cond-mat0101078/sdf.tex
1: \documentstyle[preprint,aps,eqsecnum]{revtex}
2: \input epsf
3: 
4: \newcommand{\eeq}{\end{equation}}
5: \newcommand{\vs}[1]{\rule[- #1 mm]{0mm}{#1 mm}}
6: \newcommand{\beq}{\begin{equation}}
7: \newcommand{\beql}{\begin{eqnarray}}
8: \newcommand{\eeql}{\end{eqnarray}}
9: \newcommand{\lp}{\left(}
10: \newcommand{\rp}{\right)}
11: \newcommand{\scr}{\scriptstyle}
12: \newcommand{\om}{\omega}
13: \newcommand{\al}{\alpha}
14: \newcommand{\lam}{\lambda}
15: \newcommand{\Lam}{\Lambda}
16: \newcommand{\eps}{\epsilon}
17: \newcommand{\dV}{\frac{d}{dV}}
18: \newcommand{\dxi}{\frac{dx_i}{dV}(p)}
19: \newcommand{\parV}{\frac{\partial}{\partial V}}
20: \newcommand{\sect}[1]{\setcounter{equation}{0}\section{#1}}
21: \newcommand{\NP}[3]{{\it Nucl. Phys. }{\bf B#1} (#2) #3}
22: \newcommand{\PL}[3]{{\it Phys. Lett. }{\bf B#1} (#2) #3}
23: \newcommand{\PR}[3]{{\it Phys. Rev. }{\bf #1} (#2) #3}
24: \newcommand{\PRL}[3]{{\it Phys. Rev. Lett. }{\bf #1} (#2) #3}
25: \newcommand{\IMP}[3]{{\it Int. J. Mod. Phys }{\bf #1} (#2) #3}
26: \newcommand{\MPL}[3]{{\it Mod. Phys. Lett. }{\bf #1} (#2) #3}
27: \newcommand{\JP}[3]{{\it J. Phys. }{\bf A#1} (#2) #3} 
28: \renewcommand{\theequation}{\thesection.\arabic{equation}}
29: 
30: \begin{document}
31: 
32: \draft 
33: 
34: 
35: \title{Spectral Statistics of Instantaneous Normal Modes in Liquids and Random Matrices}
36: 
37: \author{Srikanth Sastry$^1$, Nivedita Deo$^{1,3}$ and Silvio Franz$^2$}
38: 
39: \address{$^1$ Jawaharlal Nehru Center for Advanced Scientific Research, Bangalore 560064, India}
40: \address{$^2$ The Adbus Salam International Centre for Theoretical Physics,
41: Treieste, Italy}
42: \address{$^3$ Santa Fe Institute,
43: 1399 Hyde Park Road, Santa Fe, NM 87501, USA}
44: 
45: \maketitle
46: 
47: \begin{abstract}
48: We study the statistical properties of eigenvalues of the Hessian
49: matrix ${\cal H}$ (matrix of second derivatives of the potential
50: energy) for a classical atomic liquid, and compare these properties
51: with predictions for random matrix models (RMM). The eigenvalue
52: spectra (the Instantaneous Normal Mode or INM spectra) are evaluated
53: numerically for configurations generated by molecular dynamics
54: simulations. We find that distribution of spacings between nearest
55: neighbor eigenvalues, $s$, obeys quite well the Wigner prediction $~
56: s~exp(-s^2)$, with the agreement being better for higher densities at
57: fixed temperature. The deviations display a correlation with the
58: number of localized eigenstates (normal modes) in the liquid; there
59: are fewer localized states at higher densities which we quantify by
60: calculating the participation ratios of the normal modes. We confirm
61: this observation by calculating the spacing distribution for parts of
62: the INM spectra with high participation ratios, obtaining greater
63: conformity with the Wigner form. We also calculate the spectral
64: rigidity and find a substantial dependence on the density of the
65: liquid.
66: \end{abstract}
67: 
68: \pacs{PACS numbers: 75.10.Nr, 61.20.-p, 61.43.-j, 64.70.Pf, 05.40.-a, 05.45+b}
69: 
70: \section{Introduction:}
71: The local topography of the potential energy surface, as characterized
72: by the ensemble averaged spectrum (INM spectrum) of eigenvalues of the
73: second-derivative matrix (Hessian) of the potential energy function,
74: have been studied in recent years as an approach to the analysis of
75: dynamics in liquids\cite{keyes,Sciortino}. Broadly, the study
76: of the INM spectra has been directed at analysing short time dynamics 
77: as in studying solvation \cite{strattcho}, and at elucidating information
78: about pathways to long time relaxation in the form of potential energy 
79: barriers {\it etc}\cite{keyes,Sciortino}. 
80: 
81: Considerable effort has also been dedicated to developing analytical
82: theories for calculating the INM spectra within an equilibrium
83: description\cite{xustratt,wu,stratt,mona,cava}. The approach in much
84: of these attempts has been to formulate the problem of calculating the
85: INM spectrum as an exercise in random matrix theory. If one treats the
86: individual elements of the Hessian matrix as independent, and
87: distributed according to the appropriate Boltzmann weight, then the
88: Hessian may be viewed as a real, symmetric random matrix with a known
89: distribution of matrix elements. Two properties, however, distinguish
90: the Hessian from the standard corresponding case treated in random
91: matrix theory: (i) The diagonal entries of the Hessian are related to
92: the off-diagonal entries by the property ${\cal H}^{\alpha \beta}_{ii}
93: = - \sum_{j \neq i} {\cal H}^{\alpha \beta}_{ij}$, where $i,j$ label
94: the particles and $\alpha, \beta$ the spatial coordinates $x, y,
95: z$. (ii) For liquids with short-ranged interaction potentials, the
96: Hessian matrix is sparse, with the fraction of non-zero entries, $p$,
97: depending on the system size as $p \sim 1/N$.
98: 
99: In view of the above considerations, it is of interest to inquire to
100: what extent the INM spectrum displays universal features idenitied in
101: random matrix theory. In this paper, we address this question by
102: obtaining INM spectra numerically for a model atomic liqiud that has
103: been studied in the context of slow dynamics in supercooled
104: liquids\cite{sastryNature}. The statistics we consider are the
105: spacing statistics between nearest neighbor eigenvalues and the 
106: spectral rigidity, which we explain below. 
107: 
108: \noindent
109: \section{Eigenvalues of the Hessian}
110: The model liquid we study is a binary mixture\cite{kob} composed of
111: $80 \%$ of particles of type $A$ and $20 \%$ of type $B$, interacting
112: via the Lennard-Jones potential, with Lennard-Jones parameters
113: $\epsilon_{AB}/\epsilon_{AA} = 1.5$, $\epsilon_{BB}/\epsilon_{AA} =
114: 0.5$, $\sigma_{AB}/\sigma_{AA} = 0.8$, and $\sigma_{BB}/\sigma_{AA} =
115: 0.88$, and a ratio of masses $m_B/m_A = 1$. Lennard-Jones reduced
116: units are used to report all the quantities, in terms of the $A$
117: particle paramters $\epsilon_{AA}$, $\sigma_{AA}$ and $m_{A}$:
118: temperatures as $T^{*} = k_B T/\epsilon_{AA}$, densities as $\rho^{*}
119: = \rho/\sigma_{AA}^3$ and Hessian eigenvalues $\lambda^{*} = \lambda
120: {m_A \sigma_{AA}^2 \over \epsilon_{AA}}$. Further details may be found
121: in \cite{sastryNature}.  Molecular dynamics simulations of the liquid
122: are performed at ten reduced densities $\rho^{*}$ from $\rho^{*} =
123: 0.65 $ to $1.40$, at reduced temperature $T^{*} = 1.0$. Two hundred
124: sample configurations in each case are chosen from the equilibrated
125: trajectory for the INM analysis. For each of these configurations, the
126: Hessian is calculated and diagonalized numerically to obtain the
127: eigenvalues $\lambda_i$ as well as the eigenvectors ${\bf e}_i$. The
128: eigenvectors are used to calculate the localization properties of the
129: normal modes, {\it via} the participation ratio
130: 
131: \begin{equation}
132: P_i = \left[N \sum_{\alpha = 1}^{3N} ({\bf e}_i^\alpha.{\bf e}_i^\alpha)^2 \right]^{-1}. 
133: \end{equation}
134: 
135: The participation ratio thus defined is small (order of $1/N$) for localized modes and large (order of $1$) for extended modes. 
136: 
137: By constructing the histogram of eigenvalues $\lambda$ of the Hessian
138: for all configurations considered, we obtain the INM Density of States
139: (DOS) or spectrum\cite{fn1}. Figure 1 shows the INM spectrum
140: $D(\lambda)$ verses $\lambda$ for different densities. Note that the
141: DOS is very different from the Wigner semi-circle distribution,
142: obtained in RMM. We know from the literature\cite{bz} that the
143: correlation and spacing functions are universal in certain regions no
144: matter what the DOS is. Thus we use the corresponding statistics of
145: spacings between eigenvalues, Wigner-Dyson statistics, as the standard
146: of reference.
147: 
148: \noindent
149: \section{Unfolding The Spectrum}
150: The statistical analysis of the numerical data proceeds by first using
151: an {\it unfolding procedure}. The numerical calculation yields the
152: eigenvalues of the Hessian which is ordered and forms the sample
153: spectrum $\{\lambda_1,\lambda_2,...,\lambda_n\}$. In order to analyse
154: the spacing statistics, one must transform the eigenvalues $\lambda_i$
155: in such a way that the transformed eigenvalues $\zeta_i$ are uniformly
156: distributed.  That is, the spectral density $D(\zeta) = 1$. This
157: procedure is referred to as ``unfolding'' the
158: spectrum\cite{mehta,guhr}. In general such a transformation for a
159: spectral density function $D(\lambda)$ is most easily accomplished
160: through its cumulative distribution
161: \begin{equation}
162: C(\lambda) \equiv \int_{-\infty}^{\lambda} D(\lambda) d\lambda,
163: \end{equation}
164: by defining $\zeta(\lambda) = C(\lambda)$. For a discrete spectrum
165: such as the ones we consider here, the corresponding procedure would
166: be to extract from the sample spectrum the ``smooth'' part of the
167: ``staircase'' cumulative distribution\cite{guhr}. In this work, the
168: procedure we adopt to estimate the smooth part of the cumulative
169: distribution is to evaluate the cumulative distribution for the union
170: of all eigenvalues $\lambda_i$ obtained for $200$ independent
171: configurations at each density and temperature value. Such a smooth
172: cumulative spectrum along with the ``staircase'' cumulative spectrum
173: for a single configuration, is shown in Fig. 2, and the unfolding
174: procedure is indicated.
175: 
176: \noindent
177: \section{Spacing Distribution}
178: The spacing distribution $P(s)$ for the random matrix models is
179: defined as the probability of finding the next nearest neighbor
180: eigenvalue of the spectrum to be at a distance s i.e.  $
181: s_i={{\lambda_{i+1}-\lambda_i}\over {\Delta}} $ where $ \Delta $ is
182: the mean level spacing. In the present case where we use unfolded
183: eigenvalues, $s_i=\zeta_{i+1}-\zeta _{i}$.  Then $P(s)= A s^{\beta}
184: e^{-B s^2}$ where $\beta=1$ for the orthogonal random matrix model,
185: which is the ``Wigner surmise''\cite{mehta}.  This spacing
186: distribution arises in various quantum systems which show an
187: underlying classically chaotic behavior e.g. quantum billiards,
188: quantum dots, nuclear spectra, disordered mesoscopic systems. The
189: system we analyze, on the other hand, is a classical liquid, with a
190: disordered microscopic structure. The spacing distributions obtained
191: are shown in Fig. 3 for three densities of the liquid. We find that to
192: a first approximation, the cases considered display the universal
193: behavior according to the Wigner surmise, with the agreement being
194: better for the liquid at higher densities. We note here that in
195: Ref. \cite{cava} a sparse random matrix is used to calculate the INM
196: DOS analytically for a one-dimensional system which is very different
197: from the usual Wigner semi-circle, and displays qualitative features
198: very similar to the DOS we calculate numerically here.  It would be
199: very interesting to see whether the correlators of the sparse random
200: matrix proposed in ref. \cite{cava} captures the above behaviour for
201: different densities for the liquid.
202: 
203: \noindent
204: \section{Spectral Rigidity}
205: We next study the spectral rigidity, which measures the fluctuations
206: of the number of eigenvalues in a window of given size as a function
207: of the size of the window (or equivalently, the average number of
208: eigenvalues expected in the window). The number fluctuations are
209: plotted as a function of the mean number of eigenvalues in Fig. 4 for
210: the same three densities as in the preceding figures. For the case of
211: the Poisson spectrum and the harmonic oscillator\cite{mehta} the
212: number fluctuations are $ <\delta N^2> \sim <N> $ and $ <\delta N^2> =
213: constant $, while for the Gaussian random matrix ensembles $ <\delta
214: N^2> \sim ln <N> $ for $ N \gg 1 $.
215: 
216: 
217: The number fluctuations are of the form $ <\delta N^2> \sim
218: N^{\gamma} $ where $\gamma(\rho)$ depends on the liquid's density.
219: This situation is reminescent of that found for the mobility edge in
220: the quantum hall effect and in the Anderson metal-insulator
221: transition.  A detailed analysis of this awaits future work.
222: 
223: \noindent
224: \section{Participation Ratio}
225: In order to get some insight into the reason for the systematically
226: better agreement with increasing density of the spacing distribution,
227: we consider the localization properties of the normal modes in the
228: liquid.  In the standard random matrix case for orthogonal matrices,
229: the eigenvectors are all extended, while from previous numerical and
230: analytical studies we know that a fraction of the INM eigenmodes are
231: localized. As described earlier, we use the participation ratio to
232: quantify the localization of modes, averaging over modes corresponding
233: to eigenvalues in each histogram bin, for the unfolded eigenvalues.
234: In Fig. 5, the participation ratio is plotted as a function of the
235: unfolded eigenvalues for $T^{*} =1.00$, for values of the density
236: $\rho^{*} =0.65, 1.00$ and $1.35$. One notes that for the highest density
237: $\rho=1.35$ the participation ratios are highest overall, while for the 
238: lower densities the participation ratios are quite small for a substantial 
239: fraction of the eigenmodes, indicating a large number of localized modes. 
240: 
241: 
242: 
243: Next, we calculate the spacing distribution for $\rho^{*} = 1.35$ for
244: unfolded eigenvalues between $0.1$ and $0.6$ for which the
245: participation ratio is high ($> 0.75$) and relatively unchanging
246: (Fig. \ref{fig:prat}). The resulting spacing distribution is shown in
247: Fig. 6, along with the spacing distribution for the entire eigenvalue 
248: spectrum and the expectation based on the RMM result. The data shown 
249: clearly demonstrate that the spacing distribution is practically 
250: identical to the standard RMM result, confirming the speculation that 
251: the increasing fraction of localized states at lower densities are 
252: responsible for the deviations at these densities from the standard 
253: RMM result. Indeed, this observation has been used, in a different context
254: to locate the mobility edge in disordered systems\cite{carpena}
255: 
256: \noindent
257: \section{Conclusions}
258: We have presented the spacing statistics and spectral rigidity for
259: numerically calculated INM spectra. The spacing statistics is seen to
260: conform better with increasing density (at fixed temerature), with the
261: predictions for random matrix models. We demonstrate that the source
262: of deviations from RMM predictions is related to the presence of
263: localized instantaneous normal modes in the liquid, whose number is
264: greater for lower density. As the two features distinguishing the INM
265: spectra from the standard random matrix case are the nature of the
266: diagonal elements and the sparseness of the Hessian matrix, further 
267: understanding of the non-universality of the INM spectra are to be
268: sought in the manner in which these aspects affect the INM spectral 
269: statistics. 
270: 
271: 
272: \noindent
273: 
274: \begin{flushleft}
275: 
276: \underline{Acknowledgements}:\\ 
277: 
278: We would like to thank S. Jain, V. E. Kravtsov, A. Cavagna,
279: I. Giardina, P. J. Garrahan and P. Carpena for very useful discussions
280: during the course of this work. SF thanks the Jawaharlal Nehru Center
281: for Advanced Scientific Research for hospitality.
282: \end{flushleft}
283: 
284: \begin{figure}
285: %\hbox to\hsize{\epsfxsize=0.6\hsize\hfil\epsfbox{sdf-fig1.eps}\hfil}
286: \caption{INM Density of States shown for three densities. Inset shows
287: the DOS on a logarithmic scale. Eigenvalues $\lambda$ are expressed in
288: units of ${\epsilon_{AA} \over m_A \sigma_{AA}^2}$, and $D(\lambda)$ in
289: units of ${m_A \sigma_{AA}^2 \over \epsilon_{AA}}$.}
290: \label{fig:inmdos}
291: \end{figure}
292: 
293: \begin{figure}
294: %\hbox to\hsize{\epsfxsize=0.6\hsize\hfil\epsfbox{sdf-fig2.eps}\hfil}
295: \caption{A portion of the staircase cumulative spectrum from a single
296: configuration is shown, along with the estimate of the smooth part of
297: the cumulative distribution, obtained from combining eigenvalues from
298: 200 configurations, for density $\rho^{*} = 1.0$ and temperature
299: $T^{*} = 1.0$. The arrows indicate the mapping of any given eigenvalue
300: $\lambda_i$ to the unfolded eigenvalue $\zeta_i$. The inset shows the
301: cumulative spectra for the full range. Eigenvalues $\lambda$ are
302: expressed in units of ${\epsilon_{AA} \over m_A \sigma_{AA}^2}$, and 
303: cumulative probabilities $\zeta$ are dimensionless.}
304: \label{fig:unfold}
305: \end{figure}
306: 
307: \begin{figure}
308: %\hbox to\hsize{\epsfxsize=0.6\hsize\hfil\epsfbox{sdf-fig3.eps}\hfil}
309: \caption{The level spacing distribution $P(s)$ for densities $\rho =
310: 0.65, 1.0, 1.35$. Level spacings $s$ are expressed in units of
311: ${\epsilon_{AA} \over m_A \sigma_{AA}^2}$, and $P(s)$ in units of
312: ${m_A \sigma_{AA}^2 \over \epsilon_{AA}}$.}
313: \label{fig:pofs}
314: \end{figure}
315: 
316: 
317: \begin{figure}
318: %\hbox to\hsize{\epsfxsize=0.5\hsize\hfil\epsfbox{sdf-fig4.eps}\hfil}
319: \caption{Spectral rigidity: 
320: Lines are fits to the form $<N^2> - <N>^2 \sim <N>^\gamma$, and the
321: values of $\gamma$ are $1.036, 0.85, 0.765$ respectively for $\rho^{*} = 0.65, 1.00, 1.35$.  }
322: \label{fig:rigid}
323: \end{figure}
324: 
325: 
326: \begin{figure}[t]
327: %\hbox to\hsize{\epsfxsize=0.5\hsize\hfil\epsfbox{sdf-fig5.eps}\hfil}
328: \caption{Participation ratio as a function of unfolded eigenvalues
329: $\zeta$ (see caption of Fig. 2).}
330: \label{fig:prat}
331: \end{figure}
332: 
333: 
334: \begin{figure}[t]
335: %\hbox to\hsize{\epsfxsize=0.5\hsize\hfil\epsfbox{sdf-fig6.eps}\hfil}
336: \caption{The level spacing distribution shown for $\rho^{*} = 1.35$ for
337: the full eigenvalue spectrum (open rhombs), for the range of
338: eigenvalues with high participation ratio ($\> 0.75$) (filled rhombs),
339: along with the RMM prediction. Level spacings $s$ are expressed in units of
340: ${\epsilon_{AA} \over m_A \sigma_{AA}^2}$, and $P(s)$ in units of
341: ${m_A \sigma_{AA}^2 \over \epsilon_{AA}}$.}
342: \label{fig:pofs2}
343: \end{figure}
344: 
345: 
346: \begin{thebibliography}{99}
347: 
348: \bibitem[] {} Email: {$^1$sastry@jncasr.ac.in, $^2$ndeo@jncasr.ac.in,
349: $^3$franz@ictp.trieste.it}.
350: 
351: \bibitem{keyes} T. Keyes, {\it J. Chem. Phys.} {\bf A 101},
352: 2921, (1997) and references therein.
353: 
354: \bibitem{Sciortino}
355: F. Sciortino and P. Tartaglia, Phys. Rev. Lett. {\bf 78}, 2385 (1997).
356: 
357: \bibitem{strattcho} R. M. Stratt and M. Cho, {\it J. Chem. Phys.} {\bf 100},
358: 6700, (1994).
359: 
360: \bibitem{xustratt} B.-C. Xu and R. M. Stratt, {\it J. Chem. Phys.} {\bf 92}, 1923 (1990).
361: 
362: \bibitem{wu} T. M. Wu and R. F. Loring, {\it J. Chem. Phys.} {\bf 97},
363: 8568, (1992).
364: 
365: \bibitem{stratt} Y. Wan and R. M. Stratt, {\it J. Chem. Phys.} {\bf 100},
366: 5123, (1994).
367: 
368: \bibitem{mona} G. Biroli and R. Monasson, {\it J. Phys. A: Math. Gen.}
369: {\bf 32}, L255 (1999).
370: 
371: \bibitem{cava} A. Cavagna, I. Giardina and G. Parisi,
372: {\it Phys. Rev. Lett.} {\bf 83}, 108 (1999).
373: 
374: \bibitem{kob} W. Kob and H. C. Andersen, {\it Phys. Rev. E} {\bf 51}, 4626 (1995); K. Vollmayr, W. Kob and K. Binder {\it J. Chem. Phys.} {\bf 105}, 4714 (1996).
375: 
376: \bibitem{sastryNature} S. Sastry, P. G. Debenedetti and
377: F. H. Stillinger, {\it Nature} {\bf 393}, 554 (1998).
378: 
379: \bibitem{fn1} Note that we have not followed the standard practing in
380: the INM literature of plotting the DOS against $\omega \equiv sign(\lambda)\times \sqrt{|\lambda|}$.
381: 
382: \bibitem{guhr} T. Guhr, A. Muller-Groeling and H. A. Weidenmuller,
383: Physics Reports, {\bf 299}, Number 4-6, (1998) and references therein.
384: 
385: \bibitem{bz} E. Br\'ezin and A. Zee, Nucl. Phys. B {\bf 402}, 613 (1993),
386: Phys. Rev. E {\bf 49}, 2588 (1994).
387: 
388: \bibitem{mehta} M. L. Mehta (1991) Random Matrices, Academic Press, San Diego.
389: 
390: \bibitem{bouch} P. Cizeau and J. P. Bouchaud, {\it Phys. Rev. E} {\bf 50},
391: 1810, (1994).
392: 
393: \bibitem{bray} A. J. Bray and G. J. Rodgers, {\it Phys. Rev. B} {\bf
394: 38}, 11461, (1988); G. J. Rodgers and A. J. Bray, {\it Phys. Rev. B}
395: {\bf 37}, 3557 (1988).
396: 
397: \bibitem{dedom} G. J. Rodgers and C. De Dominicis, {\it J. Phys. A}
398: Math. Gen. {\bf 23}, 1567 (1990).
399: 
400: \bibitem{mirlin} A. D. Mirlin and Y. V. Fyodorov, J. Phys. A: Math.
401: Gen. {\bf 24},2273 (1991); Y. V. Fyodorov and A. D. Mirlin, {\it
402: Phys. Rev. Lett.} {\bf 67}, 2049 (1991).
403: 
404: \bibitem{mezard} M. Mezard, G. Parisi and A. Zee,
405: http://arXiv.org/abs/cond-mat/9906135.
406: 
407: \bibitem{carpena} P. Carpena and P. Bernaola-Galva\'n {\it
408: Phys. Rev. B} {\bf 60}, 201 (1999).
409: 
410: \end{thebibliography}
411: 
412: \end{document}
413:                                                                     
414: 
415: 
416: 
417: 
418: