Atomistic Simulations of Elastic-Plastic Deformation of Amorphous Polymers by Franco Mario Capaldi Submitted to the Department of Mechanical Engineering in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY June 2001 @Massachusetts Institute of Technology 2001. All rights reserved. A A A uthor ..................... Department of Mechanical Engineering May 11, 2001 Certified by................ Mary C. Boyce Professor Thesis Supervisor . I1 C ertified by ...................... I ' K) Gregory C. Rutledge Associate Professor Thesis Supervisor Accepted by. . Ain A.- Sonin Chairman, Department Committee on Grad P fn t MASSACHUSETTS INUSTITUTE OF TECHNOLOGY JUL 1 6 2001 I LIBRARIES r% Atomistic Simulations of Elastic-Plastic Deformation of Amorphous Polymers by Franco Mario Capaldi Submitted to the Department of Mechanical Engineering on May 11, 2001, in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering Abstract As the demand for polymers with superior properties increases, an understanding of the fundamental connections between the mechanical behavior and underlying chemical structure becomes imperative. In this work, the thermo-mechanical behavior and the molecular-level origins of plastic deformation of an amorphous glassy polymer were studied using atomistic simulations. Understanding of the molecular response will aid the development of physics-based continuum level models for these materials. A polyethylene-like molecular network was numerically constructed using a Monte Carlo algorithm and then subjected to uniaxial deformation over a wide range of strain rates and temperatures using Molecular Dynamics. The model exhibits many experimentally observed characteristics such as an initial elastic response followed by yield then volume preserving plastic deformation. The stress response was decomposed into intra and inter molecular components and analyzed throughout deformation. In the glassy regime, activation parameters were calculated in the context of the Eyring Model of flow in a solid. In addition, observations were made of the evolution of chain configuration and the correlation of transitions between dihedral angle states. Below the glass transition, dynamic heterogeneity is observed. It was also observed that mobility, as measured by the transitioning between dihedral angle states, increases during plastic deformation to levels observed at much higher temperatures under zero stress. At temperatures near the glass transition temperature, the mobility approaches levels of undeformed samples at the glass transition temperature. Thesis Supervisor: Mary C. Boyce Title: Professor Thesis Supervisor: Gregory C. Rutledge Title: Associate Professor 2 Acknowledgments I would like to thank Professor Mary Boyce and Professor Gregory Rutledge for their guidance and advice. I am grateful to the National Science Foundation for the fellowship funding they have given me and for their support of this research through NSF Grant No. DMR-98-08941. I would like to thank Dr. Jorgen Bergstrom, Dr. Mark Lavine, Numan Waheed, and Shawn Mowry for valuable discussions concerning various aspects of this work. I would also like to acknowledge my lab mates with whom I have had the pleasure of working and whose company I have enjoyed immensely during the past two years: Mats, Mike, Ethan, Rebecca, Dora, Sauri, Brian, Prakash, Tom, Hang Qi, Jin Yi, Matt, Rami, and Steve. I am especially grateful to my parents, my sister, and Irene for their constant encouragement, support, and their faith in me. 3 Contents 1 2 3 12 Introduction 1.1 The Glass Transition . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 1.2 Deformation of Amorphous Glassy Polymers . . . . . . . . . . . . . . 14 1.3 Eyring Model of Flow of Solids . . . . . . . . . . . . . . . . . . . . . 16 Methodology 19 2.1 Molecular Dynamics Algorithm . . . . . . . . . . 19 2.2 Monte Carlo Algorithm . . . . . . . . . . . . . . . 24 2.3 Generation of Initial Structures . . . . . . . . . . 25 2.4 A Lennard-Jones Fluid . . . . . . . . . . . . . . . 27 2.5 Polyethylene Chain Model . . . . . . . . . . . . . 28 2.6 Correlation of Torsion Angles . . . . . . . . . . . 31 Structure 34 3.1 Initial Network Characterization . . . . . 3.2 The Glass Transition Temperature . . . . . . 44 3.3 Energy Barriers for Transitioning . . . . . . . 46 3.4 Transitions and Free Volume . . . . . . . . . . 52 3.5 Correlation of Transitions along a Chain . . . 57 3.6 Tessellated Volume . . . . . . . . . . . . . . . 60 34 4 Deformation Results 4.1 64 Dynamical Mechanical Analysis . . . . . . . . . . . . . . . . . . . . . 4 64 4.2 5 Uniaxial Deformation . . . . . . . . . . . . . . . . 65 4.2.1 Compression . . . . . . . . . . . . . . . . . 68 4.2.2 Tension . . . . . . . . . . . . . . . . . . . . . . 82 4.3 Evolution of Tessellated Volume . . . . . . . . . . . . . 100 4.4 Strain Rate and Temperature Dependence of Yield Stress . . . . . . 102 4.4.1 Stress Response . . . . . . . . . . . . . . . . . . 102 4.4.2 Activation Parameters . . . . . . . . . . . . . . 105 4.5 Transitions and Yield . . . . . . . . . . . . . . . . . . . 109 4.6 Correlation of Transitions along a Chain . . . . . . 115 117 Conclusion 5.1 Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117 5.2 Deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118 5.3 Configuration and Deformation . . . . . . . . . . . . . . . . . . . . . 119 5.4 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120 5 List of Figures 2-1 Pressure-density response of a Lennard-Jones fluid at 0 = 0.9, N = 500 28 2-2 Pressure-density response of a Lennard-Jones fluid at 0 = 2.0, N = 500 29 2-3 Representation of dihedral angle rotation, bond angle bending and bond stretching. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 . . . . . . . . . . . . . . . . . 32 2-4 Energy as a function of dihedral angle 3-1 Measures of equilibration for sample system with 4500 mers at 100K. The dotted line indicates the time at which the ensemble was switched from NVT to NPT: (a) density as a function of time, the dashed line indicates the equilibrium density (b) change in energy as a function of . . . . . . . . . . . . . . . . . . time (c) stress as a function of time. 3-2 36 Distribution of bond lengths for crosslinked systems, nw-1-8000-160100K, nw-1-8000-160-200K, nw-1-8000-160-450K . . . . . . . . . . . . 37 3-3 Distribution of bond angles for crosslinked system . . . . . . . . . . . 37 3-4 Distribution of dihedral angles for crosslinked system. . . . . . . . . . 38 3-5 Pole Figures for crosslinked system at temperatures of lOOK, 200K, and 450K . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 3-6 Radial distribution function for initial structures at various temperatures. 40 3-7 Distribution of densities at 100K for 40000 mers free chains. Number of Bins/ Length of Side (Angstroms) / Average of atoms per bin and standard deviation of distribution . . . . . . . . . . . . . . . . . . . . 6 42 3-8 Comparison of bond and angle statistics for various systems: (a) bond length distribution, (b) bond angle distribution (c) dihedral angle distribution (d) radial distribution profile. . . . . . . . . . . . . . . . . . 3-9 43 Simulations of polyethylene model at atmospheric pressure with 4000 united atom beads. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 3-10 Diagram of dihedral angle rotation. . . . . . . . . . . . . . . . . . . . 50 3-11 Diagram of transition definition. . . . . . . . . . . . . . . . . . . . . . 51 3-12 Transition rates for a system of 5000 mers arranged in 20 free chains. 51 3-13 Decay of the torsional angle autocorrelation function at various temperatures. ....... ......................... .... ... 52 3-14 Spatial distribution of conformational transitions for a system of 5000 united atoms arranged in 20 free PE chains. Dihedral angles are numbered serially along the chain and the number of transitions at each dihedral angle are plotted against the dihedral angle. The total number of transitions at each temperature is 4043. . . . . . . . . . . . . . . . 53 3-15 Correlation between transition rates and density at various temperatures for a system of 5000 mers arranged in 20 free chains using a bin size of 6.31A x 6.31A x 6.31A. Each bin contains an average of 10.6 m ers. . . . .. . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . 55 3-16 Correlation between transition rates and density at various temperatures for a system of 5000 mers arranged in 20 free chains using a bin size of 5.23A x 5.23A x 5.23A. Each bin contains an average of 5.84 m ers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 3-17 Dihedral angle trajectory for a single, active dihedral angle . . . . . . 58 3-18 Neighbor Correlation for system of 5000 united atoms arranged in 20 free chains at various temperatures. . . . . . . . . . . . . . . . . . . . 59 3-19 M ap of dihedral angles . . . . . . . . . . . . . . . . . . . . . . . . . . 60 3-20 Dihedral angle value over time for active angle in networked system at 100K . . . . .. . . . . . .... .. . . . . . . . . . . . . . . . . . . . 61 3-21 T = lOOK, networked system. . . . . . . . . . . . . . . . . . . . . . . 62 7 3-22 Tessellated volume distribution at various temperatures for system of 5000 monomers arranged in 20 free chains. . . . . . . . . . . . . . . . 62 4-1 Experimental graph of DMA curve for HDPE [15] . . . . . . . . . . . 66 4-2 Estimated modulus for i = 1 x 10' 0 /s and = 5 x 10 10 /s for a system of 7775 united atoms arranged in 160 networked chains. . . . . . . . . 4-3 Estimated modulus for various strain rates at lOOK and 200K for a . . . system of 7775 united atoms arranged in 160 networked chains. 4-4 66 Compressive stress versus strain for uniaxial compression for = 67 1 x 1010/s and T = lOOK for a system of 7775 united atoms arranged in 160 networked chains . . . . . . . . . . . . . . . . . . . . . . . . . . . 4-5 Partitioning of compressive stress for uniaxial compression along load direction . 4-6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 Partitioning of normalized energy change per united atom versus true strain in com pression . . . . . . . . . . . . . . . . . . . . . . . . . . . 4-8 70 Partitioning of normalized compressive stress as sample is loaded in compression along the x-axis. 4-7 69 72 Comparison of structural statistics for compressive true strains of 0, 0.2, and 1.0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 Density versus true strain for uniaxial compression . . . . . . . . . . 74 4-10 Schematic of Lennard Jones potential and gradient. . . . . . . . . . . 75 4-9 4-11 Compressive stress versus true strain for constant volume compressive deformation with i = 1 x 1010/s and T = 100K 4-12 Chain statistics versus true strain in compression . . . . . . . . . . . . 76 . . . . . . . . . . . 77 4-13 Transition rate versus true strain for uniaxial compression 4-14 Pole figures for uniaxial compression . . . . . . 78 . . . . . . . . . . . . . . . . . . 79 4-15 Line profiles of three dimensional pair distribution for monomers along different chains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 4-16 Diagram of three dimensional distribution profile slices. . . . . . . . . 82 8 4-17 Three-dimensional pair distribution for monomers on different chains w ith c = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 4-18 Three-dimensional pair distribution for monomers on different chains w ith c = 0.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 4-19 Three-dimensional pair distribution for monomers on different chains w ith E = 1.0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85 4-20 Line profile of three dimensional pair distribution for monomers along the sam e chain. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86 4-21 Three-dimensional pair distribution for monomers on the same chain w ith c = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 4-22 Three-dimensional pair distribution for monomers on the same chain w ith c = 0.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 4-23 Three-dimensional pair distribution for monomers on the same chain w ith c = 1.0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4-24 Schematic of chain interactions along compressed axis. . . . . . . . . 89 90 4-25 Tensile stresses versus true strain for uniaxial tension with i = 1 x 1010/s and T = 100K for system of 7775 united atoms arranged in 160 networked chains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 4-26 Partitioning of tensile stresses versus true strain for uniaxial tension loaded in the x-direction. . . . . . . . . . . . . . . . . . . . . . . . . . 92 4-27 Density versus true strain for uniaxial tension . . . . . . . . . . . . . 93 4-28 Partitioning of energy change per united atom versus true strain for tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 4-29 Comparison of structural statistics for tensile true strains of 0, 0.2, and 1 .0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4-30 Chain statistics versus true strain for uniaxial tension. 95 . . . . . . . . 96 4-31 Transition rate versus true strain for uniaxial tension. . . . . . . . . . 97 4-32 Pole figures for uniaxial tension . . . . . . . . . . . . . . . . . . . . . 98 4-33 Tessellated volume distribution at various strains for system of 5000 mers tested in uniaxial tension at lOOK for a strain rate of 1 x 10 10 sec-1. 10 0 9 4-34 Tessellated volume distribution at various strains for system of 40000 mers arranged in 160 free chains tested in uniaxial compression at lOOK for a strain rate of 1x10' 0 s- 1 . . . . . . . . . . . . . . . . . . . . . . . 101 4-35 Stress versus strain for various temperatures for a system of 7775 united atoms arranged in 160 networked chains. . . . . . . . . . . . . . . . . 103 4-36 Change in density versus strain for various temperatures for a system of 7775 united atoms arranged in 160 networked chains. . . . . . . . . 103 4-37 Change in radial distribution function during the application of compressive deformation at 450K. . . . . . . . . . . . . . . . . . . . . . . 104 4-38 Chemical Structure of (a) polystyrene, (b) polyethylene-terephthalate, (c) polycarbonate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 4-39 Summary of yield stress obtained through simulations of uniaxial compression for various initial networked structures, strain rates, and temperatures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 4-40 Transition rates as a function of true strain. . . . . . . . . . . . . . . 112 4-41 Number of transitions per dihedral angle per angstrom of deformation. 113 . . . . . . . . . . . 114 4-43 Neighbor Correlation . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 4-42 Stress as a function of strain for various systems. 10 List of Tables 2.1 Potential Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1 General Structure of Chains where N is the number of mers in the sim- 32 ulation, M is the number of mers per chain, C is the number of chains, < dee > is the average end to end distance, P is given by equation 3.1, % Trans is the percentage of dihedral angles in the trans configuration, F2 , F3 and F4 are the number of crosslinks with a functionality of 2, 3, and 4 respectively 3.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . Crosslinked vs Free Chain Systems. The equilibrium bond length and angle are 1.53 A and 110.01 degrees respectively. . . . . . . . . . . . . 3.3 41 44 Glass transition temperature, T., and linear thermal expansion coefficient, LTEC, for various molecular weights. . . . . . . . . . . . . . . . 46 . . . . . . 49 3.4 Calculated Activation Energy for Dihedral Angle Switches 4.1 Activation parameters for a selection of glassy polymers [1] . . . . . . 105 11 Chapter 1 Introduction As the demand for polymers with superior properties increases, so does the need for an understanding of the relationship between mechanical behavior and underlying chemical structure. The challenge lies in the fact that mechanical behavior in polymeric materials is governed by phenomena occurring over a wide range of length and time scales. Atomistic simulations may provide a glimpse into detailed material behavior on length and time scales which are not accessible through current experimental techniques. In the present work, information obtained from atomistic simulations is used to develop a better understanding of the molecular response of amorphous polymers to deformation. Simulations of a polyethylene-like molecule subjected to uniaxial deformation are carried out over a wide range of strain rates and temperatures. From this data, activation volumes were calculated. This data was augmented by observations of the evolution of chain configuration, orientation, and molecular packing during deformation. 1.1 The Glass Transition The glassy state is a non-equilibrated state arising from the rapid cooling of a polymer melt. The glassy structure has excess volume and enthalpy with respect to the equilibrium structure at the same temperature. As the melt is cooled, molecular kinetics 12 slow such that the liquid structure is unable to relax to its lower energy crystalline state. The resulting glass maintains a disordered structure characteristic of a liquid while it exhibits mechanical properties similar to those of a crystalline solid. The highly constrained polymer chains are unable to fully explore their configurational space during cooling below the glass transition temperature resulting in an inhomogeneous microstructure with varying degrees of disorder. There are regions possessing lower local densities. The chain segments in these regions are packed less tightly and due to the diminished constraint can explore their configuration space more readily, i.e. they have higher mobility. Although the molecular kinetics may be slow, thermal motion results in continued relaxation at temperatures above absolute zero. If given sufficient time, further relaxation would occur bringing the structure closer to its equilibrium. Therefore, it is important to characterize the transition temperature and the associated time scale for a glassy structure. Attempts to correlate molecular motions with the glass transition through atomistic simulation have shown both the weaknesses and strengths of these methods. Gee and Boyd [13] have shown that the glass transition is strongly dependent on the potential used to describe the dihedral angle. In addition, the cooling rates used in atomistic simulations are much too large to allow for accurate prediction of glass transition temperatures at experimental cooling rates. Fortunately, the qualitative chain dynamics seem to be consistent across a wide range of model force fields [33, 5]. The transition of dihedral angles from trans to gauche states has been the subject of many studies. Studies have found that although transitions in the melt are evenly distributed across all dihedral angles, these transitions become heterogeneously distributed below the glass transition temperature [4]. Relatively few dihedral angles undergo the majority of transitions while other dihedral angles remain constant. This illustrates the heterogeneity of segmental mobility which is "frozen" into the glassy structure. Correlations between transitions in the glass and the melt which occur on a polymer chain have shown strong correlations between dihedral angles separated by 2 bonds which suggests a kink type motion is 13 dominant [33]. Also self correlation of transitions has been shown to become more significant as the temperature falls below the glass transition temperature making it a good indicator of vitrification. Many studies have found that as the temperature falls below the glass transition temperature, the dihedral angle transition rates decrease dramatically. But a study by Liang et al. [23] shows that conclusions drawn from transition rates may depend on the specific definition of a transition. Choosing definitions which encompass the entire range of phi leads to overestimation of transition rates by counting short lived changes between very high energy states. In the literature one finds a wide variety of definitions. Atomistic simulations have also allowed scientists to probe material states which are not experimentally accessible. For example, Karatasos, Adolf, and Hotston [21] found increasing density at constant temperature results in increased correlation between neighboring torsion angles above the glass transition temperature while decreasing the overall transition rates. Yang, et al. [46] also showed that the glass transition temperature increased with the application of pressure or an increase in density. This phenomena has also been observed experimentally by McKinney et al. [27]. Bharadwaj et al. [4] found that increasing pressure increased the torsional self-correlation. In both of these cases the glass transition was determined from the change in slope of the specific volume versus temperature. 1.2 Deformation of Amorphous Glassy Polymers Hasan et al. [16] provide a description of a free volume theory for polymer deformation. In these theories, mobility is related to the free volume, or the volume surrounding a chain segment which is available for occupation by that segment. If a segment has low free volume, it is tightly confined by its neighbors and is unable to explore its configuration space easily. Since the regions of the polymer chain with higher mobility transform at a lower applied stress, the structural inhomogeneity described in Section 1.1 leads to the localized deformation observed in polymers. Under 14 an applied load, the regions of the chain with highest mobility undergo shear transformation events. For very small strains, the transformation strain energy is then stored in the surrounding elastic material which then exerts a back stress preventing further transformations of this region from occurring. Upon unloading following small strain deformation, the back stress returns these transformed regions to their original states leading to the experimentally observed recovery. As stress levels increase, back stress is relieved, regions of the polymer chain undergo multiple transformations, and the additional deformation is no longer recoverable. A transformation in one region increases the likelihood that another transition will follow in a neighboring region. An increasing stress causes regions with lower mobility to transform decreasing the average spacing between transformed regions and percolation of these transformed regions leads to yield. The mechanisms involved in the deformation of glassy polymers are difficult to characterize due to the many length scales over which they occur. Due to its structural complexity, there are a wide range of possible responses a polymer may have to deformation. A polymer may respond through intramolecular deformation such as bond stretching, bond flexing, and bond rotation or intermolecular deformation such as chain slippage and chain separation. Creating atomistic models which capture all of these phenomena is difficult. Atomic/metallic glasses have fewer constraints and therefore are more easily studied in the context of computer simulations. Studies conducted on these glasses have shown that plastic deformation occurs by inhomogeneous atomic movements confined to well-defined volumes undergoing relatively large shear strains [9, 38, 25]. These regions act as nucleation sites for additional transformations. As these regions link, flow structures form. The original studies on polymeric glasses by Brown and Clarke [7] and Theodorou [42] proved the feasibility of the application of computer simulation techniques to studying the mechanical properties of polymer glasses. Later molecular statics studies by Mott et al. [29] and by Hutnik et al. [20] postulated that plastic deformation is the result of shear transformations occurring within a volume element undergoing a 15 transformation shear strain. Unfortunately, the size of the activation volume element was on the order of the simulation cell used in these experiments forcing the entire cell to react at each transformation event. Hutnik, for example, calculated an activation volume which had a diameter of 10.1 nm while the simulation cell size was around 3-5 nm per side. The activation volume was calculated using the transformation strain obtained through simulation and the experimental activation volume parameter. Recent experimental evidence from Zhou et al. [47] on the diffusion of species within a plastically deforming polymer suggests that atomic mobility of the polymeric glass during flow is similar to that at the glass transition temperature. Nuclear magnetic resonance studies of deforming polymeric structures has also revealed increased mobility during plastic deformation Loo et al. [24]. 1.3 Eyring Model of Flow of Solids At low temperatures or short time scales, material behavior during deformation is governed by essentially fixed energy barriers and can be described as a thermally activated process. At high temperature and long time scales, time dependence may arise due to both the thermally activated nature of the deformation process and the occurrence of structural relaxation. The Eyring Model describes deformation as a stress-biased thermally activated process assuming isostructural deformation. The Eyring model attempts to predict the effects of strain rate and temperature on yield by assuming that flow occurs through an activated-rate process. This model does not take into account the specifics of the underlying activated process which may be the motion of atoms, molecules, or, in polymers, segments of a molecule. If this transition has an associated energy barrier of height AE, the jump rate will obey the Arrhenius equation. The application of stress to the solid results in a distortion of this energy barrier favoring jumps which occur in the direction of the applied stress. The resulting strain rate is assumed to be proportional to the jump rate. Near the high yield stress, the energy barrier is distorted to such an extent that the reverse jumps may be neglected 16 resulting in an equation relating yield stress to strain rate and temperature[26]. ( AH -rYQ)( where y is the macroscopic shear strain rate, Ay0 is the rate constant, AH is the enthalpy of activation in the limit of zero applied stress, Ty is the shear yield stress, Q is the activation volume, kB is the Boltzmann constant, and T is the temperature. Equation 1.1 can be rewritten as: rY 1AH T Q T =- T ±kBln . 70 (1.2) At constant strain rate, manipulation of Equation 1.2 gives a convenient equation for determining the activation volume from yield stress versus strain rate data. d-ry dTn The activation volume, kBT Q, kBT(1.3) , is commonly interpreted as an estimate of the vol- ume which must be activated for flow to occur, Va, multiplied by the shear strain it undergoes, YT: Q = _YTVa. (1.4) It is common in polymers for the experimentally determined activation volume to be significantly larger than the segment length. This suggests cooperative motion during the flow process for polymers. It is important to realize that this analysis assumes a single activated process. Since yield in many materials may be governed by differing processes in different temperature and strain rate regimes, extrapolation of yield stress from one regime to another may be incorrect. For example, yield data for polycarbonate below -50 degrees Celsius yields different activation parameters than data at higher temperatures [22]. This suggests that yield occurs by activation of a different mechanism in these various temperature regimes. If one imagines that there are many possible 17 molecular motions in the polymer system which can be activated each with its own activation parameters. As the temperature or strain rate vary, one or another of these mechanisms may be dominant over the others. As long as one stays within a regime where that particular mechanism is dominant, the data will fit well to the Eyring model. As the active mechanism changes or more than one mode becomes active, a more complicated model is required. In addition to various modes of deformation, the barrier heights for a particular mode may vary due to the heterogeneous nature of the disorder in the system which leads to the nonlinear transition region between linear elastic behavior and yield. 18 Chapter 2 Methodology 2.1 Molecular Dynamics Algorithm Molecular Dynamics is a technique for following the progression of a system of atoms or molecules through phase space. From a given set of initial conditions and interatomic potentials, one can generate the trajectories for particles. Properties can be calculated by averaging over the trajectory: AT = - dtA(t) (2.1) where A is the property of interest and r is the time. The key assumption in molecular dynamics is that the movement of atoms can be treated using classical mechanics. If this equation of motion holds, then the same equations of motion that govern macroscopic objects may be used to model the trajectories of atoms or molecules. In fact, the motion of all but the lightest of atoms can be treated accurately by classical mechanics even though the interactions between atoms, which is a product of the motion of electrons, must be calculated using quantum mechanics. Luckily, the mass of a nuclei is on the order of ten thousand times that of an electron. The position of the electron is highly non-localized due to its high velocities and low mass and has a short relaxation time. In contrast, the heavier nucleus has a highly localized position and requires much more time to reach its equilibrium state than the 19 electrons encircling it. The two motions can therefore be decoupled and treated independently. This is known as the Born-Oppenheimer Principle. Quantum mechanical calculations can be used to determine the energy of a given configuration of atoms with equilibrated electrons [37]. By varying distances between atoms, a map of the potential energy as a function of configuration can be developed and the motion of large systems of atoms can be calculated without explicitly taking into account the electrons. By treating the atoms as localized point masses and assuming their interactions are captured in a potential function, one can use Newton's equations of motion to solve for the trajectories of these particles. Recalling that the force on a particle can be written as the negative gradient of the potential, one can then write the equations of motion in the following form: d2 r au (rN) =i dt 2 ri where mi and ri are the mass and position of atom i respectively, N is the number of atoms in the simulation, rN = {ri, r 2 , - - - rN}, and U (rN) is the potential which can be expressed as a function of all other atomic positions. This gives a set of N 2 "d order non-linear ordinary differential equations which can only be solved numerically. The corresponding Hamiltonian for this system is given by: N 2 H = +U (rN) (2.3) i12mi where pi is the particle momentum. The Hamiltonian in this case is equivalent to the total energy of the system. Therefore, solving these equations produces a trajectory within an ensemble with a constant number of particles, volume, and energy. This ensemble is termed the microcanonical ensemble (NVE). If one would like to simulate other ensembles, changes must be made to the equations of motion. To conduct simulations in an ensemble with a constant number of particles, volume, and temperature (NVT), the Nose-Hoover thermostat was em- 20 ployed [30] [19]. In this formulation, a heat bath is coupled to the equations of motion allowing the transfer of energy to and from the system. The equations of motion become: d2r DV (rN) -r N ~ dti = =- [ N _ vi (2.4) fkBT (2.5) Q f = 3(N - 1) (2.6) where q is a frictional coefficient coupling the particles to a heat bath, Q is an arbitrary constant determining the rate at which energy is transferred between the system and heat bath, f is the number of degrees of freedom for the system, N is the number of particles in the system. There are N particles each with 3 degrees of freedom with three constraint equations forcing the linear momentum of the system to be zero which reduces the degrees of freedom to 3(N-1). The Verlet algorithm is used to integrate equations of motion for a NVT system. This algorithm can be written as At 2 r(t + At) = r(t) + Atv(t) + 2 a(t) v(t + At) ma(t + At) v(t (2.7) At = v(t) + 2t (a(t) + a(t + At)) = + At) = (2.8) -VV(rN)(r(t + At)) - 77(t + At)v(t + At) (v(t) - AtVr(t + At) + 2m 1+ 71(t +At)At 2m (2.9) 2 (2.10) At 2 At) = r(t) + At (t) + 2 i(t) 22 i(t + At) = - Zv(t + At)a(t + At) r7(t + (2.11) (2.12) The equations of motion can be similarly modified to sample an ensemble with a constant number of particles, pressure, and temperature (NPT) [32]. Instead, the 21 Berendensen algorithm is used [2]. In this formulation, the coupling between motion of the simulation cell is assumed to be weakly coupled to the motion of the particles. This allows the two motions to be integrated separately. Therefore the particle motions are integrated using the Nose-Hoover equations and the volume and positions of each particle are scaled at each time step in response to stress changes. zAt A(t) = s(t + At) = As(t) (2.14) ri(t + At) = Ari(t) (2.15) V(t + At) = (detA) V (2.16) I- 3 #3 (-r(t) - r") (2.13) Tp where A (t) is the scaling matrix, Tp is the time constant, r 0 is the desired stress tensor, r(t) is the current stress tensor, At is the time step, s is the set of vectors defining the shape of the simulation cell, ri is the vector position of particle i, and V is the volume of the simulation cell. This algorithm provides an over-damped response to pressure changes, thereby removing many of the oscillations that would be found using the more rigorous ensemble techniques. The stress tensor is calculated using the atomic virial [41]. Since periodic boundary conditions are employed, the expression for the virial stress tensor is: = M - 2V Li r' ri a F-'") - % )mn,) + (2.17) Ttail where V is the volume of the cell, pf is the a component of the momentum of the ith particle, mi is the mass of particle i, r' is the a component of the position of the ith particle, rjmin(i) is the a component of the position of the image of the particle which is closest to the ith particle, Fi is the force on i due to jmin, and jth Ttail is the correction to the pressure due to the cutoff radius. The force is defined as Fi = -Vrj-rj U(rN). The gradient is calculated keeping the separation between all 22 particles other then i and j constant. This does not presuppose pairwise additivity and is applicable to many-body forces. The calculation of long range contributions to the virial stress tensor and the acceleration can be time consuming. Often interactions beyond a given distance are discarded. When a truncation is used in calculating the contribution of a set of interactions to the pressure, the following correction must be added to the calculated stress: = 7til tal oo 23 2 (2.18) g(r) dUNBr3dr dr [00 Jrc where 6oo is the Kronecker delta which is zero if a -$0 and is equal to one if a = 3, p is the number density, g(r) is the radial distribution function, UNB is the long range potential, and r, is the cutoff radius. Assuming that the van der Waals interactions are truncated at rc and are described by a 12-6 Lennard-Jones potential and that the radial distribution function can be sufficiently approximated as 1 in the region beyond r0 , the equation for the tail correction becomes ai 167r 6 3 2 fa 2( (3 9 r, or rc 12) (2.19) This correction was employed throughout this study. In order to facilitate the computation of the van der Waals interactions, a list of the potential interacting pairs was maintained. Termed a neighbor list, this technique was first used by Verlet [43]. For each atom a list of particles within a distance d = rcutof f +6r was maintained. When any particle has moved a distance greater than L, the list is updated. Due to the constrained motion of particles in the solid phase, this list is valid for a number of time steps. The link-cell method [18] was used to aid in the creation of neighbor lists. The simulation cell is divided into rectangular compartments whose dimensions are slightly larger than the cutoff distance. Then, for a given mer, only 27 adjacent compartments must be searched to find neighbors. Since searching for these neighbors requires O(N 2 ) work, the link-cell method greatly reduces the computational time required to 23 form the neighbor list. Due to computational constraints, the particle number in atomistic simulations is severely limited. Surface effects may negatively affect even the largest of simulations. In order to reduce the surface effects, periodic boundary conditions were employed. Atoms or bonds which leave one face of the simulation cell enter the opposite side. 2.2 Monte Carlo Algorithm The Monte Carlo method is a stochastic method which generates configurations of a given system within a particular ensemble. Producing a path through configuration space, it provides no information on the trajectories or velocities of particles. The original formulation is attributed to Metropolis et al. [28]. In atomistic simulations, one is interested in calculating the properties of a given potential function within a certain ensemble. It can be shown that for a NVT ensemble, the probability of finding the system in a particular configuration, pi, is given by: -, p = exp U(r N (2.20) r) where rN is the set of positions of the particles in the system, U is the potential energy function, and kb is Boltzman's constant. The average of a property can then be expressed as an integral over degrees of freedom weighted by the probability of finding that state. <For A >= f drNeXp p et (r)) A(rUk1 N) N) where A is the property. For static properties, this average is equivalent to the time average over a single trajectory. Determining a given property of a system becomes a matter of integrating this function. Only in very rare cases can this integral be evaluated analytically. Re- sorting to numerical methods, one could randomly sample configurations, calculate 24 the weight and average over the weighted samples. This is inefficient since most of the configurations sampled would be high energy states with low weights. A more efficient way to calculate this ensemble average would be to generate configurations according to the weight e-U(rN). Using this scheme, one samples more frequently those states that contribute most to the average. The property can then be expressed as a simple average of the computed property over the generated states. I <A >= - N, N Z A(r (2.22) where A is the property of interest, and N, is the number of generated configurations. To sample configurations according to this given weight, Metropolis developed the following algorithm: 1. given an initial configuration rN 2. calculate the energy of this state Uj = U(rN) 3. choose an atom at random and perturb it 4. calculate the energy of the perturbed configuration Uj = U(r ) 5. if (Uj - Uj) < 0 accept the perturbation otherwise accept configuration with a probability e- U---) 6. calculate property A 7. goto 2. The average of property A can then be evaluated using Equation 2.22. 2.3 Generation of Initial Structures The procedure for producing a well equilibrated structure is similar to that used in the study by Chui et al. [8]. A system of mono-dispersed chains of specified lengths is generated according to a random walk algorithm with fixed bond lengths and angles. The dihedral angle is randomly assigned as the walk progresses. This structure is relaxed in the NVT en- 25 semble at a temperature of 1300-2300 K using the Monte Carlo Algorithm described in Section 2.2 with the Lennard-Jones potential turned off. During this stage the chains are free to pass through each other and adopt an end-to-end distance characteristic of an idealized chain. The LJ potential is then turned on and the LJ radius is slowly increased to the desired value. This equilibrated sample of free chains is then crosslinked according to a procedure similar to that used by Duering et al. [10]. During crosslinking, functional groups are attached to all chain ends. When these groups come within a specified critical distance from one another, one functional unit is removed and all atoms bonded to this group become bonded to the remaining unit. The system configuration is evolved using the Monte Carlo scheme and periodic checks are made on the distance between functional groups. To quicken the crosslinking process, an artificial potential is added between crosslinks. ECL 1 2 -KCL$ |rijI {1 :r7 + 771 > rimax : (2.23) (2.24) Th + 77'<; 7max where KCL the potential constant, Iri I is the distance between functional unit i and j, rTi is the current functionality of group i, and rimax is the maximum functionality. The sum is carried out over all pairs of functional groups. The added potential acts to draw groups that can be joined together while repulsing those whose added functionality exceeds the maximum. The functionality, potential constant, and critical distance can be varied to change the crosslinking dynamics. The crosslinked structure is then taken as the initial configuration for the molecular dynamics algorithm described in Section 2.1. The Verlet integration algorithm requires that the position, velocity, and acceleration of each particle be specified as initial conditions. The Monte Carlo algorithm gives only configurational data. The velocities and accelerations must be specified at the start. The initial accelerations 26 are computed from the gradient of the potential and the initial velocities are randomly assigned according to a normal distribution. The probability density, p (vix), for the normal distribution of velocities is given by the following equation: m p (vix) = 2rkBT exp '-mv* kB I (2.25) where kB is the Boltzman Constant, T is the absolute temperature, m is the mass, vix is the x-component of the velocity for the ith particle. Velocities are assigned according to the normal distribution at the elevated temperature. The sample is cooled at a constant rate and equilibrated at the desired temperature. The pressure control algorithm is used to equilibrate the system at atmospheric pressure. 2.4 A Lennard-Jones Fluid In order to test the implementation of the algorithms described in Section 2.1 and 2.2, the equation of state for a Lennard-Jones fluid was determined and the results compared with previously reported data [11]. Data are presented in terms of a reduced temperature, pressure, and density: kBT (2.26) Pa3 (2.27) pr (2.28) 0 P = p* = (2.29) where kB is the Boltzman constant, T is the absolute temperature, E and a are the Lennard Jones parameters, P* is the reduced pressure, p* is the reduced density, 6 is the reduced temperature. This is done so that the results are independent of the choice of a and c. The cases of 0 = 0.9 and 0 = 2.0 are shown in Figure 2-1 and Figure 2-2 respec- 27 0.8 o Monte Carlo .1- Molecular Dynamics -- Data from Frenkel and Smith (1 996)_ 0.6- 0.4- 0.2- -0.2- -0.4 -0.6 0 0.1 0.2 0.3 065 0.6 0!4 Reduced Density p = No /V 0,7 0.8 0.9 Figure 2-1: Pressure-density response of a Lennard-Jones fluid at 6 = 0.9, N = 500 tively. A 50 ps simulation was run at each pressure and the values for temperature were averaged over the last 30 ps. The simulations were conducted with the cutoff radius set to one half of the simulation cell length. Excellent agreement with the published data was found. 2.5 Polyethylene Chain Model The quality and accuracy of atomistic simulations depend upon the quality of the force field used to model a system. The structure and dynamical properties of a simulated system are dictated by this force field. Therefore it is of utmost importance to find a force field that accurately produces the properties of the system which are to be investigated. The amount of detail that must be captured by the model also depends on the phenomena to be studied. Polyethylene was chosen as a model system due to its chemical simplicity: It consists of a backbone of carbon atoms with pendant hydrogen atoms. It has been shown in numerous studies that the explicit representation of these pendant hydrogen atoms in atomistic simulations is not necessary to recreate many static and dynamic 28 1 10 11 1 11 1 Monte Carlo <1 Molecular Dynamics -0- Data from Frenkel and Smith (1996) 10 9 8 - 7 -I 6- / 5 4- 3- 2- 0 0.1 0.2 03 0.4 0.5 0.6 3 Reduced Density p = Na N 0.7 0.8 0.9 1 Figure 2-2: Pressure-density response of a Lennard-Jones fluid at 0 = 2.0, N 500 properties [34]. In the united atom approximation, each methyl group is collapsed into a single effective potential positioned at the carbon site while leaving the essential skeletal structure of the carbon backbone unaltered. This approximation has been widely used due to the three fold reduction in the number of explicit particles in a given simulation which reduces the computational time required by an order of magnitude. For the purposes of this study, the force field optimized by Paul et al. [34] was used in a modified form. This model consists of an united atom representation of polyethylene optimized to reproduce experimental data on the properties of melts of n-alkane chains with varying molecular weights. This appears to capture the local and large scale dynamics of the chain well. The van der Waals forces between united atoms are modelled using the LennardJones functional form of the potential. EU = 4E (" 12 , 6) rj<ro (2.30) - 0 :evl dr > reo where o- is the excluded volume radius, E is the well depth of the potential, rij is the 29 vector distance separating the united atoms i and j, and reu is the cutoff radius for the Lennard-Jones potential. The rigid bonds of the Paul model are replaced with a harmonic potential: E = Ik (Iri| - bequi) (2.31) where kb is the bond potential coefficient, bequil is the equilibrium length of the bond, and rij is the vector distance separating the bonded united atoms i and j. The angle potential is given by: ko (cos (jk ) Eik = cos (Oik) - = cos (Oequii)) 2 -ri rge(233 ri3 rjk (2.32) (2.33) where ko is the angle potential coefficient, Oequil is the equilibrium angle of the potential, and 0 ijk is the bond angle between consecutive united atoms ij, and k. The functional form of the torsional energy is given by: 3 km E (cos' (Oihkl)) m=0 (2.34) rij x rjkJ (2.35) \rij rik )| Eijki = COS (qijk) where k, k2, k3, k Paul, #ijkl - 1 x rjk x rkl |rjk x rk1I|) are constants which were fit to the torsional potential used by is the dihedral angle formed between the consecutive united atoms i, j, k, and 1, and #ijki = ir radians corresponds to the trans torsional state. The calculation of the van der Waals potential is the most time consuming step in the simulation of this system. Due to the rapid decay of the Lennard-Jones potential, a cutoff was applied without greatly affecting the evolution of the system. Corrections to the calculated pressure for the truncated system were added to compensate for the interactions that were not included directly. The proper corrections are discussed in 30 ijkl 1 k . - i j k Figure 2-3: Representation of dihedral angle rotation, bond angle bending and bond stretching. section 2.1. The van der Waals interactions were not calculated for atoms which were separated by fewer then 4 bonds along a chain since the corresponding contribution was accounted for by the bond, angle, and torsional potentials for these groupings. 2.6 Correlation of Torsion Angles In order to observe the relaxation or change in the torsional angle over time, the torsional angle autocorrelation was calculated according to the following formula: PO(t) W (< cos((0))cos(#(t)) > - < cos#(O) >2) (< cos 2 (2.36) (O) > - < cos#(O) >2) where 0(t) is the value of the dihedral angle at time t, 0(0) is the initial value of the dihedral angle, and the brackets denote an ensemble average. The correlation between transitions of neighboring dihedral angles on a given chain was also computed. A transition was defined as the movement of a dihedral 31 20 - - - -- - -- - - - - .- - -- - - - 18 -~. --... --.... -... -.-.--.-----.----..----.-. 16 ..-.- . .. ... .. ... .....-. ..-. ... ..- -. . ....-. . .. .-. -..-.14 0 E 12 .... .- ---. .. ... --- . -. - --. . -...-.. -. -. .. -.. ..- .-. - 10 ..... -. -. - .....- -. .-. . .. . . . . -.-. . ... . .... .-.-.-.-.-.-.-.-.. - 8 -.--. .. --. .. --.-. -.-. -. - ... .. .. . .. .. -.-.-. . .-. -. 6 4 2 2 1 4 3 Dihedral Angle (Radians) 5 Figure 2-4: Energy as a function of dihedral angle Table 2.1: Potential Parameters 4.o1 A 0.468 rc 2.5 2000 kJ/kmol/A 2 kb bequji 1.53 A ko 5.01 kJ/kmol 0 equil 1.92 rad 14.47 kJ/kmol ki k2 -37.59 kJ/kmol k 3 6.490 kJ/kmol 74 58.49 kJ/kmol o 32 6 angle from within 10 degrees of a potential minima to within 10 degrees of another potential minima. By maintaining the time and position of each transition on a given chain, the probability that a transition will be followed by another transition a given number of segments away from it was calculated. No restrictions were placed on the time between transitions. Neighboring transitions occuring within the same time step (each time step was one femptosecond long) were not counted. 33 Chapter 3 Structure The behavior of the model polymeric system was characterized as a function of temperature. The glass transition temperature was determined as a function of chain topology from the specific volume versus temperature response. Since rotation of the dihedral angle is strongly related to the relaxation of the polymeric chain, energy barrier transitions between dihedral angle state transitioning were calculated. The correlation of local density and transition rates and the correlation between neighboring dihedral angles were investigated. These will serve for comparison to the system response during deformation. 3.1 Initial Network Characterization There are two issues associated with the short time and length scales employed in these simulations which must be addressed. Firstly, only a limited amount of phase space is explored during the course of these simulations leading to a large variation in the behavior of systems with different initial configurations. Conducting the same deformation experiment with various initial configurations, one can obtain an estimate of the variation in behavior to be expected. Secondly, relaxation in these samples occurs over many time scales. Stretched bonds, for example, equilibrate over femptosecond time scales while chain motion requires much more time. The glassy state inherently is a non-equilibrium state, but one would like to be sure that structural 34 relaxation will not occur on the time scale of the simulation. There are a number of indicators which were used to monitor the relaxation of the system. By following measures of density, energy and stress as well as distributions of bond lengths, angles, and dihedral angles, one can observe the system's progression towards equilibrium. The systems are crosslinked and cooled under NVT conditions at densities which are slightly lower than the equilibrium densities. The sample is then relaxed under NPT conditions at atmospheric pressure and the convergence of the density and energy measures are used to gauge the equilibration of the system. Figure 3-1 is a characteristic relaxation profile. The density fluctuates rapidly for the initial 50 ps as the system is switched from NVT to NPT. After this period, the density fluctuates about the equilibrium value. This particular relaxation was conducted at 300K which, as will be shown later, is above the glass transition temperature for this particular system. The energy fluctuations and stress in the system both fluctuate about zero as shown in Figure 3-1. A compilation of statistics for equilibrated structures which were used as initial conditions for the ensuing deformation studies is presented in Table 3.1. In addition to providing densities, chain statistics, and measures of conformation, the functionality of crosslinks is reported. Histograms of the bond length, bond angle, and dihedral angle distributions which are representative of equilibrated structures are shown in Figures 3-2 to 3-4. The harmonic bond length potential employed in these simulations results in a normal distribution of bond lengths. On average, the bond angles are slightly flexed closed with respect to the potential equilibrium value. Mean bond lengths for the distributions shown in Figure 3-2 at 100 K, 200 K, and 450 K were 1.5276, 1.5285, and 1.531 angstroms respectively while the mean bond angles were 109.779, 109.784, and 110.07 degrees respectively. The equilibrium bond length is 1.53 angstroms while the equilibrium bond angle was 110 degrees. With an increase in temperature the compressed bonds are relieved and then pulled in tension while the compressed angles are relieved then flexed beyond their equilibrium values. 35 0.92 2 - - 0.91 1 o9 5. -van Bond Stretching Bond Angle Bending Dihedral Angle Rotation Der Waals A w 1., V E A >0.89 V 0 w 0.5 0.88 0 0.87 50 100 200 150 Time (ps) 50 250 150 Time (ps) 100 200 250 (b) (a) 200- -150* x-direction y-direction z-direction 100 CL 50 -50 50 100 150 Time (ps) 200 250 (c) Figure 3-1: Measures of equilibration for sample system with 4500 mers at lOOK. The dotted line indicates the time at which the ensemble was switched from NVT to NPT: (a) density as a function of time, the dashed line indicates the equilibrium density (b) change in energy as a function of time (c) stress as a function of time. 36 - 0.06- x 100 K 200 K 300K 0.05- .20.04- 0.03-x XXX cc x x x x XXx x xx . x X 0.02- X x X X x. XX 0.01- XX >x Xxx x ' xX XXX . '. x XX .. 0.9 0.92 X xxxxxx --. 1 1.02 1.04 0.96 0.98 Normalized Bond Length (1/b ) 0.94 1.06 1.08 1.1 Figure 3-2: Distribution of bond lengths for crosslinked systems, nw-1-8000-160-100K, nw-1-8000-160-200K, nw-1-8000-160-450K 0.06 - 0.06x OO 100 K 200 K 300K 0.05- 0.04- (D 0.03 -X X x 0.02- X XX Y XXx .' 0.01 . X . x "x xx - - XXX xX 1.7 xx x x. x . 1.8 1.9 2 Bond Angle (Radians) xx 2.1 2.2 Figure 3-3: Distribution of bond angles for crosslinked system. 37 -- 0.09- x 100 K 200 K 300K 0.080.07.20.06- xx 16 x CO IL ,0.05- C 0.04 0.03 0.02- x x- 0.01 - x x x 2 1 0 4 3 Dihedral Angle (Radians) xx x 6 5 Figure 3-4: Distribution of dihedral angles for crosslinked system. To determine whether the initial structure had any initial bias in orientation, vector correlation functions were computed. Vectors were constructed connecting atoms separated by two bonds along a chain. The average angle between the unit vector along this direction, vi, and the load direction, es, was calculated and the value is reported in table 3.1. Pi =< Vj - ex > (3.1) Assuming that there are an infinite number of vectors, vi, oriented randomly in space, then the probability of finding a vector at particular angles # and e is uniformly distributed. The value of P can then be computed by P1 = dE f2 dqarccos(cosecos#)cose ~- 57.3' (3.2) The radial distribution function was calculated to obtain an estimate of the long range order within the sample. Thin shells of thickness 6r are constructed each centered r away from a particle. The number of particles, N(r), within each of these 38 AF & ij X ;41 Xx Ao- -W (b) 200K (a) 100K (c) 300K Figure 3-5: Pole Figures for crosslinked system at temperatures of 100K, 200K, and 450K. shells is computed. The radial distribution function is then defined as: N(r) _V g(r) - -Nwr) N 4,Fr26r (3.3) where r is the distance to the shell, 6r is the thickness of the shell, V is the volume of the cell, N is the total number of particles within the cell. In a perfect crystal, this function will have sharp peaks centered at the spacing of neighboring particles. On the other hand a liquid will have smoothed peaks at short distances, where molecular spacing is influenced by the Van der Waals interactions, and decay to a constant value of 1 at larger distances. This decay is indicative of the lack of ordering at these larger length scales. An amorphous polymer with its short range order determined by atomic bonding and lack of long range order will consist of a mixture of these two profiles. Figure 3-6 provides the radial distribution functions for initial structures at various temperatures. At 450K the radial distribution function reveals a polymer melt with peaks representing the first, second, and third neighbors along a chain. At distances beyond this, the distribution is liquid like. Decreasing the temperature below the glass transition temperature results in the appearance of fourth neighbor correlations as the dihedral distribution becomes localized at the 39 5- - 200 K x 4.5- 450 K 4- 3.53- 'a; 2.5 2- 1.5 - 1- 0.5- 1 2 4 3 Normalized Distance (d/ulb 5 6 7 Figure 3-6: Radial distribution function for initial structures at various temperatures. minima as shown in figure 3-4. The long range order, however, is frozen in as the system kinetics slow and it is unable to achieve a crystalline structure. To obtain an estimate of the length scale of the inhomogeneity captured in the structure, the simulation cell was divided into subsections. The density within each of these sections was calculated and the distribution of these densities was examined. Each atom was approximated by a cube with uniformly distributed mass with a side length given by L. L = V7(0.8u-) (3.4) If an atom was cut by the boundary of a subsection, the mass of each atom was then divided accordingly between the neighboring subsections. A representative histogram of the density distribution is shown in Figure 3-7. As the size of the bin increases, the distribution of densities becomes narrow indicating the uniformity of the sample at this scale. As the section length then approaches the Lennard Jones parameter -= 4.01 Angstroms, the distribution broadens indicating the heterogeneity at this level. 40 Sample ID I. nw-c-1-4500-90-100K nw-c-1-4500-90-200K nw-1-8000-160-100K nw-1-8000-160-200K nw-1-8000-160-450K nw-c-1-10000-200-200K nw-t-1-4000-80-100K nw-t-1-4000-80-200K nw-t-1-4000-80-450K fc-1-4000-83-100K fc-1-4000-80-300K fc-1-4000-40-100K fc-1-40000-160-100K N 4380 4380 7775 7775 7775 9722 3946 3946 3946 3984 4000 3920 40000 T (K) 100 48/90 200 48/90 48/160 100 48/160 100 48/160 450 48/200 200 100 98/40 200 98/40 450 98/40 100 48/83 300 50/80 100 98/40 250/160 100 M/C p (g/cm3 ) 0.942 0.923 0.939 0.916 0.848 0.921 0.935 0.921 0.854 0.901 0.855 0.914 0.916 < dee > (A) 20.4 20.58 17.81 18.02 18.83 19.09 23.04 23.26 24.48 17.81 16.76 23.72 30.01 P (Degrees) 54.86 54.706 58.04 57.81 58.21 58.04 54.67 55.63 56.47 57.35 59.88 56.16 57.07 Trans % 66.9 66.8 63.7 63.9 65.7 65.7 66.8 67.0 64.7 62.2 62.4 63.5 64.3 F2 F3 F 4 16 16 12 12 12 20 6 6 6 0 0 0 0 19 19 36 36 36 33 12 12 12 0 0 0 0 22 22 47 47 47 64 8 8 8 0 0 0 0 Dimensions (A) x 29.43 29.01 x 126.6 127.5 x 29.42 x 29.42 57.75 x 57.75 x 57.75 58.22 x 58.22 x 58.22 59.75 x 59.75 x 59.75 176.6 x 37.28 x 37.28 23.51 x 64.20 x 64.98 23.53 x 64.65 x 65.44 23.48 x 69.80 x 65.57 47.06 x 47.06 x 47.06 24.00 x 68.46 x 66.17 46.48 x 46.48 x 46.48 100.5 x 100.5 x 100.5 Table 3.1: General Structure of Chains where N is the number of mers in the simulation, M is the number of mers per chain, C is the number of chains, < dee > is the average end to end distance, Pi is given by equation 3.1, % Trans is the percentage of dihedral angles in the trans configuration, F 2 , F3 and F4 are the number of crosslinks with a functionality of 2, 3, and 4 respectively 0.45 - 6859/5.29/5.8 std= 0.09 5832/5.59/6.86 std = 0.115 - - 4096/6.28/9.8 std = 0.133 - 729/11.2/54.9 std = 0.0815 x 216/16.7/185 std = 0.0338 x 0 0.4- 1 - 0.35 - x 0.3- 0.25LA. 0.20 0.15- 0.1 - 0.05 00 0.4 0.5 0.6 0.7 0.9 0.8 3 Density (g/cm ) 1 1.1 1.2 1.3 Figure 3-7: Distribution of densities at 100K for 40000 mers free chains. Number of Bins/ Length of Side (Angstroms) / Average of atoms per bin and standard deviation of distribution The effect of crosslinking on the initial structure network was also investigated. The density of the crosslinked systems was higher than similar systems of free chains. Examination of the bond length distribution, bond angle distribution, dihedral angle distribution, and radial distribution functions reveal no obvious differences between the crosslinked and free chain systems. These distributions for two network systems with 48 and 98 mers per chain and a system of free chains with 250 mers per chain are shown in Figure 3-8. Referring to Table 3.2, one observes that the change in average bond length and bond angle is small between the free chain and network systems but that the density changes are significant. The density change is partly due to the increased density around crosslinks. A crosslink can have up to four mers within one bond length away as compared to two atoms within one bond length for a normal atom and one mer for a chain end. The free chain system which has less dense chain ends instead of the high density crosslinks has a smaller density. 42 0.0 5 x 0 0. - -- MM -= 4898 (network) xo MM ==4898 (network) (free chains) 0.04 - M : 48 (network) M- 98 (network) M-48 (Iree chains) M 8 (free hlns) 0,- 0.05 0.045 0.0 0 0" (free chains), 0.035 0* 0.0 1- 00 0.03 -0 0 0.025 2- 0.02 0.0 0 0oxo 2- 0 x 0.01 0.0 0.96 x 0.005 000 0.95 8 x 0.015 0.87 0.88 0.99 1 1.01 1.02 Normalized Bond Length (554).. 1.03 1.04 1.05 1.70 1.8 1.85 (a) 2 2.05 (b) - x 0.12- 1.9 1.90 Bond Angle (Radians) o M M M M - 48 - 98 =48 98 N49 (network) (network) (free chains) (free chains) 4 .5 xFC 260, - 0.13 - 0.08 !.S 2- 0. 0.04- 0.02- x 1 1.5 2 2.5 3 3.5 4 Dihedral Angle (Radians) 4.5 5 1 5.5 2 3 Normalized 4 Distance (d/b 5 6 7 (d) (c) Figure 3-8: Comparison of bond and angle statistics for various systems: (a) bond length distribution, (b) bond angle distribution (c) dihedral angle distribution (d) radial distribution profile. 43 Table 3.2: Crosslinked vs Free Chain Systems. The equilibrium bond length and angle are 1.53 A and 110.01 degrees respectively. Density (g/cm 3 ) Sample 0.942 0.935 0.901 Average Bond Length (angstroms) 1.5273 1.5272 1.5266 Average Bond Angle (degrees) 109.77 109.78 109.80 free chains, M = 98 0.914 1.5271 109.76 free chains, M = 250 0.916 1.5270 109.80 network, M = 48 network, M = 98 free chains, M = 48 3.2 The Glass Transition Temperature The glass transition temperature was determined for various network topologies. Simulations were conducted by varying temperature while maintaining the cell boundaries at atmospheric pressure. Each temperature was maintained for 150 picoseconds and the average density obtained from an average over the last 50 picoseconds. At lower temperatures, the density would continue to increase if allowed more time to relax. The times used simulate a cooling rate of 1.25 x 10- 10 K/sec. Below the glass transition temperature, one cannot speak of equilibrated density. If the simulations were performed for longer times, one would observe higher densities at these low temperatures. The glass transition is approximated from the point at which the slope changes in the graph of the specific volume versus temperature. In Figure 3-9(a), results for varying chain lengths and crosslinking densities are presented. The glass transition temperature as well as the linear thermal expansion coefficient measured before and after the glass transition temperature are reported in Table 3.3. The glass transition temperature seems to respond sensitively to changes in the chain molecular weight and to the addition of crosslinks. An increasing molecular weight or crosslink density is found to increase the glass transition temperature. This result is in agreement with previous work done by Rigby and Roe on a polyethylene like molecule and suggests that beyond a certain chain length, the glass transition temperature will be constant with further increase in chain length [35]. Extrapolation 44 of the data from Rigby and Roe suggests that for their model, chains with more than approximately 200 units have a common glass transition temperature. These phenomena are also observed experimentally [45]. The chain ends being more mobile and having higher free volume than mers along the backbone will lower the glass transition temperature. For low molecular weight chains, chain ends can increase the free volume significantly but at higher molecular weights, the additional free volume from chain ends is an insignificant fraction of the total and the glass transition is constant with molecular weight. Crosslinks are less mobile and have less free volume than a mer along the backbone. Therefore the addition of crosslinks stifles chain mobility and acts to increase the glass transition temperature. It has been shown both in experiment and through simulation that the glass transition is dependent on cooling rates and pressure[17]. Experimentally the temperature may be varied on the order of degrees per second or degrees per hour while this molecular simulation experiment was limited to cooling rates on the order 101 K per second. The simulated thermal expansion coefficient was found to be between 175 and 194 x10- 6/C. The experimentally observed glass transition temperature for polyethylene is approximately 250 K while the thermal expansion coefficient is reported to be between 110-200 xlO-6 /K [6]. Noting that the much faster cooling rate is expected to elevate the glass transition significantly and that the experimental values are for semicrystalline high density polyethylene, the simulated results compare favorably with experimental results. The glass transition temperature of polyethylene determined through atomistic studies varies between 200 and 300K depending on both the cooling rate and the specific potentials employed [46, 13, 40]. Yang et al. [46] find a glass transition temperature of 306K for a system of 5 chains with 300 united atom beads each at a pressure of 0 MPa. Takeuchi et al. [40] simulate a chain of 600 mers with a cooling rate of 1.66 x 10" K per second and find a glass transition at 201K . Gee et al. find that by varying the torsional potential from that of a freely rotating chain to barriers twice those employed in this study and a chain length of 768 united atoms cooled at rates on the order of 10" K per second, the glass transition varies between 100K and 400K [13]. 45 Table 3.3: Glass transition temperature, Tg, and linear thermal expansion coefficient, LTEC, for various molecular weights. Topology 50 mers per chain 100 mers per chain 1000 mers per chain Network Tg (K) 252 268 280 305 LTEC x10 6 /CO T = Tg -6T 71 66 63 60 LTEC x10 6 /C T = T +6T 194 188 175 175 One universal feature observed in all of these systems is that the fraction of dihedral angles in their lowest energy state, i.e. trans state, does not obtain its equilibrium value. Instead as the temperature falls below the glass transition temperature, the chains which are no longer able to fully explore phase space become "frozen" into their current states. The resulting plateau of the fraction of dihedral angles in the trans configuration is shown in Figure 3-9(b). As shown in the next section, the transition rate decreases with temperature. A longer amount of time is therefore required to allow all of the dihedral angles to explore various states and obtain their equilibrium distribution. 3.3 Energy Barriers for Transitioning As mentioned earlier, the torsion potential has three potential minima. Gauche plus, trans, and gauche minus, refer to torsion angles inhabiting these minima as shown in Figure 3-10 and Figure 3-11. The torsion angle is significantly more compliant than bond stretch and bond bending and therefore is the degree of freedom utilized for chain motion. Understanding this motion and the barriers to it is critical to understanding polymer response. Though a specific torsional potential has been specified for the rotation of dihedral angles, this potential will be modified by intermolecular interactions. There are two methods for determining the modified or "effective" barrier potential for the dihedral angles. The first method results in a full potential and involves the 46 1.3 - o x 50 mers 100 mers o 1000 mers 1.25 > Network 1.2> E 0 1.15 0 1.1- a) CO, 1.05 1' 0 100 300 200 400 500 Temperature [K] (a) Specific volume versus temperature. 0.72- 50 mers x 0.7 - 100mers 1000 mers Network 0.680.66 0.640.620.6 0 100 300 200 Temperature [K] 400 500 (b) Fraction of dihedral angles in the trans configuration as a function of temperature. Figure 3-9: Simulations of polyethylene model at atmospheric pressure with 4000 united atom beads. 47 calculation of the probability that a dihedral angle will assume a given configuration. The effective potential can then be calculated from this potential. The second method provides information on the barrier height between the minima. Transition rates can be calculated for the various states. The assumption that the rate of transition between state 1 and state 2, q 12 , will be related to the activation energy, AE 12 , by the Arrhenius relation leads to the following equation: 412 = K 12 exp (kBT (3.5) where T is the temperature and K 12 is the rate constant. The definition of a transition is ambiguous. For the purposes of this study, gauche plus, trans, and gauche minus states were assigned to dihedral angles passing within ±10 degrees from a potential minima which they maintain until they pass within this prescribed distance of another minima. Transitions are counted when the state assigned to a dihedral angles changes. Bearing in mind that the activation energy between a change in angle from one configuration to another depends on both the initial dihedral angle and the final dihedral angle due to the curvature of the potential wells, calculating activation energies between transitions thus defined gives only an average quantity. The transitions between the various states were observed for a system of 20 free chains of 250 mers at various temperatures as shown in Figure 3-12. The temperature was decreased in 25 degree increments. The sample was equilibrated for 10 ps at each temperature and was followed by a 200 ps sampling run. Transition rates were calculated as the total number of transitions divided by the total sampling time and by the total number of dihedral angles. The gauche plus and gauche minus states are indistinguishable. Therefore transitions between gauche plus and gauche minus are equivalent to transitions from gauche minus to gauche plus. The calculated activation energy for transitions reported in Table 3.4 are very close to the barrier energies in the explicit torsional potential. This suggests that the torsional potential is not altered significantly by the neighboring molecules. This is supported by similar results by Boyd et al [5]. 48 Table 3.4: Calculated Activation Energy for Dihedral Angle Switches Transition Type Gauche to Trans Gauche to Gauche Trans to Gauche AE (kJ/mol) 12.3 15.6 12.4 Also, referring to Figure 3-12, one notes that as the temperature decreases, the transition rate deviates from a strictly Arrhenius dependence. In fact, the transition rate is larger than that predicted by such a dependence. At 100K, the predicted log of the transition rate per bond is -2.5 while the actual value is -1. observed in a previous study by [5]. This has been They claim that the transition rates remain higher than predicted by the Arrhenius model because of excess free volume trapped in the system below the glass transition temperature. Analysis in Section 3.4 shows this is not the case. The decay of the torsional autocorrelation function is closely tied to transition rates. For the torsional autocorrelation function to approach zero, all angles must sample other states. The decay function was calculated according to Equation 2.36. Results at various temperatures are shown in Figure 3-13. At 1OOK one notes that there is no appreciable drop in the decay function even after a few hundred picoseconds. This indicates that the dihedral angles retain their initial state over these long time periods. This correlates with the low transition rates observed at these temperatures. The rapid decay of the torsional autocorrelation function at 450K indicates that structural relaxation is occurring rapidly while at lower temperatures structural relaxation occurs on much longer time scales. Agreement between the decay of dihedral angles at 450K and previously published data from Pant et al. [31] was found. Though Pant et al. incorporated an anisotropic united atom model verses the simple united atom model employed in this study, the torsional potential was similar to that employed in this study. The striking similarity in decay time suggests that the dihedral angle decay is relatively insensitive to this difference in force field at high temperatures. Since it was shown earlier that the 49 I ' - ~ %I - - T7 Figure 3-10: Diagram of dihedral angle rotation. torsional potential is not modified significantly by the surrounding molecules, it is expected that the decay times should be similar for similar torsional potentials at high temperatures. It is also important to note that transitions are not distributed evenly amongst the dihedral angles. There are some angles which undergo numerous transitions while others undergo none. At a temperature of 350K, 40% of the dihedral angles are involved in the first 4043 transitions, 3-14(a). The remaining 60% undergo no transitions during this period. The standard deviation about the mean number of transitions per dihedral angle is 1.65. At a temperature of 250K 23.2% of the dihedral angles are involved in the first 4043 transitions 3-14(b). In this case the standard deviation about the mean number of transitions per dihedral angle is 2.57. This heterogeneity becomes even more exaggerated as the temperature decreases further. At 1OOK 4.4% of the angles participate in the first 1941 transitions while at 225K 12.3% participate in as many transitions. 50 22 t g 20 g+ 18 16 :14 0 E 12 210 W 8 6 4 K 2 I 100 50 300 250 200 150 Dihedral Angle (Degrees) 350 Figure 3-11: Diagram of transition definition. . -.. .-.-.-. - -----.-. .-.--.-. + g--t - t ->g2.5 - - 0 - t c' 2- 1.5 S9-->9+ ->g+ -- Total Rate Total Rate (Gee --- - t g+ a.) 1- -0.5 ~~~~~~~~~~. -1 ............ . ....... . . . ..... . . .. . .. . 04-0.4 -- - -- - 1- .5 - 2 2.5 3 - 3.5 4 4.5 5 1000/(Temperature (K)) Figure 3-12: Transition rates for a system of 5000 mers arranged in 20 free chains. 51 0.9 2O.8- =0.5 C3 I0.4 - 0.4 - 0.2 _ 0.1 -- -- 0 10-1 b network, T = 100K network, T = 200K - -network, T = 450K 0 Pant et al. r. .. I 100 . 101 102 103 Time (ps) Figure 3-13: Decay of the torsional angle autocorrelation function at various temperatures. 3.4 Transitions and Free Volume An attempt was made to find whether the dihedral angles with the highest transition rates had higher larger local free volume. Two methods were employed in this attempt. In the first method, the system was split into bins. The density in each bin was calculated according to the method outlined in Section 3.1. The transitions were divided into the same bins weighted by the volume of the mer falling into each of the bins. The density and transition rate were recorded for each bin. When averaged over multiple samples, the distribution of densities for bins with highest transition rates was identical to the distribution of densities for bins with the lowest transition rates. In the same manner, the distribution of transition rates for the bins with the lowest density was found to be identical to the distribution for bins with the highest densities. Plots of transition rate versus density are shown in Figure 3-15 and Figure 3-16 for a bin size of 6.31A x 6.31A x 6.31A and 5.23A x 5.23A x 5.23A respectively. One would expect that the bins with lower density would exhibit more activity but these figures show no correlation between density and transition rate over the length 52 22 20 18F 16 14 E12 10 10 8 6 -11 Il 4 I I . il 11,|1 111. ill JIJI II 2 0 500 1000 1500 2000 2500 3000 Dihedral Angle Number (a) T 3500 4000 4500 5000 3500 4000 4500 5000 350K 40353025 CD - S20-- E m 15 z Ii 10 5 500 1000 Dihedral 15000 g b DierlAngle Number (b) T = 250K Figure 3-14: Spatial distribution of conformational transitions for a system of 5000 united atoms arranged in 20 free PE chains. Dihedral angles are numbered serially along the chain and the number of transitions at each dihedral angle are plotted against the dihedral angle. The total number of transitions at each temperature is 4043. 53 scales investigated. The second method involved correlating transition rates with the tessellated volume for the atoms along the chain making up the dihedral angle which was consistently transitioning. The tessellated volume gives an indication of the local volume surrounding a single mer as opposed to the first method which averages the density over regions on the order of 5-10 atoms. One would expect that at least one of the mers comparing the active dihedral angle has a larger than average tessellated volume which then gives it more mobility allowing transitions to occur. This was not found to be the case. In fact, the distribution of tesselated volumes was identical to those obtained by random sampling. Though these methods should be able to resolve the correlation between excess free volume and high transition rates, they were unsuccessful when applied to this model. As shown earlier the energy barrier for the dihedral angle transition is on the order of the torsional barrier. This suggests that the transition is unaffected by the surrounding mers at these temperatures. One notes that the united atom model employed in this study reduces many of the interactions between neighboring chains by condensing the methyl groups. At higher temperatures, one expects very little interaction between chains due to the low density and increased spacing between neighboring chains. If the transition rates at these higher temperatures are extrapolated to temperatures below the glass transition temperature, the predicted rate is lower than that observed. Though it had been postulated that this was caused by excess free volume trapped in the system due to cooling, the most active regions appear to be those which are most constrained, crosslinks in the case of the network and interior mers in the case of free chains. The methods employed require the positions of each atom at the time of a given transition. Due to memory constraints, positions may be maintained only at selected times. These may not correspond to the exact time of a transition. It may be possible that a given dihedral angle may have additional free volume prior to the transition but that the free volume collapses by the time the next position file is obtained. For this reason, transitions were correlated with position files saved within 1 pico-second 54 14 35 a12- 30x 10- 25 - x -x2 G20 - x ax 8x - x .x x x 6 x 4- 10 -% xxd- x 5- x x x - 2 "f x 0 0 x 0 0.2 0.4 0.8 0.6 3 Density (g/cm ) 0.2 1.4 1.2 1 0.4 0.6 x 0 0.8 1 1.2 1.4 Density (g/cm) (b) T = 300K (a) T = 400K 8 x x 0x 7- r.6 6- x Ii Cr 3 - 4 - x K xx x 4- - x x x x x x 2 x 2 . 1- x x, x x. x 0 0.5D - xx 1I 1.5 Density (g/cm3) (c) T = 250K Figure 3-15: Correlation between transition rates and density at various temperatures for a system of 5000 mers arranged in 20 free chains using a bin size of 6.31A x 6.31A x 6.31A. Each bin contains an average of 10.6 mers. 55 60 100 90 Zx -- .i 50 - 80 70- 840- 1' x 60 - x 40 x xx 1xxx 20 x 30- 10 20 0.2 0.4 x .** xx1 x x x 0 %20- - 10 xt"" x x 0.8 0.6 Density (g/cm) 1 XX 0.2 0 1.4 1.2 0.4 0.6 0.8 3 1 1.2 1.4 Density (g/cm ) (b) T = 300K (a) T = 400K 20 18- 16 x 14 12 10- x xx x a . 42-x 0.2 0.4 0.6 0.8 3 Density (g/cm ) x - 1 1.2 1.4 (c) T = 250K Figure 3-16: Correlation between transition rates and density at various temperatures for a system of 5000 mers arranged in 20 free chains using a bin size of 5.23A x 5.23A x 5.23A. Each bin contains an average of 5.84 mers. 56 of the transition. This may be too large a time slice for the given studies. Reducing the cutoff time beyond this resulted in too few data points for meaningful analysis. Upon inspection of the transition angle histories for the most active angles, Figure 3-17, it was found that in the free chain system these angles oscillated about high energy states whereas the less active angles oscillate about the energy minima. Figure 3-17 depicts the dihedral angle trajectory for a single, representative, very active dihedral angle. As shown in Figure 3-17, the active angles are found to frequently transition between two high energy states and are unlikely to oscillate about the potential minima. This behavior was also observed by Boyd et al. [5]. In the networked samples, a large fraction of the segments with the highest transition rates tended to be adjacent to crosslinks. In the case of crosslinks, the potential employed does not allow for Van der Waals interactions between the monomers adjacent to the crosslink. Atoms which are separated by fewer than 4 bonds do not include Van der Waals interactions. For this reason, the crosslink has reduced mobility but the four mers bonded to the crosslink are less constrained than their neighbors. Movement of the chain produces transitions in the dihedral angle which contains this mer with additional mobility. This may also explain why at low temperatures, the transition rates remain elevated above that predicted by an Arrhenius dependence. At lower temperatures, the transition rates are dominated by relatively few dihedral angles which are trapped in high energy states. The effective energy barrier to transitioning is therefore reduced and rates are increased. These regions of high activity persist in the system even after nano-seconds of relaxation below the glass transition temperature. Whether they are artifacts of the extraordinarily high cooling rates employed is not known. 3.5 Correlation of Transitions along a Chain The correlation between neighboring transitions was calculated according to the methods outlined in Section 2.6 as a function of temperature. Referring to Figure 3-14, one finds that as the system temperature is lowered below the glass transition tempera57 7 C -0 4 C < < < 3 2:3 AI 3080 3120 3200 3160 Time (ps) 3240 Figure 3-17: Dihedral angle trajectory for a single, active dihedral angle. ture, dihedral angle transitions become heterogeneously distributed along the chain and self correlation of transitions increases dramatically. This is consistent with the notion that chains below the glass transition are "caged in" with heterogeneous distributions of surrounding volume, Figure 3-22. As the temperature is raised, self correlation decreases, but there are still significant correlations between dihedral angles separated by even numbers of bonds, Figure 3-18. This behavior has been seen in previous studies [5]. The correlation between changes in dihedral angles is not completely captured by the distributions shown in Figure 3-18. Motion of one angle does not always cause the second neighbor to transition but may cause significant changes in the neighboring dihedral angle. It seems that motion of a single dihedral angle almost invariably causes the second neighbor angle to move also. Shown in Figure 3-20 are the dihedral angle histories for neighboring dihedral angles as enumerated in Figure 3-19 at 100K. One finds this behavior is prevalent in the samples. The dihedral angle oscillates between two high energy states and each transition is accompanied by motion of 58 0.9500K 400K x 350K -300K - - 0.8 .- - .200K - OOK 0.50.4 0.37 0.4 LI 0.2 0 0 1 2 3 6 5 4 Correlation position (i±j) 7 8 9 10 Figure 3-18: Neighbor Correlation for system of 5000 united atoms arranged in 20 free chains at various temperatures. the second neighbor transition. Figure 3-20(a) and 3-20(c) show that rotation of a dihedral angle is accompanied by movement of the ±2 neighbors while Figure 3-20(b) and 3-20(d) show that the ±1 neighbor does not seem to be affected by changes in a given dihedral angle. In the case of crosslinked systems, the situation is slightly different. The dihedral angles which are adjacent to crosslinks may oscillates about the dihedral energy minima but undergo many more short lived transitions, Figure 3-21. There seems to be no correlation between dihedral angle switching on one chain attached to a crosslink and switches on other chains attached to the same crosslink. This suggests that communication of dihedral transitions across crosslinks is hampered by the limited mobility of the crosslink. 59 -k2 Figure 3-19: Map of dihedral angles 3.6 Tessellated Volume Voronoi tessellation was used to distribute the total system volume to each of the united atoms in the system [44]. In this method the united atoms are treated as reference points and regions are formed about these reference points such that each point within this region is closer to the enclosed reference point than to any other reference point. The Voronoi volume distribution gives details of how the mers are packed locally. The code used was developed by Richards [14]. With cooling, the distribution of Voronoi volumes becomes more sharply peaked as the mean value decreases Figure 3-22. As thermal motion decreases, atoms pack more tightly decreasing the volume per mer. It is also noted that heating above the previously determined glass transition temperature of 300K results in an increase in mers with the largest volume. Similar behavior has been observed by Sylvester et al. [39]. Though the maximum Voronoi volume measured was in the range of 80-100 A 3 , fewer than one half of a percent of the mers had volumes larger than 60 A 3 . As is evident from these broad distributions, the local environment varies significantly from mer to mer. The mean value is simply proportional to the total volume and therefore exhibits the characteristic change in rate of change at the glass transition temperature. The volume reported includes the volume of the mer itself. In order to obtain an estimate of the free volume surrounding each mer, the Van der Waals volume may be subtracted from the Voronoi volume. In essence, one assumes that the mers occupy a 60 Dihedral Angle Energy (kJ/mole) 25 20 15 10 5 5 30 -o g- 350 Dihedral Angle Energy (kJ/mole) 15 20 25 10 30 350 01 300 300 250 250 o o a200 a200 -@ 0150 150 0 -TI 0 100 100" 0 -sJ 50 50 20 60 40 Time (ps) 80 n, 100 (b) (a) #-2 5 Dihedral Angle Energy (kJ/mole) 25 20 15 10 5 30 80 60 40 Time (ps) 20 100 q1 Dihedral Angle Energy (kJ/mole) 25 10 15 20 30 350 350 02 300 300 250 250 200 -~ o $ 0200 -6 [150 *150 * * 8 100 1001 50 50 20 60 40 Time (ps) 80 ci )c~cic " ci ci 20 20 100 ~ 40 40 60 60 Time (ps) 80 80 100 100 (d) q-1 (c) 02 Figure 3-20: Dihedral angle value over time for active angle in networked system at 100K. 61 I A Dihedral Angle Energy (kJ/mole) 11 IA A in 1n 1a 9n 350- 25 12 0- 00 - 0 oa --+ . 5 10 20 30 40 50 60 70 80 90 100 Figure 3-21: T = 100K, networked system. 0.2 - 0.18- - - 0.16- - - 200K 300K 400K 500K 0.14 - o0.120.1 - EI z 0.08 - 0.06 0.04 0.02 0 10 15 20 25 40 35 30 Volume (Angstromse) 45 50 55 60 Figure 3-22: Tessellated volume distribution at various temperatures for system of 5000 monomers arranged in 20 free chains. 62 region of space given by the Van der Waals radius which no other atom can occupy. In reality, the mers are not hard spheres. Hence neighboring mers may penetrate this radius. Additional complexity arises from overlap in the Van der Waals volumes due to the connectivity of the mers. The bond length is less than half of the Van der Waals Volume. Overlap of the Van der Waals Volume occurs between adjacent mers and even second and third neighbors depending on the configuration of the chain. 63 Chapter 4 Deformation Results The material response was characterized as a function of deformation rate. The glass transition temperature is located using DMA type simulations. Both uniaxial compression and tension experiments are performed at a variety of temperatures and strain rates. The evolution of stress, internal energy, and chain conformation are observed with deformation. The Eyring model is applied to determine the activation volume for the system. Observations of the configurational transitions of the chain backbone are studied as indicators of mobility and the correlation between these transitions is investigated during deformation. 4.1 Dynamical Mechanical Analysis As stated previously, the state of a material is determined not only by a temperature but also a time scale. For this reason, additional specification was necessary to ensure that the sample was indeed glassy for the deformation rates employed. In Figure 39(a), the specific volume versus temperature indicated that the Tg for a network with 50 mers between crosslinks was 305K. Here, in order to verify that this T. is valid for the time scales of the deformation to be imposed, dynamical mechanical tests were first simulated. Experimental data obtained from dynamical mechanical analysis (DMA), Figure 4-1, shows that polyethylene undergoes a decrease of the storage (real) modulus in the vicinity of the glass transition temperature. During a 64 DMA experiment, a sample is subjected to small amplitude strain at high frequency and the generated stress is measured as the temperature is varied. The modulus is sensitive to both the temperature and the frequency of oscillation. A rapid change in the modulus over a given temperature range is an indicator that a transition in material behavior will occur for the given frequency in that range. The Young's modulus was obtained by estimating the slope of the stress strain curve over the first 3% of true strain for the simulation data at various temperatures and strain rates in order to characterize the behavior of the material. For temperatures above the glass transition, this becomes increasingly prone to statistical error as the sample displays a non-linear response even at these small strains. The results presented in Figure 4-2 agree well with the data presented in Section 3.2 suggesting that the glass transition is in the range of 275-300 K for a strain rate of 1010/s. At a strain rate of 5 x 1010 the glass transition temperature is significantly elevated to 330-370K. Data obtained at constant temperature and varying strain rate are plotted in Figure 4-3. This data indicates that the sample is glassy throughout the sampled regime at lOOK but that a transition may be in the vicinity of a strain rate of 108/s for a temperature of 200K. 4.2 Uniaxial Deformation Having established the glass transition temperature for this material model, the material behavior during deformation in the various regimes may be explored. Samples were loaded in the x-direction while the stresses in the y and z direction were maintained at atmospheric pressure. Stress components, energies, and a variety of chain statistics were monitored throughout deformation. Figures 4-4 through 4-12 were obtained from a constant strain rate uniaxial compression simulation conducted at 100K containing 7775 united atom beads arranged in a network with 160 chains. Figures 4-25 through 4-30 were obtained from a constant strain rate uniaxial tension simulation conducted at 100K containing 7775 united atom beads. These results are illustrative of the qualitative material response observed at all of the simulated strain 65 HDPE Dynamic Machanikal Analysis .2 - i,000,000 +2 -1uu -as. 40.000 100 E E-T Mari, Law, "F Tan DdIta + .. .L..... 0 -300 -200 0 -100 j 0 Z00 20X) '00 -. Temnperatw% 'IF Figure 4-1: Experimental graph of DMA curve for HDPE [15]. 2.6- 0 2.4 0 2.2C_ 2 x Ci) _ 1.8 - 0 2 .0 1.6- U) i 1.4 1.2 - 0.8 100 ' 150 ' 200 ' 300 250 Temperature E= E=5*1010 66 400 350 Figure 4-2: Estimated modulus for = 1 x 10 10 /s and i 7775 united atoms arranged in 160 networked chains. 1*1010 x o 1 - = 450 5 x 10 10 /s for a system of 2.'.A 0 x 2.5- X xX 0 0 - 2.42.3 CO S2.2 0 Cn2. 1 2- 0 1.9 x 1.818 10 o 10 1 10 101 100OK 200K - 10 11 Strain Rate (/s) Figure 4-3: Estimated modulus for various strain rates at lOOK and 200K for a system of 7775 united atoms arranged in 160 networked chains. rates below the glass transition temperature. Prior to deformation, the sample was relaxed at atmospheric pressure, but the internal stresses within the network remain quite high. There exist highly tensile Van der Waals forces and highly compressive bond stresses with negligible stresses due to bond angle flexure and dihedral angle rotation. Previous studies note this competition as the source of high internal stresses, but there are conflicting results in the literature as to the expected nature of these stresses within the dense network structure [12]. Studies by Chui et al. [8] and Sylvester et al. [39] find compressive Van der Waals stresses and tensile bonded stresses whereas bonds are compressed in this study and in studies by Roe et al [36]. This difference in initial stress state is attributable directly to the force fields employed. Chui and Sylvester employ a torsion potential which does not energetically differentiate between trans and gauche states. Biasing was accomplished by adding a repulsive Van der Waals interaction between atoms separated by three bonds along the chain. The equilibrium bond lengths in the Chui 67 and the Sylvester studies were shorter than in other studies. This was done so that the stretching of the bonds by the repulsive Van der Waals interactions resulted in a bond length of approximately 1.53 angstroms. In this study and that by Roe, the torsional potential incorporates a biasing between states and does not require additional biasing from the Van der Waals interactions. For this reason Van der Waals interactions are computed for atoms separated by 4 or more bonds. Atoms separated by 4 or more bonds will have an attractive Van der Waals interactions which leads to compressed bonds. 4.2.1 Compression Results from simulated compression experiments on a network of 7775 mers arranged in 160 chains is shown in Figure 4-4. The network response to the first few percent of strain is elastic with a calculated Young's modulus of 2.4 GPa and a Poisson ratio of 0.4. This agrees well with the results of Brown and Clarke [7] who find the modulus of a similar polyethylene model to be 2.4-2.5 GPa. Yield initiates at a strain between 10 and 12 percent and is followed by softening and then by deformation at a relatively constant stress and then strain hardening. The stress perpendicular to the load direction fluctuates about atmospheric pressure throughout the simulation. Qualitatively the network response is similar to that obtained in experimental studies of glassy polymers, but the quantitative results cannot be compared to experiment due to the high strain rates employed. The yield stress measured in this simulation was 260 MPa for a strain rate of 1.0 x 10 10 sec-1. The stress response was decomposed into its various contributions, thermal stresses, Van der Waals induced stresses, stresses due to bond streching, bond angle bending stresses, and stresses due to dihedral angle rotation, Figure 4-5. This provides the detailed stress response of the network to the imposed strain. In this figure, one notes the compressive bonded stresses and the tensile Van der Waals stresses at 6 = 0. In the load direction, the x-direction, the elastic response is governed by the linear relief of the tensile Van der Waals stresses. Though in this case the Van der Waals stresses go from tensile to compressive at the yield point, this was not characteristic of all 68 300- 250- 200- w150a) 100- 50- 0 0.1 0.2 0.3 True Strain 0.4 0.5 0.6 Figure 4-4: Compressive stress versus strain for uniaxial compression for = 1 x 1010/s and T = 100K for a system of 7775 united atoms arranged in 160 networked chains. systems tested. Some networks yielded before the Van der Waals stresses reached zero. The change in slope of the Van der Waals stresses and the bonded stresses, on the other hand, were characteristic of the network response. During the elastic portion, the stresses due to bond streching and bond angle bending both become increasingly compressive while the stress due to the rotation of dihedral angles is negligible throughout the deformation. For clarity, the initial stress was subtracted from the curves and the results of the stress partitioning in the x, y, and z directions are shown in Figure 4-6. Nonlinear behavior of the macroscopic stress strain curve, Figure 4-4, begins at strains of approximately 6 percent. Prior to this the Van der Waals stress, Figure 4-6, increases linearly in the load direction. At 10-12 percent strain, the Van der Waals stress is negligible in the load direction and with further deformation becomes compressive. The rate of change of the Van der Waals stress also changes at this point. In the y and z directions, the Van der Waals stresses become increasingly tensile until yield occurs and this tensile stress begins to be relieved. In the y and z directions the stresses due 69 250 200% 150 100- Thermal Stress --- -200 - a Bond Angle Bending 2- -200 0.1 0.2 0.3 J Interactions Bond Length 0.4 0.5 0.6 0.7 Bond Torsion 0.8 0.9 True Strain Figure 4-5: Partitioning of compressive stress for uniaxial compression along load direction. to bond angle streching become increasingly tensile throughout deformation. The contribution to the stress from dihedral angle rotation is negligible in all directions throughout deformation. Energy changes during the first 10-20% of strain are dominated by increases in Van der Waals energy and an increasing torsion energy, Figure 4-7. Due to the stiff nature of the bond stretching and bond angle bending potentials, very little energy is absorbed by these degrees of freedom during elastic deformation. There is an increase in the dihedral angle energy, Figure 4-7, as the shoulders of the torsion angle distribution, Figure 4-8, become broadened. High energy angles are adopted making transitions between states increasingly favorable. The bonded stresses pass through an inflection point at approximately 10 percent strain and yield follows in the range of 10 to 15 percent strain, Figure 4-6. During the elastic portion, the bonded stresses become more compressive in all directions as the average bond length decreases slightly. The mean bond length decreases from 1.52765 Angstroms, 1.5275 Angstroms , and 1.52713 Angstroms at 0%, 2%, and 10% strain 70 Thermal Stress .."."'Van Der Wasa - - Bond Streching 350 a 300 250 ..... ...... - 200 Interactions Bond Angle Bending Dihedral Angle Rotation 150 0 100 5 0 0 [ -S -100'0 0.1 0.2 0.3 0.4 0.5 0.6 True Strain 0.8 0.7 0.9 1 (a) x-direction 100 80 60 Thermal Stress ....Van Der Waals Interactions - - Bond Streching 0 Bond Angle Bending '-.Dihedral Angle Rotation .. ,, ' 0.7 0.8 0.9 1 0.8 0.9 1 40 20 CL S- 0 -20 -40 -60 -80 0.1 0.2 0.3 Strain 0.4 True0.5 0.6 (b) y-direction 100 - 80 - - 60 - a Thermal Stress - Van Der Weals Interactions Bond Streching Bond Angle Bending Dihedral Angle Rotation ,.."' 40- 0- I.2- 20- -20 -40 -60 -80 0 0.1 0.2 0.3 0.4 0.5 0.6 True Strain 0.7 (c) z-direction Figure 4-6: Partitioning of normalized compressive stress as sample is loaded in com- pression along the x-axis. 71 0.5- 4 Bond Stretching Bond Angle Bending - - Dihedral Angle Rotation 0.4 - a Van der Waals Interactions 0.30.2 # # X # a a#a###aa # - -0.1 -0.2 0 0.1 0.2 0.3 0.4 0.5 0.6 True Strain 0.7 0.8 0.9 1 Figure 4-7: Partitioning of normalized energy change per united atom versus true strain in compression. respectively. This change in mean bond length and a slight redistribution of bond lengths from 0% to 10% strain results in a 22% change in the bonded stresses, Figure 4-6, while only changing the bonded energy by 1%, Figure 4-7. Following yield, the mean bond length increases approaching the equilibrium length of 1.53 Angstroms, but remains slightly compressed throughout deformation. The distribution of bond lengths also becomes increasingly peaked about the mean value, Figure 4-8. The density, Figure 4-9, increases during elastic deformation with a characteristic rapid dilatation occurring just after yield. The decrease in density persists through softening then the density becomes constant beyond a strain of 25%. The average Poisson ratio in flow is 0.5. As the activation barrier for structural rearrangement is crossed, the material softens and the density decreases. The Van der Waals interactions, which are in effect holding the material together in the lateral direction, is repulsive at distances smaller than 4.50 Angstroms, has a maximum attraction at a distance of 4.99 Angstroms, and is weakly attractive at large distances, Figure 4-10. The re-shuffling of neighboring mers in a polymer as simple as polyethylene becomes 72 -- 0.0 E-0 -0.2 x E=1.0 0.08 0.07- -E0.2 0.04 0.0 x x 5- , 0.03 0.06- x , x 0.024-. 0.05- .x 5 0.04x 0.0 0.030.01 x 0.021 0.01 0.0 0.92 0.94 " xx x 0.9 2- -x 0.00 5 - x 1.02 0.98 1 0.96 Normalized Bond Length (/b q. 1.04 1.06 ~ .5 18 18 x . .9 Rdas xx .5 21 21 . An Bon 1.08 (b) Bond angle distribution (a) Bond length distribution Sr 0.08- j=0.2 x -.. 1x E= 1.0 0.07- 4-. 0.06- 3.5- 0.2 E.1.0 3- T 0.05 0.04 0.03- 1.5- x 0.020.01 0 0.5- 0.5 1 1.5 2 2.5 3 3.5 4 Dihedral Angle (Radians) 4.5 5 1 5.5 2 5 3 4 Normalized Distance (d/b,,) 66 7 7 (d) Radial distribution function (c) Dihedral angle distribution Figure 4-8: Comparison of structural statistics for compressive true strains of 0, 0.2, and 1.0. 73 0.95 0.945- E .940-- 6 0.935- 0.93- 0.925 0 0.1 0.2 I 0.3 0.4 0.5 0.6 True Strain 0.7 0.8 0.9 1 Figure 4-9: Density versus true strain for uniaxial compression possible when this maximum is crossed. This dilatation continues to a strain of approximately 20-25% strain. The energy stored in Van der Waals interactions and in dihedral angle rotations increase during this dilatation while the energy stored in bond stretching decreases slightly, Figure 4-4. In order to address concerns that the dilatation following yield was not a result of the barostat, the same initial structure was subjected to deformation with a prescribed Poisson ratio of 0.5, Figure 4-11 (i.e. under NVT conditions). The lateral pressure was negative throughout the elastic regime indicating that the material would compact if the sides were released. The pressure then increases to a positive value just after yield reproducing the tendency of the system to expand just beyond yield. Perhaps due to the suppression of the dilational response, the strain softening observed in the uniaxial simulations was not observed in the NVT case. A constant pressure regime follows suggesting the deformation is naturally volume preserving at this point. Similar observations were made in molecular mechanics experiments on atactic polypropylene by Mott et al. [29]. By observing system pressure in a model polypropylene system, they found that there is a tendancy to dilate in the early stages of plastic deformation. 74 0.5 0.5 Uci 0.5 O ---- -0.5 W -1 4 5 6 7 Distance (Angstroms) 8 9 -1 10 Figure 4-10: Schematic of Lennard Jones potential and gradient. In studies by Hutnik et al. [20] on a polycarbonate model system, a tendancy for the system to contract upon yield is found. The measures of rotation and elongation of bonds and chains for the uniaxial simulation are given in Figure 4-12. Compressive deformation results in an elongation of the average end to end distance of the chain. The rate at which the chain elongates is very small during the elastic portion of deformation and quickly increases. At a true strain of 1, the chain is approximately 60% of its maximum elongation. The chains rotate into the plane perpendicular to the load direction as is evident from the < P 3 > values in Figure 4-12. The percentage of dihedral angles in the trans configuration changes very little during the course of the simulation, Figure 4-8. It is notable that during the application of 100% true strain, the fraction of dihedral angles in the trans configuration changes by only 3%. This suggests that the 27 % change in chain length can be accomplished with the conversion of relatively few gauche configurations to trans configurations. Although the total number of dihedral angles in the trans state changes little, there is a significant amount of switching between states during deformation. Prior to and during yield the transition rate between dihedral angle states increases rapidly 75 350 --- - 300 -x x-direction y-direction z-direction 250 200a U) 0 (D 10- 0 IIPI 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 True Strain Figure 4-11: Compressive stress versus true strain for constant volume compressive deformation with e = 1 x 1010/s and T = 100K indicating that structural rearrangement of the chain is now energetically feasible, Figure 4-13. The transition rate plateaus as the dihedral angle energy plateaus. Flow initiates soon after suggesting that any additional input of energy is dissipated through rotation. The Van der Waals stress in the axial direction increases at a slower rate as it becomes compressive in nature, and the bonded stresses become less compressive. Figure 4-7 shows that the energy stored in dihedral angle rotations and Van der Waals interactions plateau as the bonded energy decreases due to the tightening distribution of bond lengths. Pole figure diagrams were constructed to obtain information concerning the rotation of bonds away from the load direction, Figure 4-14. The random distribution of bond directions in the initial state is shown for E = 0. Rotation of bonds away from the load direction is evident at strains as small as E = 0.1. The continued deformation promotes further rotation into the plane perpendicular to the load direction. A more complete view of the deformation induced orientation can be seen from 76 24 61 2375 70 21 20 6 65 1960 18- 17 0 0.1 0.2 0.3 0.4 0.5 0.6 True Strain 0.7 0.8 0.9 0 0.1 (a) Average end to end distance 0.2 0.3 0.4 0.5 0.6 True Strain 0.7 (b) Average chain angle, < 0.25 63 0.8 0.9 P3 > -x-direction y-direction ... z-direction -x 0.2 0.15 62 0.1 0.05 61 a.j 60 0 -0.05-0.1- 59 -0.15-0.2 58 -0.251 0 0.1 0.2 0.3 0.4 0.5 0.6 True Strain 0.7 0.8 0.9 0.1 (c) Average bond angle with respect to the load direction, < P 2 > 0.2 0.3 0.4 0.5 0.6 True Strain 0.7 0.8 (d) Bond orientation function, < P > -50.665 0.66 0.655 0.65 0.645 " 0.64 L 0.635 0 0.1 0.2 0.3 0.9 0.4 0.5 True Strain 0.6 0.7 0.8 0.9 (e) Fraction of Dihedral angles in the trans configuration Figure 4-12: Chain statistics versus true strain in compression 77 1 3+ gauche to trans o trans to gauche 0 gauche to gauche 20 0 CO - z () 0.0 0)0 0 C: 0 C: -- 0 0.1 0) 0 * . 0 Tru6 Strain . 0 . 0 . 0 . . 0.7 0.8 0. 1 0.9 1 2-- -4 0 I 0.1 0.2 0.3 0.4 0.5 0.6 True Strain Figure 4-13: Transition rate versus true strain for uniaxial compression constructed three dimensional distribution functions, Figures 4-15 through 4-23. These were constructed for two cases. The first considering only atoms which are on differing chains, Figure 4-15 through 4-19, and the second considering only atoms along the same chain, Figure 4-20 through 4-23. Though the full three dimensional profiles are provided, profiles of the distribution are provided in the x, y, and z directions for clarity, Figure 4-15 and Figure 4-20. In the case of atoms along different chains, the compaction of mers occurring during the initial elastic deformation is evident from the G(x, 0, 0) curve in Figure 415. The steady development of additional packing order in the load direction is evident from the growth and refinement of the peaks in G(x,0,0). In the plane perpendicular to the load direction, the distance to the first neighbor grows and the peaks flatten to a value of 1. This suggests that there is little organization of the mers located on different chains in this plane. The three dimensional distribution of mers along the same chain, Figure 4-20, shows that the chains are rotating out of the load direction even at small strains. The first and second neighbor peaks decrease in the load direction, i.e. becoming 78 xx 3, 'I. x x x x x x 0.1 0 (a) c XX xx x x x x. x - x XW x x I WX .Xx k I., x x xX x XA x x x x x x x it xxx x. x ax It )k.-x w x X#x x x xo Xk X w xxx AM . **xx4xx xx-t x -k x xx x xx x"0J x-Ax 'Am;*xxo xl-x -kxf : xx x x 9 k I ;,,VIA IA x x xx x 0 xx x x x Ixx Y'x x x x x x x xx 9 'no -x. it r x xx xA 3vt x x 0.4 (d) 0. 2 (c) x x x x x x x x (e) c = 0.6 Figure 4-14: Pole figures for uniaxial compression 79 more peaked in the plane perpendicular to the load direction. The three dimension radial distribution data suggests that as chain segments rotate out of the load direction, they become on average stacked one on top of the other. Already directed in the plane perpendicular to the x direction, further compression can no longer rotate the chains easily. Instead, the Van der Waals interactions become compressive and mers on neighboring chains are compacted together. This action tends to stretch the bonds which counteracts its initial compression, Figure 4-24. This also suggests that the Van der Waals stresses in the axial direction are being relieved for two reasons. Firstly, the < P2 > correlation function in Figure 4-12, the pole figure plots in Figure 4-14, and the radial distribution profiles in Figures 4-15 through 4-23, indicate that the highly compressed chains are rotating out of the load direction. The change in the value of < P 2 > is on the order of 1.5 degrees and linear in the elastic regime. Slight orientation is also evident in pole figure diagrams at strains of only 10%, Figure 4-14, and becomes increasingly noticeable at larger strains. Since these segments have highly tensile Van der Waals interactions aligned with the chain axis, rotation of the axis out of the load direction results in a relief of the tensile stresses oriented along the load axis. The Van der Waals stresses become increasingly tensile in the plane perpendicular to the load direction (i.e. the y and z directions). These results reinforce the data by Chui et al. [8] and Bergstrom et al. [3] which suggest that it is rotation of the chain segments which is responsible for the initial network response. Secondly, the mer mobility is severely limited by the Van der Waals attractions. Unable to undergo large structural rearrangements, compression results in a decreased mean spacing of united atoms along the load direction and relief of the tensile Van der Waals stresses. Density increases during the elastic portion of deformation necessarily require a decrease in the mean spacing of mers. This decrease in mean spacing is seen even in the radial distribution profiles. Since bond angles and bond length distributions and mean values change very little, this necessarily translates into a compaction of neighboring chains. As the mean spacing in the load direction decreases 80 f ,,*0~ / 108 -. *8 / 0.0 w I (0.4 - -0G(0.0z)1 0.4 I) b)2 _4 (a) c = (b) 0 010. =0. 1 1.2 - 0~''*~ 0.6 - 0.4 - 0.2 - .0 -- GO ~* 0 * / 2 ~0,0) (d) e = 0.4 1.4 -- 1.4 / / I (c) E = 0. 2 1.2 I 2 1.2 0 0.8 7 -. 0. 0..4- o.4.- * ~ 0,0) 0.2 - 0.4 0.2 no. 2 (Ang.t-) (e) c = 0.6 at ne (ArNstr-m) (f) C = 1 Figure 4-15: Line profiles of three dimensional pair distribution for monomers along different chains. 81 z Slice 1 Slice 2 X Slice 3 Slice 4 Slice 5 Figure 4-16: Diagram of three dimensional distribution profile slices. and bonds are compressed, the bonded stresses in the y and z directions become more compressive triggering expansion of the chains in plane perpendicular to the load direction. This results in a decreased mean spacing and increasingly tensile Van der Waals stresses in this plane. The lack of structural relaxation and dissipation is also reflected in the rapidly increasing system energy throughout elastic deformation. 4.2.2 Tension The results for a simulation of 7775 united atoms arranged in a network of 160 chains which was deformed at a true strain rate of 1 x 1010 per second are presented in Figures 4-25 through 4-30. The initial elastic response in tension was characterized by a Young's modulus of 2.3 GPa and a Poisson ratio of 0.35, Figure 4-25. The Young's modulus was calculated using the stress strain data for the first 3% of strain. Nonlinear behavior is observed in the stress strain data at 5-7% strain and yield occurs in the range of 10-15% strain at a stress level of 200 MPa, Figure 4-25. It is well known that yield in polymers is pressure dependent [26]. The tensile yield stress is lower than that found in compression due to the difference in the state of hydrostatic pressure between compression and tension simulations. Note that the macroscopic 82 Slice 1 1.5 20 20 15 G'15 0 N 10 15 52 N 5 20I a 0 0 5 10 x 15 20 10 20 y (Angstroms) 0 0 1 x (Angstroms) Slice 3 Slice 2 N 0.5 20 10 20 20 15 15 N 10 10 5 0 0 5 10 y 15 0 20 5 > 1I 5 10 x 15 20 15 20 Slice 5 Slice 4 0 0 10 y 15 10 0 20 5 10 x Figure 4-17: Three-dimensional pair distribution for monomers on different chains with c = 0 83 Slice 1 1.5 20 20, 015 E 0 110.10 Cn .. 1 N 0 20 0 10 x 5 15 0.5 E10 20 0 0 y (Angstroms) x (Angstroms) Slice 3 Slice 2 20 15 5 N 10 0 5 10 15 10 y 5 0 20 y 15 20 15 20 Slice 5 Slice 4 : .10 :5 - : x' d5 : ,: : -. :: .::o . : v 15 0 5 10 x 15 0 20 5 10 x Figure 4-18: Three-dimensional pair distribution for monomers on different chains with c = 0.1 84 Slice 1 1.5 20 20 ' 151- 15 E 2 1 10 N 10 <5 N 20jq 5 120 10 5 10 x 15 10 0 y (Angstroms) 20 0.5- 0 x (Angstroms) Slice 3 Slice 2 20 20 15 15 N N 10 in. 5 0 0 5 10 y 15 0 20 5 10 y 15 20 15 20 Slice 5 Slice 4 20 15 >'10 5 0 0 5 10 x 15 0 20 5 10 x Figure 4-19: Three-dimensional pair distribution for monomers on different chains with c = 1.0 85 - 2 (2,y,2) I 3- 35- 2.52- j2. 2 - .2 0.6 - 12 a 6 4 0421222 (A22.12111) 2 0 ,0. DIM222 (A22212.2) (b) c = 0.1 (a) c = 0 .-..0(2 y.2) .2.5 - 22 2-- .2- 1.6 0.- 110.5 -0 2 2 Dist-nc2 (Angst-rn2) 02 4 04 12a Distwnc (Angst-.m) a 10 (d) e = 0.4 (c) c = 0.2 4.4.44 3.5I 3.5 32.1 '12. - .5-1.2 1s1. 0.5 - 0.5 1 2 4 (Ang2t-.) 042.22. (Angt-.211) (e) c = 0.6 (f) E = 1 D4.121n Figure 4-20: Line profile of three dimensional pair distribution for monomers along the same chain. 86 Slice 1 20. 0.9 0.8 15. 0.7 0.6 0.5 0.4 E N 5 10 0.3 0.2 201 20 10 0 5 10 x 15 20 0 0 y (Angstroms) x (Angstroms) Slice 3 Slice 2 N 0.1 20 20 15 15 10 N10 5 0 0 0U 0 5 10 y 15 20 5 10 y 15 20 15 20 Slice 5 Slice 4 20 20 >,10 5: 00 5 10 x 15 0 0 20 5 10 x Figure 4-21: Three-dimensional pair distribution for monomers on the same chain with c = 0 87 Pp Slice 1 20 0.9 15 20, 0.8 ------- ' 15. 0.7 E 0.6 10 N 0.5 0.4 10- 51 - 110.3 0 20 5 0.2 20 ---- --0 0 5 10 x 15 20 10 . 0 0 y (Angstroms) 0.1 x (Angstroms) Slice 3 Slice 2 20 20 15 14 N 10 5 0 5 10 y 15 0 20 5 10 y 15 20 15 20 Slice 5 Slice 4 20 20 15 15 >'10 >'10 5 5 00 0 5 10 x 15 0 20 0 5 10 x Figure 4-22: Three-dimensional pair distribution for monomers on the same chain with E = 0.1 88 Slice 1 0.9 ...... 20 . 0.8 [lilt15 , 0.7 0.6 0.5 E 5 [ 0.2 20A 201 10 0 5 20 10 x 15 0 0 y (Angstroms) 20 0.4 I0.3 0.1 x (Angstroms) Slice 3 Slice 2 20 N 15 15 10 N 10 5 0 0 5 10 y 15 00 0 20 5 15 20 15 20 Slice 5 Slice 4 20 20 15 15 > 10 >110 5 0 0 10 y 5 5 10 x 15 0 0 20 5 10 x Figure 4-23: Three-dimensional pair distribution for monomers on the same chain 1.0 with c 89 A -41$ 4- f t t ft ft Figure 4-24: Schematic of chain interactions along compressed axis. density during deformation in the system undergoing tensile deformation is lower than that for a system undergoing compression deformation precisely because of this hydrostatic pressure difference. Following yield, the material softens. The stress remains relatively constant between 20% and 50% strain. Beyond 60% strain there are indications of hardening. The stresses in the y and z directions fluctuate about atmospheric pressure throughout deformation. The initial structure contains bonds which are compressed and tensile Van der Waals stresses. The initial stress is subtracted from the response and the results for stress in the x, y, and z direction are shown in Figure 4-26. As in the case of the compression tests, the Van der Waals interactions dominate the elastic response in the load direction. These interactions become increasingly tensile in a linear fashion in the load direction while the tensile Van der Waals stresses are relieved in the y and z directions. The stresses due to bond angle bending and dihedral angle rotation remain negligible throughout deformation in all directions. In the x, y and z direction the bonded stresses become less compressive. Bonds are being slightly decompressed as indicated by the relief of compressive bond stresses in all directions. This decompression is minor and does not affect the bond energies significantly, Figure 4-28. The mean bond length is 1.52765 Angstroms, 90 300 x-direction - - y-direction 250- z-direction 200a 150 - 100 (0 50 -5 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 True Strain Figure 4-25: Tensile stresses versus true strain for uniaxial tension with = 1 x 1010/s and T = 1OOK for system of 7775 united atoms arranged in 160 networked chains. 1.52766 Angstroms, and 1.5284 at 0%, 2%, and 10% strain respectively. There is an accompanying redistribution of bond lengths and a 32% change in bonded stresses and a 0.5% drop in bond energy from 0% to 10% strain. At yield, the Van der Waals stresses decrease. As the atoms are re-shuffled they find lower energy positions with an accompanying lower stress. The stresses remain more tensile during flow than they were prior to deformation. Also the level of the Van der Waals stress in the axial direction remains relatively constant with further deformation. In the plane perpendicular to the load direction, the stress response is characterized by an alleviation of internal stresses throughout deformation. The bonded stresses become less compressive and the Van der Waals stresses become less tensile, Figure 4-26. This is attributable to the highly stressed backbone rotating out of the y-z plane and to the decompression of the bonds. The structure has an initial Poisson ratio of 0.35 in the elastic regime. The density decreases by 4% from 0% to 15% strain, Figure 4-27. This large decrease in volume 91 200 - Thermal Stresa ....Van Der Weals Interactions Bond Streching a Bond Angle Bending 16 0Dihedral Angle Rotation - 0 1 14 0 12 0 0 I' - s agtr % .5 - 10 0 * 5 60 4 0. . 2 0 0.2 0.1 0.3 0.4 0.5 0.6 True Strain 0.7 0.9 0.8 1 (a) x-direction - 150 . - a -- - Thermal Stress Van Der Waals interactions Bond Streching Bond Angle Bending Dihedral Angle Rotation 100 50 ' 0 1= o - .. -+- -50 . . . . .. - - - -------------- -.- I .,-,.PIN -100 0.1 0.2 0.3 0.4 0.7 0.5 0.6 True Strain 0.8 0.9 1 (b) y-direction Thermal Stress ..... Van Der Weals interactions - - Bond Streching a Bond Angle Bending -Dihedral Angle Rotation 150 , 100 . Is- so CL -- 0 ---.-.- --- --- --- -... --.. -50 -100 0 0.1 0.2 0.3 0.4 0.5 0.6 True Strain 0.7 0.8 0.9 1 (c) z-direction Figure 4-26: Partitioning of tensile stresses versus true strain for uniaxial tension loaded in the x-direction. 92 0.95 I 0.940.930.92K 0.91 - 0.9 - 0.890.880.870.860.851 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 True Strain Figure 4-27: Density versus true strain for uniaxial tension is sensitive to the deformation rates employed. The magnitude of the drop is smaller for lower strain rates. The same large density drops were observed in work by Brown and Clarke [7]. The slope of this drop then decreases after yield as the Poisson ratio approaches 0.5. There is a minor decrease in the density over the remainder of the deformation. The inter molecular spacing is increased during this dilation since the bond length and angle changes are too small to account for the density changes. The increasing mean spacing would result in an increasingly tensile interaction until neighboring mers cross the maximum in the gradient of the potential, Figure 4-10. The energy change is dominated by increasingly energetic Van der Waals interactions, Figure 4-28. During elastic deformation, atoms are trapped by their neighbors unable to undergo large structural rearrangements. The imposed deformation results in an increased mean spacing between chains and hence an increase in energy as atoms are displaced from their minimum energy configurations. Energy changes seem to be confined to the Van der Waals interactions as the energy stored in bond stretching, bond angle bending, and dihedral angle rotation remain relatively constant during elastic deformation. Further deformation results in decompressed bonds and hence 93 0.7Bond Stretching Bond Angle Bending 0.6 - - - Dihedral Angle Rotation a Van der Waals Interactions - 4 0.5 - a33 aaa, so#3#*### 0.40.23W 0.2 - -0.1 -0.2 0 0.1 0.2 0.3 0.4 0.5 0.6 True Strain 0.7 0.8 0.9 1 Figure 4-28: Partitioning of energy change per united atom versus true strain for tension. a decrease in the bond energy. The torsion energy increases then plateaus beyond yield. Observations of the structural statistics, Figure 4-29, shows that the bond length distribution narrows with deformation and the shoulders of the dihedral angle distribution broaden. The radial distribution shows an increasing peak corresponding to the trans configuration. There is also a slight shift in the mean bond length as the bonds are decompressed. During the elastic portion of deformation the chain end-to-end distance does not change. Chains are elongated by 44% at a true strain of 1, Figure 4-30. The chains rotate into the load direction as the sample is strained as indicated by the decrease in < P3 >, which measures the average angle made between the end points of a chain and the load direction. both segments and bonds are rotated into the load direction as indicated by < P 2 > and < P >. Though the fraction of dihedral angles occupying the trans configuration changes twice as much as it does in compression, there is only a 6% change in the fraction trans as the chain is elongated by 44%. 94 0.07 =0 .x x 0.06 .X ~ ~ 0.0 0.1 = x X E E E= 0.1 4 -I =1.0 x E =1.0 0.03 x x 5- 0.05- 0.0 '3 -a 0.02 0.04- 0.0 '2 .0.03 - 2- '5 0.01 xx x 0.020.0 0.01 0.005- x 0.94 xx 1.02 1 0.98 Nonnalized Bond Length (/b u) 0.96 1.06 1.04 1.7 1.8 1.85 1.9 Bond Angle 1.95 2 (Radians) 2.05 2.1 2.15 (b) Bond angle distribution (a) Bond length distribution 0.08 - 5- _E-0 E=0.1 x E.1.0 0.07- 1.75 4.5- x x E=0.1 E11.0 40.063.50.05 - 3- 0.04- 62.5 - c0 0.03- 2 1.5 - 0.02 x - 1- 0.5 - 0.01 - 1 1.5 2 3.5 2.5 3 Dihedral Angle (Radians) 4 4.5 5 5.5 Normalized Distance (d/b,,) (d) Radial distribution function (c) Dihedral angle distribution Figure 4-29: Comparison of structural statistics for tensile true strains of 0, 0.2, and 1.0. 95 27 60 26 I !55023'U 45- 2221 - W 40- 2019- 35- 18 17 -- 0 0.1 0.2 0.3 0.4 0.5 0.6 True Strain 0.7 0.8 0.9 1 30 0.1 0 0.2 0.3 0.4 0.5 0.6 True Strain 0.7 0.8 0.9 (b) Average chain angle, < P 3 > (a) Average chain end to end distance 57- 0.3 -- 5655- 0.2 -- S54- o53 0.1 - - A 52- -L0 ,- 51 50 0.1 -- 4948 0 0.1 0.2 0.3 0.4 0.5 0.6 True Strain 0.7 0.8 -0. 0.9 - 0.2 0.3 0.4 0.5 0.6 True Strain 0.7 0.8 0.9 (d) Bond order parameter, < P > (c) Average segment angle with respect to the load direction, < P 2 > 0.69 0.1 ---- 0.68 0.67 0.66 0.65 0.64 06 0.1 0.2 0.3 0.4 0.6 0.5 True Strain 0.7 0.8 0.9 (e) Fraction of dihedral angles in the trans configuration Figure 4-30: Chain statistics versus true strain for uniaxial tension. 96 3- CL cld) C 0 I- C-2 -3) - + o o -4''''' 0 0.2 0.4 0.6 True Strain gauche to trans trans to gauche gauche to gauche 0.8 Figure 4-31: Transition rate versus true strain for uniaxial tension. As the material flows at constant stress levels, energy changes and transition rates plateau. At this point any additional input of energy is dissipated through structural rearrangements. Though the transition rates are significantly higher following yield than they are in an undeformed sample, the fraction of dihedral angles in the trans configuration changes by only 6% percent from 0% to 100% true strain, Figure 4-30(e). In fact, this fraction does not change at all from 0% to 50% true strain. Referring to Figure 4-31, one finds that the transitions from the Gauche state to the Trans state are matched on average by the transitions from the Trans to the Gauche state. Therefore, the average configuration of the system does not change while individual dihedral angles are in fact sampling multiple states. Construction of pole figures, Figure 4-32, shows an increasing alignment of chains with the load direction. This alignment is evident at strains of 0.1 and becomes increasingly visible in the plots. From these measures of orientation, one deduces that the change in the Van der Waals stresses is attributable both to an increase in mean spacing of particles and to rotation of the chains in the load direction. As there are highly tensile Van der Waals 97 x X ,I- K x x T' W. x, x x x = 0.1 (b) (a) f = 0 xm x x xx 'k Ix x x x Nx gab ; A Wx fi 1; x x _111 v x x Vi x x x 4 ex x x x x x x x 04 x 4x t li x x x x 0. 4 0. 2 (d) x x I x# xxxxx xx x rx IFx x xx waxw x x ON x x x wN 'A (c) c xxx x 'k x t Ox x x awxx x x x x x 17 10 fxf * x x x x ; x, -. x x x x x x 'kx x* x it . x 0 (e) c jF xmks * x xr x x x x i-x _x x il x x.0 xIN x xx!x x;wx Ipx x 0. 6 Figure 4-32: Pole figures for uniaxial tension, 98 IF x x x*N j x stresses along the chain backbone, rotation of these chains into the load direction results in the observed increasingly tensile Van der Waals stresses. 99 0.15 - 1=0.1 = 0.2 - F = 0.3 - -- 0.1 E =0.6 S---=0.8 - .0 E z 0.05- 10 15 20 25 30 35 40 Volume (Angstrom3) 45 50 55 60 Figure 4-33: Tessellated volume distribution at various strains for system of 5000 mers tested in uniaxial tension at 1OOK for a strain rate of lx 10 10 sec- 1 . 4.3 Evolution of Tessellated Volume The evolution of the tessellated volume with deformation indicates how the volume is redistributed locally amongst the atoms. During elastic deformation in both compression and tension, one notes a broadening of the free volume distribution, Figures 4-33 and 4-34. The curve corresponding to those mers with free volumes from 10 to 21 A, i.e. those which are tightly packed, does not change much. The distribution then becomes static beyond a strain of 0.3. This corresponds to the initiation and continuation of flow. This indicates that both the local volume distribution and the macroscopic volume are static during flow. In addition, the volume distributions during flow in tension and compression are similar to those found at the temperatures around the glass transition temperature. The broadened shoulder and peak heights are similar. This suggests that the volume distribution in flow evolves to a state near the glass transition temperature after which molecular mobility can accommodate further deformation without increases in volume. 100 0.15-- CE=0.1 - e=0.2 --- e=0.3 - =0.6 S=0.8 0.10 CO) .0 E z 0.05- 0 10 ' 15 20 25 30 35 40 Volume (Angstrom3 ) 45 50 55 60 Figure 4-34: Tessellated volume distribution at various strains for system of 40000 mers arranged in 160 free chains tested in uniaxial compression at lOOK for a strain rate of 1x10 10s- 1 . 101 4.4 Strain Rate and Temperature Dependence of Yield Stress 4.4.1 Stress Response The response of the network at various temperatures and strain rates was determined. Presented in Figure 4-35 are stress strain data for two temperatures each tested at two strain rates. As previously stated, the temperature and strain rate are both important in determining the material response. If the strain rate is slow enough or temperature high enough, the response is leathery or rubbery. This difference is exhibited by the samples shown in Figure 4-35. At 1OOK the model is well below its glass transition temperature while at 450K it is well above the glass transition temperature. At 100K, one notices that the magnitude of strain softening increases with increasing strain rate but the proportion of softening to yield stress remains relatively constant. The flow stress as increases with increasing strain rate. Though as the strain gets larger the response is similar. At 450K there is no softening as the material exhibits a rubbery behavior. The density change as a function of strain is provided in Figure 4-36. Note also that the dilation seen after yield in samples below the glass transition is sensitive to the strain rate. At slower rates the magnitude of the dilation becomes smaller but is always present. Though one might expect constant volume deformation for a rubbery material, the model being tested is known to crystallize easily given enough time and temperature. At 450K, one must worry about crystallization of the model. This may explain why the density at 1 x 10 10s- 1 increases more rapidly than at 5 x 10's--1. Given more time, the sample may crystallize further. This may result in additional hardening of the material as the strain is increased. Attempts to characterize amounts of crystallinity through observation of the radial distribution function were unsuccessful. There is little change in the radial distribution function as the sample is deformed. The thermal fluctuations at higher temperatures broaden the peaks and make interpretation of the distribution changes 102 4 00 -- - - - - -- - - - ' ---- -A-- F 350 - - -a~ S300 C 0 0 250 200 -- C lO-,E-110- S150 U) S100 -- 100K, E=51X 10 -- 450K,sE= 5x101 50 0 0 0.2 0.1 0.4 0.3 0.5 0.6 0.7 True Strain Figure 4-35: Stress versus strain for various temperatures for a system of 7775 united atoms arranged in 160 networked chains. -. 0.05 ...- . ...... . . - 100K, E = 1Xl -- - 100K, F =5x10 -- 450K, E = 1X10 - - -450K= 5 10 0.04 0.03 ... .. .... 450K~e ......... - - - . .. ...... sx *. ... 9. 0.02 -. -. .... ..... 0.01 - 0 ..... -. ... ..... .. ..... -.. >-0.01 . -0.02 -0.03F-00 A .......... I 0 0.1 0.2 0.3 0.4 True Strain 0.5 0.6 0.7 Figure 4-36: Change in density versus strain for various temperatures for a system of 7775 united atoms arranged in 160 networked chains. 103 2-- 1.8 450K, e = 0 450K, e = 0.8 1.6- 1.4 1351.2- 0- 0.2 - 0 0 1 2 3 Normalized Distance M/ 4 5 6 Figure 4-37: Change in radial distribution function during the application of compressive deformation at 450K. difficult. In addition, there may be crystalline and amorphous regions within the same sample cell. From the radial distribution function, Figure 4-37, there is little change in the short range peaks at a normalized distance of 1 and 1.63. One finds a slight sharpening of the peak at a normalized distance of 3.25. The long range liquid like structure is compacted as g(r) is shifted to the left at normalized distances above 3.5. It appears that the long range order is being altered through the application of deformation. Although changes in the density suggest crystallization, the crystallites may be dispersed in an amorphous structure and so the radial distribution function does not capture the appearance of crystallization clearly. 104 Table 4.1: Activation parameters for a selection of glassy polymers [1] Polymer £ (nm3) Polystyrene Polycarbonate Polyethylene-terephthalate 4.4.2 4.95 18.4 17.2 Activation Parameters The yield stress was determined for various strain rates, shown in Figure 4-39, in an attempt to determine the activation volume as described in section 1.3 for the Eyring model. The close agreement in yield stress measurements for different initial samples leads to confidence in the results. A least squares fit to the yield stress vs strain rate data as shown in Figure 4-39 results in an activation volume of Q = 0.18 nm 3 at 100K and 0.21 nm 3 at 200K. The RVE in our simulations ranged in size from 105 nm 3 for a cell of 4500 mers to 240 nm 3 for a cell of 10000 mers. Referring to Table 4.1, the activation volume obtained for this model is over an order of magnitude smaller than the experimentally determined values for various other polymers. Noting the simpler chemical structure of polyethylene compared to polycarbonate, polystyrene, and polyethylene-terephthalate, Figure 4-38, one would expect it to have a lower activation volume since it does not contain large side groups or ring structures. Molecular mechanics studies on atactic PP with an RVE of at most 39 nm 3 by Mott and Argon [29] suggest the shear transformation strain should be on the order of 1-5%. Assuming a transformation strain of 1% and 5%, the transformed volume, Va, according to Equation 1.3 is then 4 nm 3 and 20nm 3 , respectively. At the given densities, this represents cooperation by 150-800 mers. At 100K and strain rates above 2 x 10"/s, there is an obvious deviation in the behavior of the yield stress. A more rapid increase with strain rate which corresponds to a smaller activation volume of 0.03 nm 3 . Assuming a strain of 1% and 5%, this suggests cooperation by 20 mers and 120 mers, respectively. Though this transition may be gradual and may involve many intermediate modes, the inaccuracy of the 105 a) I I -F-C-Cm4n H 0) polystyrene b) C) / -- C) "7 C--CHsCH4t 4: terephthalate group ethylene group carbonate group Figure 4-38: Chemical Structure of (a) polystyrene, (b) polyethylene-terephthalate, (c) polycarbonate 106 measured yield stress and the scatter in data does not allow for the reliable interpretation of these intermediate modes. The structural relaxation of the glassy polymer may be imagined as being composed of many relaxation modes each of which operates with a given time constant at a given temperature. Deformation processes which occur on time scales shorter than this relaxation time constant will not sample that particular relaxation mode. On the other hand, a mode will be extremely active in deformation processes occurring on time scales much longer than its time constant. The yield stress at realistic strain rates was determined by extrapolation using the calculated activation volume at the lower strain rates. The yield stresses at 100K and 200K at a strain rate of 0.1 per second are -138 MPa and -429 MPa respectively. These negative values for the yield stress do not come as surprises. As suggested by the modulus changes observed as a function of strain rate in Figure 4-3, this model polymer may no longer be glassy at experimental strain rates. The assumption of isostructural deformation may not hold under these circumstances. In addition, one expects additional relaxation modes to be activated as the deformation rate decreases. An analogous error would be made if one were to extrapolate the behavior at 100K for strain rates above 2 x 10 10 /s to lower strain rates at that temperature. 107 4- 3.5 - - o 1OOK > 100K 200K + 200K x 200K - - 8000 mers 4500 mers 8000 mers 4500 mers 10000 mers - - a. 0.55 9 . Iog 1o(e) 0 05 1 Figure 4-39: Summary of yield stress obtained through simulations of uniaxial compression for various initial networked structures, strain rates, and temperatures. 108 4.5 Transitions and Yield By monitoring transitions rates as a function of deformation, it was possible to observe changes in material behavior during yield. Referring to Figure 4.5, one finds that transition rates quickly increase during yield to relatively constant levels as flow initiates. Since very few transitions are recorded during deformation at low temperatures, ensemble averaging would be necessary to obtain more accurate profiles of the transition rates. Figure 3-13 shows that various initial samples produce very similar transition rate profiles. In fact, the cross-linking of the system seems to make no difference in the profile obtained. This indicates that at the deformation rates used in these studies, the rate of configurational transitioning is governed by the local interactions and not by chain connectivity. The general features of a significant increase in transition rate leading to and just after yield and a decrease in the slope of this transition rate as flow initiates is seen at all strain rates and at all temperatures tested. Increasing the temperature, Figure 440(b) and 4-40(c), or increasing the strain rate, Figure 4-40(d), both serve to increase the transition rates. The difference in transition rates at a constant strain rate for various temperatures cannot be explained by taking into account the transition rate of a sample at the given temperature with no deformation. One might expect that not all transitions contribute equally to the accommodation of deformation. Some extraneous transitions occur. In an attempt to correct for this, the transition rate of a sample with no deformation was subtracted, Figure 3-12, from the transition rate at that same temperature for the various strain rates but failed to collapse the data, Figure 4-40(b) and Figure 4-40(c), into a single curve. The initial drop in transition rate seen in the first few percent of strain is most probably attributable to the increase in density during compression. The increased density in the cell stifles transitioning. At low transition rates, the values measured are statistically uncertain. Interpretation of this initial drop may simply be a consequence of the small number of transitions occurring within the given time period. One finds the most variation amongst different samples in this low transition rate regime, 4- 109 40(a). Before yield, transition rates at a given temperature appear to rise at similar rates at various strain rates, Figure 4-40(d). Beyond yield, the transition rates increase as the strain rate is increased. This suggests that as the strain rate increases, accommodation of deformation through the re-shuffling of neighboring mers is decreased. Instead, the deformation is accommodated through transitional rotations. This is consistent with the observed decrease in activation volume at the higher strain rates, Figure 4-39. It is clear that after yield the transition rates have increased significantly above the pre deformation values. At 250K the transition rates after yield are very similar to the transition rates at the glass transition temperature for the given strain rate. For example, at 250 K and a strain rate of 1 x 1010 sec~ 1 the transition rate is approximately 14-20 transitions per angle per nano-second, Figure 4-40(b). This transition rate corresponds to a temperature between 285K and 310K, Figure 3-12. The glass transition temperature at this strain rate as measured by the DMA type simulation was 275-300K, Figure 4-1. Similarly at a strain rate of 5 x 10 10sec- 1 the transition rate at 250K is approximately 30-40 transitions per angle per nano-second, Figure 4-40(c). This corresponds to a temperature between 335K and 360K. The glass transition temperature was measured to be 330-370K, Figure 4-1. As the temperature was lowered, the transition rates for a given strain rate decreased. If one compares the transition rates at high strain to the transition rates for static samples at various temperatures, these decreased transition rates correlate to a lower temperature. For example, at 1OOK for a strain rate of 1 x 10 5 x 10 10 sec- 1 10 sec- 1 and the transition rates were 3-5 and 10-20 transitions per angle per nano- second, respectively, Figure 4-40(b) and 4-40(c). These transition rates correspond to temperatures 210-235K and 265-310K respectively, i.e. the mobility has increased to a level observed in the material at much higher temperatures when under zero stress. Although the mobility is that observed at much higher temperatures, it is not as enhanced as that observed at Tg. The glass transition temperatures for the employed strain rates is much higher than this. At 1 x 10 10 sec 110 1 and 5 x 10 10 sec 1 the estimated glass transition temperature obtained from DMA type experiments is 270K and 350K respectively. The dependence of the transition rates on strain rate and temperature suggest it is a stress-biased thermally activated process. If one normalizes the number of transitions obtained in a given time unit by the displacement in that time unit, one obtains a curve which peaks during flow, Figure 4.5. This suggests that the number of transitions required to accommodate a given unit of deformation becomes constant during flow. The plateau value decreases with decreasing temperature or increasing strain rate. This suggests that as the time scale of molecular relaxation becomes smaller in comparison to the strain rate, the amount of deformation accommodated by dihedral angle rotation lessens. Instead the sampling of the dihedral angle configuration is stifled. At slower strain rates and higher temperatures more of the deformation is being accommodated by rotation. One expects that there are two sets of transitions occurring in the system. Those which contribute to accommodation of deformation and those which do not. The trend of an increasing number of transitions per unit of deformation as the temperature is increased or the strain rate is decreased might be an artifact of the fact that at higher temperatures there are more transitions even in the static case. To account for this, the initial transition rates were used to calculate the number of transitions which would have taken place in the static case during the time required to deform the sample by one Angstrom. This number was subtracted from the curves and the trend of an increasing number of transitions with increasing temperature or decreasing strain rate persisted. Referring to the stress strain plots for these systems, Figure 4-42, one finds that the plateau in the number of transitions per unit deformation seems to coincide with the end of softening. The transition rate appears to decrease in the initial elastic regime while increasing with the application of stress beyond the initial few percent of strain. The decrease in slope of the transition rate corresponds to a leveling off of the stress with increasing strain. 111 21 21 1.5 1.5 a) 1 00. 0.5 (D 0 .5 0 -L0.5 05 -0.5 1 + -p- + -.- 1 0 1 nw-c-1-4500-90-100K, E= 1X1010 ' s1 nw-1-8000-160-100K, E = 1x10 s7 fc-1-40000-160-100K, E= 1Xl0101 0S-1 nw-c-1-4500-90-100K, e = 5x10 s~1 1 nw-1-8000-160-100K, E = 5x10 ' s-' fc-1-40000-160-100K, E 0.6 0.4 True Strain 0.2 = 5X1010 S- 2 0 -9- 1 - -- 0.2 1 0.8 0 fc-1 -40000-160-100OK, -= 1 x0 aS1 10 tc-1-40000-160-200K, c = 1x10 s-1 fc-1-40000-160-250K, E = 1X100 S 0.4 0.6 True Strain 0.8 (b) Varying the temperature with e 1 x 10 10 /s (a) Comparison between free chains and network systems. 2- 1 = 21 1.5 1.5- Q- 0.50- , .5 - 02 CD 0 0- 0- c -0.5 - .5 - --- fc-1-40000-160-100K, E = 5xII10 S-1 -40000-160-200K, E = 5xl 10 s-1 fc-1 -40000-160-250K, E = 5xl 10 S-1 +fc-1 -eS 0.2 0.4 0.6 True Strain 0.8 (c) Varying the temperature with 1 i= -1 - tc-1-40000-160-100K, - - fc-1-40000-160-100K, e = 1Xl Sfc-1 -40000-160-100K, E = 5x10 s- -4- 0 0.2 E 0.4 0.6 True Strain = 5x101 0 s- 0.8 (d) Varying the strain rate with T 100K 5 x 10'1/s Figure 4-40: Transition rates as a function of true strain. 112 1 1 = lor 9a 0 a a 70) 6-' a 1 5- 0- 4- CL 05 3- 2- -- - 0 1 - fc-1-40000-160-100K, E = -- _C) r-1 0 0.1 0.2 0.3 0.4 True Strain a 0=5x1010 s0 0.5 0.6 0.7 - -- 0 0.1 0.2 fc-1-40000-160-200K, E= fc-1-40000-160-250K, E = 0.3 0.4 True Strain 0.5 0x10 0 aI 1x1 0 "s1 ' a's' 0.6 0.7 (b) Varying the temperature with i 1 x 1010/8 (a) Ensemble average at 100K. = 4.5- 42 .5- E 3.5 E E 2 2- 3- EC 2.5- -1 CL .50.5 -- -. -, 00 + -.- 0.1 0.2 fc-1 -40000-160-1OOK, E = 5x10 S-1 0 fc-1--40000-160-200K, E= Sxl1 s- 1 fc-1-40000-160-250K, E = 5x1 010 S7 0.3 0.4 True Strain 0.5 9 5x10 slx1 010 S~1 0 -4- fc-1-40000-160-100K, E = 5x1O' S-1 0.7 0.6 (c) Varying the temperature with c 5 x 101s 00K, E= fc-1 -40000-160-1 fc-1-40000-160-1OOK, E= 0 0.1 0.2 0.3 0.4 True Strain 0.5 0.6 (d) Varying the strain rate with T lOOK. = 0.7 = Figure 4-41: Number of transitions per dihedral angle per angstrom of deformation. 113 ---.----.-.-.--.-.-.-- 300 500 450 - 250 400f -- , -. -. - .. A, -. -. .... Cal ----. --. -----.. .. -. . ...... -.. - .... -.-.. lt a- 350 C 0 C 0 2200 6 300 a -D 0 -j 50 as 5 00 200 - -0 (75 65 1 0 0 2 150 2 a F- 50 100 8000, =1x10 - -. ---- 50 --0 C 0.2 0 secsec~40000, e = 1x10 sec1 40000, E = 5x1 0 sec- -8000, E=5x10 0.6 0.4 True Strain 40000 mers, T=100OK, E=1x100 sec-l 10 0.8 10 1 ---.--.-. --.. -. -. ..... --- -- 30 0 ......-.. 250 - -. --- 3 50 - - 0 0 --. 0.8 - ..... 1 e= .---- - -- - - -- - - 300 -. ..-. C - -. ,cs250 - (2 0.4 0.6 True Strain sec- (b) Varying the temperature with 1 x 1010/s. (a) Varying sample and strain rate. 35 0 0.2 - 0 K 40000 mers, T = 250K ,E= 1 x1 0 - ..-. .- 0200 ---- 0 150- 100 1000 -sec I - -- 40000 mers, T = 200K, e =5x1 00 sec -- 40000 mers, T = 250K ,E = 5xI 10 Sec 50 ..--..--.--. 0 -- 40000 mers, T = 100K, 0.2 0.4 0.6 True Strain 50 e =5xI1 0.8 (c) Varying the temperature with 1 e= -- - - -- - - -- 40000 mers, T = 100K, e= 1x1099SeC-- 40000 mers,T=100K,E_=1x10 sec- - - 40000 mers, T = lOOK, E =1 Xl10 sec' 40000 mers, T = 1OOK , = 5x1 010 sec' 0 0.2 0.4 0.6 True Strain 0.8 (d) Varying the strain rate with T 100K. 5 x 10-0/s. Figure 4-42: Stress as a function of strain for various systems. 114 = 4.6 Correlation of Transitions along a Chain During deformation at strain rates of less than 1 x 10 10sec- 1 , it was shown in Section 4.5 that the transition rate increases and then plateaus as flow commences. In addition to this increase in rate, the nature of coopertivity between these transitions changes. Prior to yield, the transitions are highly self correlated. Following yield and throughout flow, the transition coopertivity resembles that at the glass transition temperature. Throughout flow, one finds a decreased self correlation and strong correlations with dihedral angles separated by an even number of bonds. This behavior seems to be consistent across the investigated temperature and strain rate regimes. Unfortunately, statistical error in these correlations is much greater than in the determination of transition rates due to the small number of each type of transition. 115 0.9 0.9 0.8 0.8 0.7 0.7 0.6 0.6- !40.4 0.4 0.3 0.3 0.2 0.2 -static -- ____-'-----. ------'--_0.1 0.1 "0 2 4 6 Correlation position (i±j) (a) fc-40000-160-100K, 100K, i = 5 x 10 9 /s 0.9 8 10 T 0 2 4 6 Correlation position (itj) _ 8 10 T = 0.7 -- static --- f i static low 0.8 -- -- (b) fc-40000-160-100K, 100K, = 1 x 1010/s = flow --- 0.6 0.7 0.5 0.6 -0.4 0.4 -- - - - - - - - - - - - - - 0.3 0.3 0.2 0.2 0.1 0.1 0 2 4 6 Correlation position (i±j) (c) fc-40000-160-100K, 100K, E = 5 x 10 10 /s 10 8 T "0 2 4 6 Correlation position (i±j) (d) fc-40000-160-200K, 200K, i = 5 x 10 10 /s = Figure 4-43: Neighbor Correlation 116 8 T 10 = Chapter 5 Conclusion The results of this investigation are divided into three segments, structural behavior as a function of temperature, stress response to deformation, and chain configuration as a function of deformation. The chapter closes with remarks concerning future studies. 5.1 Structure The initial structure of the network was characterized as a function of temperature. Networks and free chain systems were created and simulations run at various temperatures. The structural response can be characterized as follows: " The glass transition temperature was found to decrease with decreasing molecular weight for free chain systems and increase upon the addition of crosslinks. As the temperature decreases below the glass transition temperature, dihedral angles no longer explore their configuration space and are frozen into a particular configuration. " Studies of transition rates as a function of temperature reveal that dihedral configuration transitions occur as single barrier events. At temperatures above the glass transition temperature, the activation energy for transitioning is found to be similar to the explicit dihedral angle potential. 117 " Deviation from an Arrhenius behavior is detected at temperatures below the glass transition. This occurred as dihedral angles became trapped by the surrounding matrix and oscillated between two high energy states. The effective barrier height for these transitions is lower and thus a higher transition rate than predicted by the Arrhenius extrapolation was observed. Attempts to correlate transition rate with local density and tessellated volume were unsuccessful using the techniques described in Chapters 3 and 4. " The cooperative nature of transitions along a chain was investigated and found to be significant at all temperatures below the glass transition. A strong correlation between dihedral angles separated by an even number of mers was found. The especially strong ±2 which is characteristic of kink type motion was observed as temperatures fell. The self correlation increases below the glass transition temperature indicating that many dihedral angles are oscillating between two or more states while neighboring dihedral angles are not transitioning at all. 5.2 Deformation The system response to deformation was determined as a function of strain from simulations which mimic uniaxial tension and compression experiments. The system density, energy, and stress response as well as analysis of yield stress as a function of strain rate and temperature can be characterized as follows: " Deformation response below the glass transition temperature captured qualitatively the experimentally observed behavior. DMA type experiments and studies of specific volume as a function of temperature were used to characterize the glassy regime. The system response is marked by a stiff elastic regime, yield, softening, flow, and hardening. * Yield in the system is accompanied by dilation which is followed by volume preserving deformation. The distribution of tessellated volume during flow ap118 proaches the distribution for an undeformed sample at the glass transition temperature. " Analysis of the yield stress as a function of temperature and strain rate below the glass transition temperature using the Eyring model reveals a small activation volume parameter. As the strain rate was increased at 100K, the activation volume describing the dependence of yield stress on strain rate changed. This may correspond to the change in activated processes. " Though the polyethylene chain theoretically provides the opportunity to directly observe shear events due to the small activation volumes involved, it has not yet been possible to correlate local motion in the system directly. 5.3 Configuration and Deformation The changes in configuration of the chains during deformation were observed as measures of mobility in the system. Characterization of the transition rates and correlations between transitions led to the following observations: " The self correlation decreases drastically during flow as the cages around atoms are destroyed by the deformation. A sharp increase in the transition rate per dihedral angle during and following yield occurs at all temperatures and strain rates. The transition rate depends only on temperature and strain rate while being insensitive to the initial structure and the chain connectivity. " It appears that in the elastic regime, the response is dominated by rigid motions of the chain. The application of stress modifies the dihedral angle transitioning barriers leading to yield and then flow. " The number of transitions per unit deformation also levels off during flow suggesting deformation is being accommodated by configurational changes along the backbone. A decrease in temperature or an increase in strain rate decreases the number of transitions per angstrom of deformation indicating a competition 119 between accommodation through chain angle rotation and chain slippage. At high temperatures or low strain rates, the chain is allowed to explore its configurational space and accommodates the deformation through these configurational changes leading to a higher number of transitions per angstrom of deformation. At lower temperatures or higher strain rates, this mechanism is being frozen out and the chains react in a more rigid fashion with fewer configurational changes per angstrom of deformation. * The flow state for samples deformed at temperatures near the glass transition temperature seems to resemble the state of the material at the glass transition temperature. The transition rates during flow approach those of an undeformed sample at the glass transition temperature. The neighbor correlation function calculated during flow also resembles that calculated for an undeformed sample at the glass transition temperature. 5.4 Future Work In the future, attempts will be made to correlate the particle trajectories to determine the activation volume directly. The influence of side groups and ring structures on the mobility and mechanical properties of the system could also be studied. Observations of mobility changes during unloading will also be made. 120 Bibliography [1] A.S. Argon and M.I. Bessonov. Plastic deformation in polyimides, with new implications on the theory of plastic defromation of glassy polymers. Phil. Mag., 35(4):917, 1977. [2] H.J.C. Berendsen, J.P.M. Postma, W.F. van Gunsteren, A. DiNola, and J.R. Haak. Molecular dynamics with coupling to an external bath. J. Chem. Phys., 81:3684, 1984. [3] J. Bergstrom. Large Strain Time-Dependent Behavior of Elastomeric Materials. PhD thesis, MIT, 1999. [4] R.K. Bharadwaj and R.H. Boyd. Effect of pressure on conformational dynamics in polyethylene: A molecular dynamics simulation study. Macro., 33:5897-5905, 2000. [5] R. H. Boyd, R. H. Gee, J. Han, and Y. Jin. Conformational dynamics in bulk polyethylene: A molecular dynamics simulation study. J. Chem. Phys., 101:788, 1994. [6] J. Brandrup and E.H. Immergut. Polymer Handbook, 3rd ed. Wiley-Interscience, 1989. [7] D. Brown and J.H. Clarke. Molecular dynamics simulation of an amorphous polymer under tension. 1. phenomenology. Macromolecules, 24:2075, 1991. [8] C. Chui. Application of Monte Carlo Techniques to the Study of Amorphous Polymer Deformation. PhD thesis, MIT, 1998. 121 [9] D. Deng, A.S. Argon, and S. Yip. Topological features of structural relaxations in a two-dimensional model atomic glass ii. Philos. Trans. R. Soc. London, A329:613, 1989. [10] E.R. Duering. Structure and relaxation of end-linked polymer networks. J. Chem. Phys., 101(1):8169-8192, 1994. [11] D. Frenkel and B. Smit. UnderstandingMolecular Simulation. Academic Press, 1996. [12] J.P. Gao and J.H. Weinger. Nature of stress on the atomic-level in dense polymer systems. Science, 266(5186):748, 1994. [13] R.H. Gee and R.H. Boyd. The role of torsional potential in relaxation dynamics: a molecular dynamics study of polyethylene. Comp. and Theor. Polymer Sience, 8(1):93, 1998. [14] M. Gerstein, J. Tsai, and M. Levitt. The volume of atoms on the protein surface: Calculated from simulation, using voronoi polyhedra. J. Mol. Biol., 249:955, 1995. [15] E.R. Griffin. Analysis of thermoplastics used in fold-and-form pipe liners. Website, 1997. [16] O.A. Hasan. An experimental and analytical investigation of the thermomechanical properties of Glassy polymers. PhD thesis, MIT, 1994. [17] R.N. Haward and R.J. Young, editors. The Physics of Glassy Polymers, chapter Relaxation processes and physical aging. Chapman and Hall, 1997. [18] R.W. Hocknet, S.P. Goel, and J.W. Eastwood. Quit high-resolution computer models of a plasma. J. Comput. Phys., 14:48, 1974. [19] W. Hoover. Canonical dynamics: Equilibrium phase-space distributions. Physical Review A, 31:1695, 1984. 122 [20] M. Hutnik, A.S. Argon, and U.W. Suter. Conformational characteristics of the polycarbonate of 4,4' - isopropylidenediphenol. Macromolecules, 24(22):5956- 5961, 1991. [21] K. Karatasos, D.B. Adolf, and S. Hotston. Effects of density on the local dynamics and conformational statistics of polyethylene: A molecular dynamics study. J. Chem. Phys., 112(19):8695, 2000. [22] A.J. Kinloch and R.J. Young. Fracture behavior of polymers. Applied Science Publishers, 1983. [23] T. Liang, Y. Yang, D. Guo, and X. Yang. Conformational transition behavior around glass transition temperature. J. Chem. Phys., 112(4):2016, 2000. [24] L.S. Loo, R.E. Cohen, and K.K. Gleason. Chain mobility in the amorphous region of nylon 6 observed under active uniaxial deformation. Science, 288:116, 2000. [25] K. Maeda and S. Takeuchi. Atomistic process of plastic-deformation in a model amorphous metal. Philos. Mag. A, 44:643, 1981. [26] N.G. McCrum, C.P. Buckley, and C.B. Bucknall. Principles of Polymer Engineering. Oxford University Press, 1988. [27] J.E. McKinney and M. Goldstein. J. Res. Nat. Bur. Std, 1974. [28] N. Metropolis, A. Rosenbluth, M. Rosenbluth, A. Tellar, and M. Tellar. Equation of state calculations by fast computing machines. J. Chem. Phys., 21:1087, 1953. [29] P.H. Mott, A.S. Argon, and U.W. Suter. Atomistic modeling of plastic- deformation of glassy-polymers. PhilosophicalMagazine A, 67(4):931-978, 1993. [30] S. Nose. A unified formulation of constant temperature molecular dynamics methods. J. Chem. Phys., 81(1):511, 1984. [31] P.V. Pant, G. D. Smith, and R.H. Boyd. A molecular dynamics simulation of polyethylene. J. Chem. Phys., 99(1):597, 1993. 123 [32] M. Parrinello and A. Rahman. Polymorphic transitions in single-crystals - a new molecular-dyanmics method. J. Appl. Phys., 52(12):7182, 1981. [33] W. Paul, G. D. Smith, and Y. D. Yoon. Static and dynamic properties of an n-clooh 202 melt from molecular dynamics simulations. Macromolecules, 30:77727780, 1997. [34] W. Paul and Y. D. Yoon. An optimized united atom model for simulations of polymethylene melts. J. Chem. Phys., 103(4):1702, 1995. [35] D. Rigby and R. Roe. Molecular dynamics simulation of polymerliquid and glass. i. glass transition. J. Chem. Phys., 87(12):7285, 1987. [36] D. Rigby and R.J. Roe. Molecular dynamics simulation of polymerliquid and glass. 3. chain conformation. Macromolecules, 22:2259, 1988. [37] M. Sprik. Effective pair potentials and beyond. In M.P. Allen and D.J. Tildesley, editors, Computer Simulation in Chemical Physics, volume 397 of Series C: Mathematical and Physical Sciences, pages 211-259. Kluwer Academic Publishers, 1993. [38] D. Srolovitz, V. Vitek, and T. Egami. An atomistic study of deformation of amorphous metals. Acta Metall., 31:335, 1983. [39] M.F. Sylvester. Molecular Dynamics Studies of the Liquid-Glass Transition in Atactic Poly(Propylene). PhD thesis, Massachusetts Institute of Technology, 1992. [40] H. Takeuchi and R. Roe. Molecular dynamics simulation of local chian motion in bulk amorphous polymers. ii. dynamics at glass transition. J. Chem. Phys., 94(11):7458, 1991. [41] D. N. Theodorou, T. D. Boone, L. R. Dodd, and K. F. Mansfield. Stress tensor in model polymer systems with periodic boundaries. Makromol. Chem., Theory Simul., 2:191-238, 1993. 124 [42] D.N. Theodorou. Local-structure and the response of a polymeric glass to elasticdeformation. Macro., 19(2):379, 1986. [43] L. Verlet. Computer experiments on classical fluids. i. thermodynamics properties of lennard-jones molecules. Phys. Rev., 159:98, 1998. [44] G.F. Voronoi. Z. Reine Angew. Math., 134:198, 1908. [45] I.M. Ward. Mechanical Properties of Solid Polymers. Wiley-Interscience, 1971. [46] L. Yang, D.J. Srolovitz, and A.F. Yee. Molecular dynamics study of isobaric and isochoric glass transitions in a model amorphous polymer. J. Chem. Phys., 110(14):7058, April 1999. [47] Q.Y. Zhou, A.S. Argon, and R.E. Cohen. Enhanced case-ii diffusion of diluents into glassy polymers undergoing plastic flow. Polymer, 42:613, 2001. 125