SIMULATIONAL ANALYSIS OF GLASSES PREPARED VIA DIFFERENT INTERATOMIC POTENTIALS

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

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

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

Competition between face-centered cubic and icosahedral cluster structures

Rethinking atomic packing and cluster formation in metallic liquids and glasses

MOLECULAR DYNAMICS SIMULATION OF HETEROGENEOUS NUCLEATION OF LIQUID DROPLET ON SOLID SURFACE

Local order in a supercooled colloidal fluid observed by confocal microscopy

T. Egami. Model System of Dense Random Packing (DRP)

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

Physics of Materials: Classification of Solids On The basis of Geometry and Bonding (Intermolecular forces)

A MOLECULAR DYNAMICS SIMULATION OF WATER DROPLET IN CONTACT WITH A PLATINUM SURFACE

VIRTUAL LATTICE TECHNIQUE AND THE INTERATOMIC POTENTIALS OF ZINC-BLEND-TYPE BINARY COMPOUNDS

AP* Chapter 10. Liquids and Solids. Friday, November 22, 13

The samples used in these calculations were arranged as perfect diamond crystal of

Glass Formation and Thermodynamics of Supercooled Monatomic Liquids

Monte Carlo Study of the Crystalline and Amorphous NaK Alloy

A MOLECULAR DYNAMICS SIMULATION OF A BUBBLE NUCLEATION ON SOLID SURFACE

Structural order in glassy water

Class 29: Reciprocal Space 3: Ewald sphere, Simple Cubic, FCC and BCC in Reciprocal Space

A Molecular Dynamics Simulation of a Homogeneous Organic-Inorganic Hybrid Silica Membrane

Classification of Solids, Fermi Level and Conductivity in Metals Dr. Anurag Srivastava

Chapter 2: INTERMOLECULAR BONDING (4rd session)

Amorphous state studies with the Gaussian core model

Mal. Res. Soc. Symp. Proc. Vol Materials Research Society

Critical Temperature - the temperature above which the liquid state of a substance no longer exists regardless of the pressure.

States of Matter; Liquids and Solids. Condensation - change of a gas to either the solid or liquid state

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

From Atoms to Materials: Predictive Theory and Simulations

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

Phase Transitions in Relaxor Ferroelectrics

EGN 3365 Review on Bonding & Crystal Structures by Zhe Cheng

Supporting information for: Anomalous Stability of Two-Dimensional Ice. Confined in Hydrophobic Nanopore

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

Analysis of the simulation

Imperfect Gases. NC State University

STRONG CONFIGURATIONAL DEPENDENCE OF ELASTIC PROPERTIES OF A CU-ZR BINARY MODEL METALLIC GLASS

Local Electronic Structures and Chemical Bonds in Zr-Based Metallic Glasses

MOLECULAR DYNAMICS STUDY OF THE NUCLEATION OF BUBBLE

Liquids & Solids: Section 12.3

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

Chapter 10. Liquids and Solids

Dynamics of Supercooled Liquids The Generic Phase Diagram for Glasses

A local order metric for condensed phase environments

MOLECULAR DYNAMICS SIMULATION OF VAPOR BUBBLE NUCLEATION ON A SOLID SURFACE. Tatsuto Kimura and Shigeo Maruyama

2. As gas P increases and/or T is lowered, intermolecular forces become significant, and deviations from ideal gas laws occur (van der Waal equation).

Percolation structure in metallic glasses and liquids

Molecular Dynamics Study on the Binary Collision of Nanometer-Sized Droplets of Liquid Argon

Introductory Nanotechnology ~ Basic Condensed Matter Physics ~

Introduction to Solid State Physics or the study of physical properties of matter in a solid phase

3.091 Introduction to Solid State Chemistry. Lecture Notes No. 5a ELASTIC BEHAVIOR OF SOLIDS

Disordered Hyperuniformity: Liquid-like Behaviour in Structural Solids, A New Phase of Matter?

E x p e r i m e n t a l l y, t w o k i n d s o f o r d e r h a v e b e e n d e t e c t e d

Phys 460 Describing and Classifying Crystal Lattices

ECE 474: Principles of Electronic Devices. Prof. Virginia Ayres Electrical & Computer Engineering Michigan State University

2. As gas P increases and/or T is lowered, intermolecular forces become significant, and deviations from ideal gas laws occur (van der Waal equation).

New from Ulf in Berzerkeley

Analysis of the Lennard-Jones-38 stochastic network

CHEM Principles of Chemistry II Chapter 10 - Liquids and Solids

Molecular dynamics simulations of EXAFS in germanium

1.3 Molecular Level Presentation

T=293 K is shown in Fig. 1. The figure shows that in. Crystalline fragments in glasses ' ' PHYSICAL REVIEW B VOLUME 45, NUMBER 13

Study of Ni/Al Interface Diffusion by Molecular Dynamics Simulation*

STRUCTURAL AND MECHANICAL PROPERTIES OF AMORPHOUS SILICON: AB-INITIO AND CLASSICAL MOLECULAR DYNAMICS STUDY

Chapter 10: Liquids and Solids

Chapter 12. Insert picture from First page of chapter. Intermolecular Forces and the Physical Properties of Liquids and Solids

IAP 2006: From nano to macro: Introduction to atomistic modeling techniques and application in a case study of modeling fracture of copper (1.

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

Structural and Morphological Transitions in Gold Nanorods: A Computer Simulation Study

Chapter 10. Liquids and Solids

CRYSTAL STRUCTURE, PHASE CHANGES, AND PHASE DIAGRAMS

Liquids and Solids. H fus (Heat of fusion) H vap (Heat of vaporization) H sub (Heat of sublimation)

University, Ames, Iowa, 50011, USA PLEASE SCROLL DOWN FOR ARTICLE

Solids / Crystal Structure

Chemistry 801: Nanostructured Materials and Interfaces

Intermolecular Forces, Liquids, Solids. IM Forces and Physical Properties

Everything starts with atomic structure and bonding

Jamming and the Anticrystal

Physics 211B : Problem Set #0

Chapter 5. Effects of Photonic Crystal Band Gap on Rotation and Deformation of Hollow Te Rods in Triangular Lattice

Phase Field Crystal (PFC) Model and Density Functional Theory (DFT) of Freezing

arxiv:cond-mat/ v2 7 Dec 1999

Heterogenous Nucleation in Hard Spheres Systems

CHAPTER ELEVEN KINETIC MOLECULAR THEORY OF LIQUIDS AND SOLIDS KINETIC MOLECULAR THEORY OF LIQUIDS AND SOLIDS

UNIT I SOLID STATE PHYSICS

Supplementary Information for Observation of dynamic atom-atom correlation in liquid helium in real space

Chem 728 Introduction to Solid Surfaces

Precursors of a phase transition in a simple model system

arxiv: v1 [cond-mat.soft] 26 Feb 2015

Velocity cross-correlations and atomic momentum transfer in simple liquids with different potential cores

From the ideal gas to an ideal glass: a thermodynamic route to random close packing.

THREE-DIMENSIONAL PACKING OF PERFECT TETRAHEDRA

MSE 102, Fall 2014 Midterm #2. Write your name here [10 points]:

Self-assembly of the simple cubic lattice with an isotropic potential

Quantum Monte Carlo simulation of spin-polarized tritium

Ionic Bonding. Chem

The yielding transition in periodically sheared binary glasses at finite temperature. Nikolai V. Priezjev

Close-packing transitions in clusters of Lennard-Jones spheres

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

Disordered Energy Minimizing Configurations

Review of Last Class 1

Transcription:

Journal of Optoelectronics and Advanced Materials Vol. 7, No. 4, August 2005, p. 1915-1922 SIMULATIONAL ANALYSIS OF GLASSES PREPARED VIA DIFFERENT INTERATOMIC POTENTIALS Y. Sano, J. K ga, F. Yonezawa Department of Physics, Keio University, 3-14-1 Hiyoshi, Kohoku-ku, Yokohama 223-8522 Japan All glasses and amorphous solids have common characteristics; That is to say, they have no long-range order (LRO), but maintain the short-range order (SRO) reflecting the types of cohesion. In order to classify glasses beyond this common feature, therefore, it is necessary to introduce some other measure by which we can distinguish different atomic structures of glasses. Under this situation, it is the purpose of the present article to propose such a quantity. To this end, we first carry out, by means of molecular-dynamics (MD) method, rapid-quench simulations to construct glassy structures of the Lennard-Jones (LJ) systems and of the systems with the Stillinger-Weber (SW) potentials. Secondly, we evaluate several physical properties of the glasses thus obtained. Note that the crystalline phase of an LJ system is fcc which is the closest-packing structure, while that of an SW system is bcc where the atoms are more sparsed than in fcc. Among several quantities we calculate, we pay special attention to the pair-distribution function g(r), r being the interatomic distance, the bond orientational parameter W 6, and the frequency spectrum F(ω),ω being the vibrational frequency. The behavior of g(r) confirms the absence of the LRO. On the other hand, the behaviors of W6 for the nearest-neighbor atoms show the existence of the SRO in the form that the nearest-neighbor configurations are icosahedral as expected for these glasses. Interesting results are found in F(ω) in the sense that, in some glasses, the peaks appear in F(ω) at almost the same ω at the peak position in the case of an fcc crystal. From the analyses of W6 for atoms in the second-neighbor shell of each central atom, we find that the origin of the peaks in F(ω) comes from the cubic-like configurations of the atoms in the second-neighbor shell. We discuss this problem in detail and propose that F(ω), which is a macroscopically-observable quantity, is used as a measure to detect the degree of the IRO. (Received July 4, 2005; accepted July 21, 2005) Keywords: Glass, Simulation, Interatomic potential 1. Introduction The absence of long-range order (LRO) and the existence of short-range order (SRO) are characteristic features common to all glasses and amorphous solids. Here, the term LRO is used to indicate that the atomic configurations are specified by the periodicity as found in crystals, while the term SRO is used to denote that the positions of nearest-neighbor atoms around a given atom in a disorderd system are almost identical to those in a corresponding crystal. Therefore the intermediaterange order (IRO) is expected to be a deciding factor in distinguishing one disordered system from another. Here, we use the term IRO as reprensenting the order in the atomic distributions beyond the nearest neighbors. The purpose of the present paper is to propose a way detecting the IRO by studying the local symmetry of atomic configurations beyond the first-nearest neighbor. For this porpose we carry out rapid quench simulations to construct glassy structures by means of molecular dynamics method. The subjects we pay attention to include (1) the glass structure of the Lennard-Jones (LJ) system and Corresponding author: ysano@rk.phys.keio.ac.jp

1916 Y. Sano, J. K ga, F. Yonezawa of the Stillinger-Weber (SW) system, (2) quench rate dependences of the structures both from the point of view of macroscopic properties and microscopic strucrues through the glass transition. The reason why we study the SW system in addition to the LJ system, is to find if there is any common features in glasses of di_erent nature such as those with potentials leading to close-packed structures and those with potentials causing relatively open structures. We calculate several physical properties; the pair distribution function g(r) as structural properties, the frequency spectrum F(ω) for studying vibrational properties, the bond orientational parameter W6 to investigate local symmetries of atomic configurations in the first and second shell around a given atom. In section 2 we describe our model and method, while the results are presented in section 3. Finally summary is given in section 4. 2. Model and method We use the constant-pressure MD simulation method by Andersen [1] to study two model systems composed of 864 spherical particles; in one system, atoms interact with one another via 12-6 Lennard-Jones pair potential and in another system, atoms interact via Stillinger-Weber pair potential modified by Nos e and Yonezawa [2]. Note that an LJ system leads to a closest-packed crystalline structure such as an fcc lattice at low enough temperatures, while an SW system leads to a relatively open crystalline structure such as a bcc lattice. For our SW system, we use the following pair potential which has been modified from the original three-body SW potential [3,4] so as to realize a bcc structure in the crystalline phase. In our simulations, the atoms are placed in a simulation cell with periodic boundary conditions. The calculations are conveniently performed in dimensionless variables by scaling the energy in ε, length in σ mass in M, temperature in ε/k B, pressure in ε/σ 3, and time in (Mσ 2 /ε) 1/2, where ε and σ are the potential parameters of the Lennard- Jones potential, M is the atomic mass, and k B is the Boltzman constant. The magnitudes of the units relevant to the argon are taken from the paper of Dangaard Kristensen [5] ε/k B = 118K, σ = 3.84Å atomic mass M 0 = 0.66 10-22 g, t 0 = 2.44ps, a time step for integration is taken to be t = 23 in reduced unit which corresponds to 0.5 10-14 s for argon. We choose the pressure P = 1. In order to construct different initial conditions, we heat an fcc crystal from T = 0.1 to T = 4.0, and then anneal the system at T = 4.0 for the time duration of various timesteps. For two different potentials we cool the initial systems down to T = 0.2 with three quench rates; Q A = 4.2 10 10 K/s, Q B = 4.2 10 11 K/s and Q C = 4.2 10 12 K/s. 8.56438 1 φ( r) = 0 r r 1 1 [ r < 1.8] exp r 1.8[ r > 1.8] (1) SW LJ V(r ) Fig. 1. The Lennard-Jones potential is presented by dotted curve, while the Stillinger- Weber potential defined by Eq. (1) is shown by solid curve, where r = r/σ. r

Simulational analysis of glasses prepared via different interatomic potentials 1917 3. Results 3.1 Macroscopically observable properties In this subsection, we study some of macroscopically observable properties such as volume V and pair distribution function g(r). The left side of Fig. 2 shows the V -T relation in the LJ system, where V is the volume per atom. For the Q A, a discontinuous decrease of volume is observed between T = 0.3 and T = 0.2, indicationg crystallization through the first-order transition. For higher quench rates Q B and Q C, on the other hand, the volume decreases continuously all the way in the process down to T = 0.2, which shows freezing without forming crystal, suggesting the formation of glasses. In the case of an SW system, the relation is shown in the right side of Fig. 2. With all quench rates Q A, Q B and Q C, the volume decrease continuously down to T = 0.2, suggesting that the systems become glassy state. V V T T Fig. 2. The right figure shows volume versus temperature in the LJ system on quenching with quench rate Q A, Q B and Q C by the dots, the open circles and the solid triangles respectively. The left figure shows the result in the SW system, in which the volume decrease continuously for all quench rates. Fig. 3 (a) and (b) show the pair distribution function g(r) for the quench rates Q A and Q B respectively in the LJ system. The peaks at the low temperature (T = 0.2) in Fig. 2 (a) appear in the positions of the characteristic peaks of an fcc crystal, thus confirming the occurrence of crystallization in addtion to the result of the V -T relation. Fig.3 (b) for the quench rate Q B in the LJ system presents that the atomic configurations are disordered all the temperatures. There appear in particular at low temperatures the splitting of the second peak, which is a characterictic feature of g(r) for glasses. For the quench rate QC as well as for Q A, Q B and Q C in the SW system, g(r) behave the same manner as Fig.3 (b), which confirms the formation of glasses. g(r ) g(r ) r r Fig. 3. The pair distribution function g(r) between temperature T = 0.8 and 0.2. (a) g(r) with quehch rate Q A, for which there appear peaks characteristic of the crystallization at T = 0.2. (b) g(r) with QB, a glass-forming quench rate.

1918 Y. Sano, J. K ga, F. Yonezawa 3.2 Frequency spectrum of vibrational modes In order to investigate the vibrational motions of atoms in the glassy states, we calculate the velocity autocorrelation function (VAF) υ(t), defined by υ( t) = ( 1/ N ) ivi ( t0 + t) vi ( t0) t0 2 v( t0) t0 (2) and the frequency spectrum F(ω) which is the Fourier transformation of the VAF given as follows; 6 F ( ω) = υ( t)cos( ωt) dt, π 0 (3) where v i is the velocity vector of the ith atom at t, and the brackets h i denote the average over initial conditions {t 0 }. Fig. 4 shows the F(ω) of each quench rate in the LJ system. The left side of Fig. 4 presents the results with quench rate QA at T = 0.3 and T = 0.2. Note that according to the results of V - T relation and g(r), the first-order transition takes place at temperature slightly lower than T = 0.3. Though there is no peak at T = 0.3, there appears at T = 0.2 a wide peak around frequency ω = 13 which corresponds to the distinct peak in F(ω), as shown by a broken curve, of an fcc crystal under P = 1.0 and T = 0.2. Q B Q C T=0.4 T=0.4 Q A T=0.5 F( ) T=0.3 T=0.3 F( ) --fcc 0 10 20 30 40 0 10 20 30 0 10 20 30 40 Fig. 4. Figures on the left show frequency spectra F(ω) versus ω for Q A at T = 0.3 and T = 0.2, between which the occurrence of the first-order transition is indicated. The broken curve corresponds to an fcc crystal at T = 0.2 with the same volume as for the Q A. The three figure in the center show the F(ω) for the Q B and on the right the results of Q C. The glass transition occurs at about T = 0.4. The right side of Fig. 4 shows the results with quench rate Q B and quench rate Q C, both of which are glass-forming quench rates and the glass transition occurs at about T = 0.4. There is no peak in the F(ω)at T = 0.4 similar to F(ω) of liquid state. As the temperature decreases, with both

Simulational analysis of glasses prepared via different interatomic potentials 1919 quench rates, the peak appears gradually at ω ~ 13 which corresponds to the peak position of fcc crystal. It shows that the higher the quench rate, the higher the peaks grow up. As for the SW system, the calculated results of F(ω) are presented in Fig. 5, which show the behaviors similar to F(ω) for Q B and Q C in the LJ system. In the SW system, the peaks appear at ω ~ 9, which corresponds to the peak position of a bcc crystal as illustrated by a broken curve in the right-bottom figure. These results of F(ω) are very interesting in following respects; In spite of the fact that, from g(r), the atomic configurations are shown to be disordered, there appears a peak at the same frequency as the peak position of bcc crystal. The appearance of this peak suggests that there exist some aspects of cubic symmetry, which is contradictory to the concluions drawn from g(r). We try to pin down the origin of the peak in F(ω) by investigating the local symmetry in a succeeding section. Q A Q B Q C T=0.4 T=0.4 T=0.4 T=0.3 T=0.3 T=0.3 F( ) --fcc Fig. 5. From left to right, the results of Q A, Q B and Q C in the SW system are presented. The broken curve in the bottom right shows F(ω) of a bcc crystal under P = 1 and T = 0.2. 3.3. Local symmetry of atoms We use the bond orientatilnal parameter W 6 as a quantity to calculate the local symmetry of atoms. This parameter is suitable for distinguishing the local icosahedral symmetry from the local cubic symmetry, which was introduced by Steinhardt et al [6], and modified by Yonezawa et al [7] to study the symmetry formed by a given group of atoms such as atoms in the first and second shell. Since the local idosahedral configuration is expected in disordered sysmtens such as liquid and glasses composed of isotropic atoms, this parameter is useful to examine the local symmetry. In the present paper, we calculate not only W6 parameters for atoms in the nearest-neighbor shell but also W6 parameters for atoms in the second-neighbor shell in order to investigate the symmetries of local atomic configurations beyond the nearest neighbor. Symmetry of the first shell 0 10 20 0 10 20 0 10 20 30 In the nearest-neighbor shell of a given atom, the value of W6 is -0.169 for a perfect icosahedron, while the value of W 6 is about 13 for a cluster with crystalline symmetry, such as a simple cubic cluster, an fcc cluster and a bcc cluster. In order to investigate the local symmetry of atoms, we classify the atoms into three categories [7]; (1) atoms with icosahedral symmetry, where an atom is regarded as having icosahedral symmetry when 5 < W 6 < 0.17; (2) atoms with cubic symmetry, where an atom is regarded as having cubic symmetry when 5 < W 6 < 2; (3) the other atoms.

1920 Y. Sano, J. K ga, F. Yonezawa 9 9 5 4 3 5 4 3 2 2 5 4 3 2 5 4 3 5 4 3 2 5 4 3 2 2 0.2 0.4 0.6 0.8 0.2 0.4 0.6 0.8 T T Fig. 6. The proportions of atoms with the icosahedral (the solid circles) and cubic symmetry (the open diamonds) in the first shell on the temperature reduction estimated from the bond orientational parameters W 6. The left side shows the results in LJ system, while the left side shows that in SW systems; from top to bottom, the processes of Q A, Q B and Q C are shown respectively. Figs. 6 show the proportion of atoms with icosahedral and cubic symmetries as far as the nearest-neighbor shell is concerned. In Figs. 6, the results of the LJ and SW are respectively demonstrated on the left and right. For each case, the process of quench rate Q A, Q B and Q C are illustrated from top to bottom. Let us first look at the top-left figure for Q A in the LJ system. The number of atoms with local icosahedral symmetry decreases suddenly between T = 0.3 and T = 0.2, while the number of atom with local cubic symmetry increases drastically, which reflects the occurence of crystallization around this temperature. On the other hand, we can observe in all the other five figures that the local icosahedral symmetry remains intact all the way down to T = 0.2. This behavior is consistent with the conclusion drawn from the V - T relation and g(r) that a glass is formed in each of these five cases. This persistence of local icosahedral symmetry in these glasses, however, contradicts with the appearance of a peak in F(ω). In order to solve this problem, we study the symmetries with respect to the atomic configurations in the second shell of each atom. Symmetry of the second shell We use the modified W 6 to calculate symmetry formed by atoms in the second shell. The value of the W 6 is -0.169 for the second shell in a perfect icosahedral structure, while that of the W 6 in a perfect cubic structure is 13, this parameter is useful to distinquish the icosahedral and cubic symmetries concerning the atoms in the second shell as well. Fig. 7 show the distributions of W 6 at

Simulational analysis of glasses prepared via different interatomic potentials 1921 T = 0.8, T = 0.4 and T = 0.2 in the second shell where the distribution is taken only over those atoms whose nearest-neighbor atoms form local icosahedral symmetry. Here, T = 0.8, T = 0.4 and T = 0.2 respectively correspond to a liquid state, a temperature near the glass-transition temperature, and a glassy state. The left side in the Fig. 7 represents the results with quench rates Q B and Q C in the LJ system, while the right side shows the results with Q A, Q B and Q C in the SW system. All the shape of the distribution have the peak at W 6 ~ +1 which corresponds to cubic symmetry in the second-neighbor shell. We can see that there are no temperature dependence in both cases. T=0.8 T=0.4 T=0.8 T=0.4 T=0.8 T=0.4 T=0.8 T=0.4 T=0.8 T=0.4 second shell w 6 second shell w 6 Fig. 7. The distribution of W 6 for the second shell at T = 0.8, 0.4 and 0.2, for which the atoms have the local icosahedral symmetry in the first shell. The left side presents Q B and Q C in the LJ system, while the right side presents Q A, Q B and Q C in the SW system. As for the quench-rate dependence, the results show no dependence in the LJ case while a slight dependence is descernible in the SW case. It has been argued in our previous study [8]~[12] that, when the quench rate is very high, the atoms do not have enough time to search for low-energy configuration over a wide space, but instead they are solidified by minimizing energy in the immediate vicinity of each atom, thus yielding the icosahedral configuration around each atom. This is the reason why we have the larger number of atoms with the icosahedral symmetry in the first shell when the quench rate is the higher. This tendency can be seen more clearly in Fig.8 where the number of atoms with icosahedral symmetry in the first shell is presented versus quench rate both for the LJ and SW system. These results however seem to be totally contradictory to the fact that the peak appears in F(ω) of glasses at the same frequency as the typical peak position in F(ω) of an fcc crystal. Another problem about the local icosahedral configuration is that it cannot be packed in a consistent manner over the whole space. These problem are all resolved by our results as shown in Fig. 7,

1922 Y. Sano, J. K ga, F. Yonezawa which illustrates that, even when the local symmetry of atoms in the first shell is icosahedral, the atomic configurations of atoms in the second shell have cubic symmetry. In other words, each cluster including atoms up to the second shell has a shape with cubic symmetry, and accordingly the close packing over a wide space is possible and the peak in F(ω) showing a vestige of cubic crystals is accounted for. Fig. 8. The relations between the quench rate and the proprtion of the icosahedral atoms in the nearest-neighbor shell at the temperature T = 0.2. Open triangles show the results in the LJ system while open circles show the SW system. 4. Summary By using constant-pressure molecular dynamic simulations, we have prepared glasses by quenching liquids. We adopted model systems composed of 864 spherical particles with the Lennard-Jones potential as well as the Stillinger-Weber potential. We have analyzed the structures of the glasses both from the view point of macroscopic and microscopic properties in detail; we have discussed the SRO and LRO by calculating the bond orientational parameter W 6 and the pair distribution function g(r). In particular, interesting results have been found in frequency spectrum F(ω). In spite of the fact that the system is glassy, the peak appears in F(ω) at almost the same frequency ω as the peak position in the case of a cubic crystal both in the LJ system and in the SW system. By calculating W6 for atoms in the second-neighbor shell of each atom, we have found that the peak in F(ω) arises from the cubic-like configurations of the atoms in the second-neighbor shell. By taking the above-described results into account, we propose that the frequency spectra F(ω) of vibrational modes can be used as a measure for the order in the atomic configurations beyond the nearet-neighbor shell of each atom. References [1] H. C. Andersen, J. Chem. Phys. 72(4), 2384 (1980). [2] S. Nos e, S. Sakamoto, F. Yonezawa Zeitschrift fur Physikalische Chemie Nemie Folge, Bd. 156, S.91 (1988). [3] F. H. Stillinger, T. A. Weber J. Chem. Phys 70(11), 4879 (1979). [4] F. H. Stillinger, T. A. Weber J. Chem. Phys 68, 3837 (1978). [5] W. Damgaard Kristensen, J. Non-Cryst. Solids 21, 303 (1976). [6] P. Steinhardt, D. R. Nelson, M. Ronchetti Physical Review B 28, (1983) [7] Fumiko Yonezawa Solid State Physics. 45, (1991). [8] M. Kimura, F. Yonezawa, in Topological Disorder in Condensed Matter (F.Yonezawa and T. Ninomiya, eds.). Springer, Computer Glass Transition, 1983. [9] S. Nos e, F.Yonezawa, Solid State Commun. 56, 1005 (1985). [10] F. Yonezawa, S. Nos e, S. Sakamoto, J. Non-Cryst. Solids 95(96), 83 (1987). [11] F. Yonezawa, S. Nos e, S. Sakamoto, J. Non-Cryst. Solids 95(96), 373, (1987). [12] F. Yonezawa, S. Nos e, S. Sakamoto, Neue Folge 156, 77 (1988).