Stratt, J. Chem. Phys. 100, 5123, (1994). [7] G. Biroli and R. Monasson, cond-mat/9902032. [8] A. Cavagna, I. Giardina and G. Parisi, Phys. Rev. Lett. 83, 108 (1999). [9] S. Sastry, P. G. Debenedetti and F. H. Stillinger, Nature 393, 554 (1998). 11

[10] Note that we have not followed the standard practice in the INM literature of plotting the DOS against ω sign(λ) λ. [11] T. Guhr, A. Muller-Groeling and H. A. Weidenmuller, Physics Reports 299(4-6) June 1998 and references therein. [12] E. Brézin and A. Zee, Nucl. Phys. B 402, 613 (1993), Phys. Rev. E 49, 2588 (1994). [13] M. L. Mehta (1991) Random Matrices, Academic Press, San Diego. [14] P. Cizeau and J. P. Bouchaud, Phys. Rev. E 50, 1810, (1994). [15] A. J. Bray and G. J. Rodgers, Phys. Rev. B 38, 11461, (1988), Phys. Rev. B 37, 3557, (1988) [16] G. J. Rodgers and C. De Dominicis, J. Phys. A: Math. Gen. 23 (1990) 1567-1573 [17] A. D. Mirlin and Yan V. Fyodorov, J. Phys. A: Math. Gen. 24 (1991) 2273-2286, Phys. Rev. Lett. 67, (1991) 2049. [18] M. Mezard, G. Parisi and A. Zee, Spectra of Euclidean Random Matrices, cond-mat/9906135. [19] P. Carpena and P. Bernaola-Galvań Phys. Rev. B 60, 201 (1999). 12