Quantum Disordering versus Melting in Lennard-Jones Clusters.

Similar documents
J. Phys. Chem. A XXXX, xxx, 000 A. Effects of Quantum Delocalization on Structural Changes in Lennard-Jones Clusters

Structural Transitions and Melting in LJ Lennard-Jones Clusters from Adaptive Exchange Monte Carlo Simulations

Pressure Dependent Study of the Solid-Solid Phase Change in 38-Atom Lennard-Jones Cluster

Global Minimum Structures of Morse Clusters as a Function of the Range of the Potential: 81 e N e 160

Analysis of the Lennard-Jones-38 stochastic network

A Heat Capacity Estimator for Fourier Path Integral Simulations

Global minima for rare gas clusters containing one alkali metal ion

The effect of the range of the potential on the structures of clusters

arxiv:cond-mat/ v2 8 Jan 2004

Close-packing transitions in clusters of Lennard-Jones spheres

Computational study of the structures and thermodynamic properties of ammonium chloride clusters using a parallel jump-walking approach

Competition between face-centered cubic and icosahedral cluster structures

MatSci 331 Homework 4 Molecular Dynamics and Monte Carlo: Stress, heat capacity, quantum nuclear effects, and simulated annealing

Geometry Optimization and Conformational Analysis of (C 60 ) N Clusters Using a Dynamic Lattice-Searching Method

Superfluidity in Hydrogen-Deuterium Mixed Clusters


Exploring the energy landscape

Analysis of the simulation

An Extended van der Waals Equation of State Based on Molecular Dynamics Simulation

Department of Chemical Engineering University of California, Santa Barbara Spring Exercise 2. Due: Thursday, 4/19/09

INTERMOLECULAR FORCES

Thermodynamics of Three-phase Equilibrium in Lennard Jones System with a Simplified Equation of State

Melting transitions in aluminum clusters: The role of partially melted intermediates

Phase transition phenomena of statistical mechanical models of the integer factorization problem (submitted to JPSJ, now in review process)

Ideal Gas Behavior. NC State University

Modeling the structure of clusters of C 60 molecules

q lm1 q lm2 q lm3 (1) m 1,m 2,m 3,m 1 +m 2 +m 3 =0 m 1 m 2 m 3 l l l

J.C. Hamilton Sandia National Laboratories, Livermore, California 94551

Imperfect Gases. NC State University

ICCP Project 2 - Advanced Monte Carlo Methods Choose one of the three options below

Quantum Momentum Distributions

THE REAL MEANING OF THE HUBBLE DIAGRAM

Dipartimento di Sistemi e Informatica

SIMULATIONAL ANALYSIS OF GLASSES PREPARED VIA DIFFERENT INTERATOMIC POTENTIALS

Geometry explains the large difference in the elastic properties of fcc and hcp crystals of hard spheres Sushko, N.; van der Schoot, P.P.A.M.

arxiv: v1 [physics.bio-ph] 26 Nov 2007

Metropolis Monte Carlo simulation of the Ising Model

Structure of the First and Second Neighbor Shells of Water: Quantitative Relation with Translational and Orientational Order.

arxiv:cond-mat/ v1 27 Mar 1998

Thermostatic Controls for Noisy Gradient Systems and Applications to Machine Learning

Atoms, Molecules and Solids. From Last Time Superposition of quantum states Philosophy of quantum mechanics Interpretation of the wave function:

The Wave Function. Chapter The Harmonic Wave Function

Molecular Dynamics Simulations of Glass Formation and Crystallization in Binary Liquid Metals

The Potential Energy Surface (PES) And the Basic Force Field Chem 4021/8021 Video II.iii

G. Gantefdr and W. Eberhardt Institut fiir Festkiirperforschung, Forschungszentrum Jiilich, 5170 Jiilich, Germany

Adsorption and Phase Properties of Inert Gases on Suspended Carbon Nanotube Resonators

Lecture 6 - Bonding in Crystals

Nucleation rate (m -3 s -1 ) Radius of water nano droplet (Å) 1e+00 1e-64 1e-128 1e-192 1e-256

Structural optimization of Lennard-Jones clusters by a genetic algorithm. Abstract

arxiv:quant-ph/ v1 1 Jul 2003

What did you learn in the last lecture?

1+e θvib/t +e 2θvib/T +

UB association bias algorithm applied to the simulation of hydrogen fluoride

Citation PHYSICAL REVIEW E (2002), 65(6) RightCopyright 2002 American Physical So

Two-Dimensional Spin-Polarized Hydrogen at Zero

Long Range Frustration in Finite Connectivity Spin Glasses: Application to the random K-satisfiability problem

The chemical bond The topological approach

Structural Bioinformatics (C3210) Molecular Mechanics

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 5 Jun 2003

Basics of Statistical Mechanics

CHEM-UA 127: Advanced General Chemistry I

Physics 127b: Statistical Mechanics. Landau Theory of Second Order Phase Transitions. Order Parameter

arxiv: v1 [hep-lat] 7 Sep 2007

Example questions for Molecular modelling (Level 4) Dr. Adrian Mulholland

Wavelength Effects on Strong-Field Single Electron Ionization

Physics 119A Final Examination

Atomic Structure. Niels Bohr Nature, March 24, 1921

HW posted on web page HW10: Chap 14 Concept 8,20,24,26 Prob. 4,8. From Last Time

Energy-Decreasing Dynamics in Mean-Field Spin Models

UNIT I SOLID STATE PHYSICS

Lennard-Jones as a model for argon and test of extended renormalization group calculations

G. L. KELLOGG. Sandia National Laboratories Albuquerque, NM USA. 1. Abstract

The effect of the range of the potential on the structure and stability of simple liquids: from clusters to bulk, from sodium to C 60

Gibbs ensemble simulation of phase equilibrium in the hard core two-yukawa fluid model for the Lennard-Jones fluid

Entropy online activity: accompanying handout I. Introduction

The (magnetic) Helmholtz free energy has proper variables T and B. In differential form. and the entropy and magnetisation are thus given by

Minimization of the potential energy surface of Lennard Jones clusters by quantum optimization

3.091 Introduction to Solid State Chemistry. Lecture Notes No. 9a BONDING AND SOLUTIONS

3.320 Lecture 23 (5/3/05)

EFFECTS OF STOICHIOMETRY ON POINT DEFECTS AND IMPURITIES IN GALLIUM NITRIDE

A Study of the Thermal Properties of a One. Dimensional Lennard-Jones System

CE 530 Molecular Simulation

The Hubbard model for the hydrogen molecule

arxiv:cond-mat/ v1 6 Jan 2000

So far in talking about thermodynamics, we ve mostly limited ourselves to

Atomic and molecular interaction forces in biology

Ions. LESSON 19 Noble Gas Envy. Think About It. How is chemical stability related to the arrangements of electrons in atoms?

Physics 211B : Problem Set #0

New Physical Principle for Monte-Carlo simulations

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester Christopher J. Cramer. Lecture 30, April 10, 2006

A Dynamic Lattice Searching Method for Fast Optimization of Lennard Jones Clusters

Structural transition in (C 60 ) n clusters

Density Functional Theory of the Interface between Solid and Superfluid Helium 4

Formation Mechanism and Binding Energy for Icosahedral Central Structure of He + 13 Cluster

A mathematical model for a copolymer in an emulsion

Critical Properties of Isobaric Processes of Lennard-Jones Gases

PDEs in Spherical and Circular Coordinates

Chapter 11 Intermolecular Forces, Liquids, and Solids. Intermolecular Forces

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

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

Transcription:

Quantum Disordering versus elting in ennard-jones Clusters. Jason Deckman and Vladimir A. andelshtam Chemistry Department, University of California at Irvine, Irvine, CA 92697, USA (Dated: December 7, 28) The ground states of ennard-jones clusters for sizes up to n = 147 are estimated as a function of the de Boer quantum delocalization length Λ, and the n Λ phase diagram is constructed. The increase of Λ favors more disordered and diffuse structures over more symmetric and compact ones, eventually making the liquid-like motif most energetically favorable. The analogy between the quantum- and thermally-induced structural transitions is explored. PACS numbers: The ennard-jones (J) clusters exhibit rich thermodynamic properties and display complex behavior that is sensitive to their size and temperature (see refs. [1 6] for examples). Both classical and quantum J clusters have been studied very extensively in the last several decades. Even without taking into account the quantum effects these systems have always made accurate simulations challenging due to the high complexity of their potential energy landscapes. Consequently, these studies have stimulated the development of various numerical techniques, such as onte Carlo or olecular Dynamics, or various algorithms of global optimization, etc. With some notable exceptions, for sizes n 1 the global minima structures of J n clusters are dominated by the icosahedral motif [7, 8], while the closed-packed structures start to be favorable only around n 1 5 [9]. An icosahedral structure may be characterized by the existence of a complete icosahedral core (n = 13, 55, 147,...) surrounded by an incomplete overlayer, which can adopt two different packings: the anti-ackay (n =14-3, 56-81,85,...) or ackay (n =31-55, 82-147,...). There are several exceptions of clusters (n =74-76, 98, 12-14,...) that adopt non-icosahedral structures as their global minimum configuration. At finite temperatures local energy minima start to contribute to the equilibrium properties of the cluster. The presence of different structural motifs that may be favorable for certain ranges of the parameters (n or T ), gives rise to size- or temperature-induced structural transitions, resulting from competition between the energy and entropy. Even for relatively small sizes (n 2) the number of the thermodynamically relevant minima for a J n cluster is already very large and keeps increasing exponentially with size. oreover, these minima are often separated by large barriers leading to severe sampling problems. The Replica-Exchange onte-carlo method [1] is commonly employed to deal with such problems. The first successful application of the latter to a J cluster (n = 38) had been reported relatively recently [2]. Despite an apparent superiority of the novel methodology, at the time this calculation seemed to be heroic. ore recently a series of Replica-Exchange calculations covering a wide range of sizes was carried out [3 6]. Notwithstanding the difficulties of a classical simulation of J clusters, accurately modeling the quantum effects is further challenging. Calculations using Path Integral onte-carlo (PIC) methods, when converged, provide essentially exact results. However, when trying to accurately describe structural transformations using PIC, the computation times can be excessively long and increase dramatically with decreasing temperature. Because in the case of a structural transition at least two basins of attraction, usually separated by a large energy barrier, must be sampled, regardless of whether the quantum effects are included in the simulation or not, the sampling problem already exists. In the PIC framework this problem is alleviated by the presence of an additional (short) time scale associated with the path variables describing the shape of the path. Strong quantum effects (such as quantum delocalization and/or tunneling), as e.g. in the helium systems, may effectively remove the ergodicity problem. However, it is the intermediate quantum regime, which is most challenging for accurate quantum simulations, especially for systems that undergo structural transformations at low temperatures. Examples of such systems include neon clusters, and, very likely, hydrogen clusters. Perhaps not surprisingly, Ne 13 cluster is the largest neon system undergoing a low-temperature (T 1K) structural change, for which a converged heat capacity curve computed by the PIC was reported [11]. arger neon clusters, e.g. Ne 39, which undergo low-temperature structural changes [12], seem at the moment too difficult to be treated accurately by PIC [13]. As yet another example, we believe that the recent attempt [14] to compute the ground states of (H 2 ) n and (D 2 ) n clusters using a variant of the PIC methodology seem to fail to accurately characterize the size-induced structural changes in these systems. In another paper [15] the Diffusion onte Carlo methodology was used to also compute the ground states of (H 2 ) n clusters, but the results are in disagreement with ref. [14] and are possibly non-converged as well. A rare gas atomic system or a molecular system (e.g., a hydrogen cluster) interacting through van der Waals forces can be mapped to a model J n system with a particular value of the de Boer quantum delocalization length Λ = ( / mɛ)/σ, where m is the particle mass, ɛ and σ are the parameters of the J pair potential, [ U(r) = 4ɛ (σ/r) 12 (σ/r) 6].

2 Λ is an effective measure of delocalization of the ground state wavefunction relative to the interatomic distance. Although only particular discrete Λ-values can be realized in nature, by looking at the properties of J n cluster as a function of continuously varying Λ one can gain new insights, which are not available otherwise. Faced with the above difficulties, for weakly quantum systems, it seems to be practical to use approximate methods that have better sampling properties, but may still account for the quantum effects adequately. One such method is the harmonic approximation (HA), which can be used for estimation of the ground state energies and wavefunctions for the entire range of Λ. As an example, ref. [9] presents a HA analysis of the ground state structures of J clusters corresponding to different rare gas atomic clusters. While the HA is straight-forward to implement and it may be adequate for nearly classical cases, as e.g., xenon (Λ.1), krypton (Λ.16) or argon (Λ.3), in a more quantum regime, e.g., corresponding to neon (Λ =.95), the HA predictions could be very crude [16, 17]. Our method of choice is the Variational Gaussian Wavepacket (VGW) method [12, 16, 17]. The VGW is exact for a harmonic potential, while it is manifestly approximate for a general anharmonic potential, yet, this method demonstrated its practicality, specifically, for the case of Ne 38 [16], for which the VGW energies agreed very well with those computed by the PIC method. In our previous work [17] the VGW method was applied to investigate the J 31 45 clusters, which, with the exception of J 38, posses ackay global energy minima. It was shown that the quantum delocalization always induces the ackay anti-ackay ( ) transition. Consequently, a phase diagram showing the stability ranges of the two structural motifs was constructed. The present paper extends the ground state calculations of J n clusters up to n = 147, the largest size of a three-layer cluster with icosahedral symmetry. As n increases the VGW calculations become increasingly more time-consuming. In order to reduce the computational burden only the following sizes were included: n = 31-55, 6, 65, 7, 75-77, 8, 82-9, 95, 98, 1, 12-14, 15, 11, 115, 12, 125, 13, 135, 14, 145, 147. J 82 is the smallest cluster with ackay global energy minimum and 55-atom core. The sizes, n = 38, 75 77, 98, 12 14, include all the cases of non-icosahedral global energy minima, namely: truncated octahedron (O h ) for n = 38, arks decahedra ( ) for n = 75 77, 12 14, and tetrahedron (T h ) for n = 98. The VGW method for estimating the cluster ground state energy and structure is based on minimization of the energy functional, Ψ Ĥ Ψ E = Ψ Ψ 1, (1) with the trial wavefunction (or VGW) chosen in the form [ Ψ(q, G, γ) = exp 1 ] 2 (r q)t G 1 (r q) + γ, (2) where the variational parameters are: the 3n 3n real symmetric matrix G, real 3n-vector q, and factor γ. For each cluster size the Replica-Exchange method [1] (using the system potential energy) is utilized to sample the configuration space. Every once in many onte- Carlo steps a configuration is selected from one of the random walks for (quantum) quenching to yield a stationary VGW that minimizes the energy functional (1), given the Gaussian constraints (2). Typically the quenched configuration is found in the vicinity of the initial configuration. The energy of the particular stationary solution is then compared to the energies of the solutions found previously. Configurations of sufficiently low energy are retained to be used later to produce energy correlation diagrams. The latter consist of the energy curves E i (Λ) generated for a range of Λ using all the retained configurations. Given a value of Λ, the lowest energy configuration is then assumed to represent the the cluster ground state. ore details of the method can be found in ref. [17] and in a forthcoming paper. The analysis of all the correlation diagrams revealed several distinct structural motifs that turned to be stable within certain ranges of Λ. Those include the ackay (), anti-ackay () and liquid-like () structures, which are generic for all clusters considered. In addition, for clusters with non-icosahedral global energy minima, the corresponding structural motifs remain stable within certain (sometimes very small) ranges of Λ values. One such example is the J 12 cluster, which in addition to the two generic transitions ( and ) undergoes a transition at Λ 4.1 1 3. The corresponding correlation diagram, which only shows one energy curve per structural motif, is given in Fig. 1. The four ground state structures involved in this figure are shown in Fig. 2. (We use reduced variables for energy and temperature, i.e. the energy is measured in units of [ɛ] and temperature, in units of [ɛ/k B ].) All the identified Λ-values corresponding to the structural transformations in J n clusters were used to produce the n Λ phase diagram (see Fig. 3), which shows the stability regions for different structural motifs as a function of n and Λ. The size-temperature (n T ) phase diagram of the classical (Λ = ) J n clusters is also shown for comparison. The latter diagram was constructed based on the data from refs. [4, 6]. There is a striking resemblance between the two diagrams, in which one can find the same structural motifs stable over certain ranges of parameters. Structurally, motifs denoted by the same symbol are indistinguishable. Note though that unlike the high-temperature liquid configurations in the classical diagram, the quantum liquid-like structures, although being disordered, are completely frozen. The quantum-induced ( order-disorder ) transition in the present case has nothing to do with quantum tunneling, no does it involve any super-fluidity as the exchange symmetries are not taken into account within the VGW approximation. Quantum delocalization (as well as the temperature

3 increase) always stabilizes the anti-ackay relative to the ackay motif. Further increase of Λ (or T ) eventually destabilizes the anti-ackay structure resulting into the least-ordered liquid-like structure. However, for small, two-layer clusters, we were not able to clearly distinguish between liquid-like and anti-ackay structures. The probable reason is that the 13-atom icosahedral core is typically not unique for an anti-ackay structure (i.e., there may be several 13-atom sub-clusters Λ.4? He H 2 D 2 J 12 38 O h 75-77 98 T 12-14 (x1) h Dh 4 6 8 1 12 14 n Ne Ar Xe E i (Λ).4 Λ FIG. 1: Correlation diagram for J 12 showing the relative energies, E i(λ) := [E i(λ) E o(λ)] /n, as a function of Λ for four quantum states, each being the lowest-energy state for one of the four different structural motifs. i = corresponds to the global classical energy minimum. The structural transitions occur, where the curves intersect each other. T 38 O h 75-77 98 12-14 4 6 8 1 12 14 n T h.41 15 22 FIG. 3: (Color online) The n Λ phase diagram (top) showing the stability ranges for the ground state structures of quantum J n clusters (n = 3 147). The points are connected for better visualization. The bar for n = 98 is scaled by a factor of 1, also for better visualization. The question-mark indicates the difficulty in distinguishing between anti-ackay and liquid-like structures for two-layer clusters. Bottom: the corresponding n-t phase diagram, which bears similarity to the top figure. This diagram was constructed using the data from refs. [4, 6].12 54 21 FIG. 2: The four configurations of J 12 cluster involved in Fig. 1. The darker color is used to identify the core atoms in the ackay and anti-ackay structures. The Λ and T axes identify the values of the corresponding structural transitions for respectively quantum (T = ) and classical (Λ = ) cases. T having the same arrangement of the complete icosahedron), while the 12-atom coordination is also characteristic of liquid-like structures. (This problem may be resolved in the future if an appropriate order parameter is found.) A similar difficulty was encountered when analyzing the temperature-induced structural changes in classical J n clusters [4], for which the liquid and anti- ackay structural motifs could clearly be distinguished only for three-layer clusters (55 < n 147), the existence of the well-ordered complete-icosahedral 55-atom core being a unique property of the icosahedral motif.

4 For the latter clusters the heat capacity curve displays a sharp peak at the temperature of the transition (also identified as the core melting transition ). For the classical two-layer J clusters only the transition can be identified unambiguously and is always accompanied by a sharp heat capacity peak. Furthermore, this peak changes gradually as a function of cluster size and continues into the size region (n > 45), where the twolayer anti-ackay structures do not exist. That is, if one follows the transition as a function of cluster size, starting with small sizes (n 31), it gradually changes in character, so that for larger clusters it is perhaps more representative of just cluster melting. Fig. 4 shows two examples of anti-ackay ground state structures for n = 115 (the stability range is Λ [116, 54]) and n = 147 (Λ [12,.4121]). Although the largest structure with a single-anti-ackay overlayer surrounding the 55-atom ackay icosahedral core corresponds to J 115 (having a perfect, nearly spherical shape), for n > 115 structures with double-anti-ackay overlayer can become energetically favorable and are easily identified, as in the case of J 147. Note also ref. [5], where two temperature-induced structural transformations were observed for the classical J 39 cluster, whose global minimum is a complete four-layer ackay icosahedron. In the latter work the lower-temperature transition was interpreted as surface roughening of the overlayer surrounding the 147-atom ackay icosahedral core. n=115 n=147 n=147 side view FIG. 4: Images of anti-ackay structures of J 115 and J 147, each characterizing the ground state of the corresponding cluster stable over certain Λ-range. n = 115 is the largest size with 55-atom ackay icosahedral core and anti-ackay overlayer. For n > 115 the extra atoms fill the fourth layer forming a double-anti-ackay overlayer, which can be seen in the side view. For the lack of both a rigorous and simple method, one is often tempted to incorporate the quantum effects into the molecular dynamics simulations by using an effective temperature. The comparison of the classical n T (Λ = ) and quantum n Λ (T = ) diagrams suggests that such a mapping exists on a certain qualitative level, although it is not universal. For example, in the n T (Λ = ) diagram the and transition temperatures merge for n 14, while in the n Λ (T = ) diagram the and transitions for the same size range occur at different values of Λ. Another interesting parallel exists between the present n Λ diagram for the quantum J n clusters and n ρ diagram for the global minima structures of the classical orse clusters [18, 19], where the orse pair potential is U(r) = ɛ exp [ρ(1 rσ)] {exp [ρ(1 r/σ)] 2} with ρ defining the range of the potential, the larger the ρ-value the shorter the range. Increasing the range of the potential destabilizes the ackay (relative to the anti- ackay) structures. Further increase of the range eventually makes the liquid-like structures energetically most favorable. In view of this comparison we conclude that the quantum delocalization, besides making the effective pair interaction softer, also increases its range. In conclusion, we believe that for the weakly quantum regime, which at least includes the neon clusters, our results are generally correct, but may only be qualitatively correct for the more quantum regime, perhaps starting with hydrogen clusters. When the quantum effects are sufficiently strong, the Gaussian approximation may fail to adequately describe the delocalized ground state wavefunction. oreover, it is believed that the particle exchange symmetry becomes important, at least for hydrogen clusters, and may even result in super-fluidity [14], while this effect is completely ignored within the present approximation. At the same time, due to the sampling problems, the Path-Integral and related methods may be hard-to-apply in the weakly quantum regime, while they become much better suited for the strongly quantum regime. Acknowledgment. This work was supported by the NSF grant CHE-8918. Contribution of Pavel Frantsuzov in writing a significant part of the code used for the present calculations is acknowledged. We also thank David Wales for many useful discussions. [1] J. P. K. Doye,. A. iller, D. J. Wales J. Chem. Phys. 11, 6896 (1999); J. P. K. Doye,. A. iller, D. J. Wales J. Chem. Phys. 111, 8417 (1999) [2] J. P. Neirotti, F. Calvo, D.. Freeman, J. D. Doll J. Chem. Phys. 112, 134 (2); F. Calvo, J. P. Neirotti, D.. Freeman, J. D. Doll J. Chem. Phys. 112, 135 (2) [3] Frantz, D. D. J. Chem. Phys. 115, 6136 (21) [4] V. A. andelshtam, P. Frantsuzov, J Chem. Phys. 124, Art. No. 24511 (26) [5] E.G. Noya, J.P.K. Doye, J. Chem. Phys. 124, 1453 (26) [6] V.A. Sharapov, V. A. andelshtam, J. Phys. Chem. A 111, 1284 (27) [7] J. A. Northby, J. Chem. Phys. 87, 6166 (1987) [8] R.H. eary, J. Global Optimization 11, 35 (1987) [9] J.P.K. Doye, F. Calvo, J. Chem. Phys. 116, 837 (22) [1] C. J. Geyer, in Computing Science and Statistics: Proceedings of the 23rd Symposium on the Interface, ed. E.

5. Keramigas, (Interface Foundation: Fairfax, 1991), pp. 156-163; K. Hukushima, K. Nemoto J. Phys. Soc. Jpn. 65, 164 (1996) [11] C. Predescu, D. Sabo, J.D. Doll, D.. Freeman J. Chem. Phys. 119, 12119 (23); D. Sabo, C. Predescu, J.D. Doll, D.. Freeman, J. Chem. Phys. 121, 856 (24) [12] P.A. Frantsuzov, D. eluzzi, V. A. andelshtam, Phys. Rev. ett. 96, 11341 (26) [13] C. Predescu, Private communication. [14] J. E. Cuervo, P.-N. Roy, J Chem. Phys. 128, 22459 (28) [15] R. Guardiola, J. Navarro, C ent. Eur. J. Phys. 6, pp. 33-37 (28) [16] C. Predescu, P. A. Frantsuzov, V. A. andelshtam, J Chem. Phys. 122, 15435 (25) [17] J. Deckman, P. A. Frantsuzov, V.A. andelshtam Phys. Rev. E 77, 1 (28) [18] Doye, J. P. K.; Wales, D. J.; Berry, R. S. J. Chem. Phys. (1995, 13, 4234. [19] Doye, J. P. K. and Wales, D. J. Science. (1996), 271, 484.