fiber lubrication: a molecular dynamics simulation study

182
ABSTRACT LIU, HONGYI. Fiber Lubrication: A Molecular Dynamics Simulation Study. (Under the direction of Wendy E. Krause and Orlando J. Rojas). Molecular and mesoscopic level description of friction and lubrication remains a challenge because of difficulties in the phenomenological understanding of to the behaviors of solid-liquid interfaces during sliding. Fortunately, there is the computational simulation approach opens an opportunity to predict and analyze interfacial phenomena, which were studied with molecular dynamics (MD) and mesoscopic dynamics (MesoDyn) simulations. Polypropylene (PP) and cellulose are two of most common polymers in textile fibers. Confined amorphous surface layers of PP and cellulose were built successfully with xenon crystals which were used to compact the polymers. The physical and surface properties of the PP and cellulose surface layers were investigated by MD simulations, including the density, cohesive energy, volumetric thermal expansion, and contact angle with water. The topology method was employed to predict the properties of poly(alkylene glycol) (PAG) diblock copolymers and Pluronic triblock copolymers used as lubricants on surfaces. Density, zero shear viscosity, shear module, cohesive energy and solubility parameter were predicted with each block copolymer. Molecular dynamics simulations were used to study the interaction energy per unit contact area of block copolymer melts with PP and cellulose surfaces. The interaction energy is defined as the ratio of interfacial interaction energy to the contact area. Both poly(proplene oxide) (PPO) and poly(ethylene oxide) (PEO) segments provided a lipophilic character to both PP and cellulose surfaces. The PPO/PEO ratio and the molecular weight were found to impact the interaction energy on both PP and cellulose surfaces. In aqueous solutions, the interaction energy is complicated due to the presence of

Upload: others

Post on 10-Nov-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Fiber Lubrication: A Molecular Dynamics Simulation Study

ABSTRACT

LIU, HONGYI. Fiber Lubrication: A Molecular Dynamics Simulation Study. (Under the direction of Wendy E. Krause and Orlando J. Rojas).

Molecular and mesoscopic level description of friction and lubrication remains a

challenge because of difficulties in the phenomenological understanding of to the behaviors

of solid-liquid interfaces during sliding. Fortunately, there is the computational simulation

approach opens an opportunity to predict and analyze interfacial phenomena, which were

studied with molecular dynamics (MD) and mesoscopic dynamics (MesoDyn) simulations.

Polypropylene (PP) and cellulose are two of most common polymers in textile fibers.

Confined amorphous surface layers of PP and cellulose were built successfully with xenon

crystals which were used to compact the polymers. The physical and surface properties of the

PP and cellulose surface layers were investigated by MD simulations, including the density,

cohesive energy, volumetric thermal expansion, and contact angle with water.

The topology method was employed to predict the properties of poly(alkylene glycol)

(PAG) diblock copolymers and Pluronic triblock copolymers used as lubricants on surfaces.

Density, zero shear viscosity, shear module, cohesive energy and solubility parameter were

predicted with each block copolymer.

Molecular dynamics simulations were used to study the interaction energy per unit

contact area of block copolymer melts with PP and cellulose surfaces. The interaction energy

is defined as the ratio of interfacial interaction energy to the contact area. Both poly(proplene

oxide) (PPO) and poly(ethylene oxide) (PEO) segments provided a lipophilic character to

both PP and cellulose surfaces. The PPO/PEO ratio and the molecular weight were found to

impact the interaction energy on both PP and cellulose surfaces.

In aqueous solutions, the interaction energy is complicated due to the presence of

Page 2: Fiber Lubrication: A Molecular Dynamics Simulation Study

water and the cross interactions between the multiple molecular components. The polymer-

water-surface (PWS) calculation method was proposed to calculate such complex systems. In

a contrast with a vacuum condition, the presence of water increases the attractive interaction

energy of the diblock copolymer on the cellulose surface, compared with that on the PP

surface. Water decreases the interaction energy of the triblock copolymer on the cellulose

surface, compared with that on the PP surface.

MesoDyn was adopted to investigate the self-assembled morphology of the triblock

copolymer, in aqueous solution, confined and sheared at solid-liquid interfaces. In a bulk

aqueous solution, when the polymer concentration reached 10% v/v, micelles were observed

with PPO blocks in the core and PEO blocks in the shell of the micelles. At the

concentrations of 25% and 50%, worm-like micelles and irregular cylinders were observed in

solutions, respectively. The micelles were formed faster in aqueous solutions confined by

cellulose surfaces than that in the bulk. The formed micelles were broken under shearing,

which led to a depletion of polymers at the interfaces. During the shearing on the PP surfaces,

the polymers were adsorbed on the surfaces protecting the PP surfaces.

This simulation study in the fiber lubrication was in good agreement with the

experimental results and so provided an approach to visualize the polymer configuration at

the liquid-solid interface, predict the lubricant-surface systems, and theoretically guide the

experiments of designing new/efficient lubricants for fibers.

Page 3: Fiber Lubrication: A Molecular Dynamics Simulation Study

Fiber Lubrication: A Molecular Dynamics Simulation Study

by Hongyi Liu

A dissertation submitted to the Graduate Faculty of North Carolina State University

in partial fulfillment of the requirements for the degree of

Doctor of Philosophy

Fiber and Polymer Science

Raleigh, North Carolina

2009

APPROVED BY:

_______________________________ ______________________________ Dr. Melissa A. Pasquinelli Dr. Harvey Charlton _______________________________ ______________________________ Dr. Orlando J. Rojas Dr. Wendy E. Krause Co-chair of Advisory Committee Chair of Advisory Committee

Page 4: Fiber Lubrication: A Molecular Dynamics Simulation Study

DEDICATION

To

my wife, my lover, my helper,

Yan Vivian Li,

for her marriage with me.

And also to

my parents,

Qitong Liu and Sulan Song,

who made all of this possible,

for their feeding, teaching, love and patience.

When I consider your heavens,

the work of the your fingers,

the moon and the stars,

which you have set in place,

what is man that you are mindful of him,

the son of man that you care for him?

--------Psalms 8:3-4

ii

Page 5: Fiber Lubrication: A Molecular Dynamics Simulation Study

BIOGRAPHY

Hongyi Liu was born in Hua County, Henan, P.R. China. He enjoyed the happy time

there, and then he went to college in 1993. After a four-year study, he graduated from Wuhan

University of Science and Engineering with his B.E. degree in Textile Engineering. By virtue

of his professional knowledge, he had worked in Zhongyuan University of Technology since

1997. In 2000, he went to Donghua University as a graduate student in Fiber Science and

Textile Design. After three years, he joined Zhongyuan University of Technology again with

his Master degree and held a position of senior lecturer. In 2005, he came to the U.S. and

started working towards his Ph.D. degree in Fiber and Polymer Science, College of Textile,

North Carolina State University. His research field is molecular dynamics simulation of thin

films and the boundary lubrication system. He got married with Yan Vivian Li in May 2008.

iii

Page 6: Fiber Lubrication: A Molecular Dynamics Simulation Study

ACKNOWLEDGEMENTS

The author would like to offer my sincerest thanks to my committee members: to my

co-chairs, Dr. Wendy E. Krause and Dr. Orlando J. Rojas, for their wise counsel, direction

and support; to Dr. Melissa A. Pasquinelli for her time and help in principles of molecular

dynamics simulation; and Dr. Harvey Charlton as the minor representative. And thanks to Dr.

Brenner as the graduate school representative. Also, the author deeply appreciates Dr. Juan P.

Hinestroza (Cornell University) who firstly provided the opportunity of this Ph.D. study,

supported and helped the author to grow in many aspects.

The author also acknowledges support of the National Textile Center under grant

C05-NS09 from 2005-2008.

Much cooperation helps provided by Dr. Yan Vivian Li (Cornell University), Dr.

Junlong Song, Dr. Jing Liang and Dr. J. L Trochmann (UNICAMP, Brazil) are also

appreciated. Dr. Senli Guo (Duke University) and Dr. Hujun Shen (Pennsylvania State

University) are acknowledged for giving valuable advice and warm discussion. The author

also thanks Carolyn Krystoff and Angie Brantley for their help on administrative matters at

NCSU. And Rebecca Klossner is appreciated for her help.

Great appreciation goes to Mark James Haven for his technical support with hardware

and Dr. Alex Goldberg (Accelrys Software Inc.) with software.

iv

Page 7: Fiber Lubrication: A Molecular Dynamics Simulation Study

TABLE OF CONTENTS

LIST OF TABLES ··········································································································· viii

LIST OF FIGURES ········································································································· ix

ABBREVIATIONS ········································································································· xvi

Chapter 1 Introduction ··································································································· 1

1. A history of the study of friction ······························································· 2

2. Problem definition ····················································································· 6

3. Simulation background ············································································· 10

4. Summary ··································································································· 25

References ····································································································· 28

Chapter 2 Molecular Modeling of Amorphous Fiber Surfaces ····································· 32

1. Introduction ······························································································· 32

2. Simulation method ···················································································· 36

3. Results and discussion ·············································································· 46

4. Conclusion ································································································ 56

References ····································································································· 57

Chapter 3 The Structure-property Correlations of PAG Block Copolymer Properties Predicted by Topological Method ································································· 60

1. Introduction ······························································································· 60

2. Simulation method and modeling ····························································· 63

3. Results and discussion ·············································································· 69

v

Page 8: Fiber Lubrication: A Molecular Dynamics Simulation Study

4. Conclusion ································································································ 78

5. Supplement ································································································ 79

References ····································································································· 80

Chapter 4 Interaction Energies of PAG Block Copolymer Melts with Polymeric Surfaces in Vacuum ······················································································· 82

1. Introduction ······························································································· 82

2. Interaction energy and simulation method ················································ 84

3. Results and discussion ·············································································· 88

4. Summary ··································································································· 93

References ····································································································· 94

Chapter 5 Configuration and Interaction Energies of a Single PAG Copolymer Molecule in Aqueous Solutions on Fiber Surfaces ······································· 95

1. Introduction ······························································································· 95

2. Modeling and methods ·············································································· 99

3. Results and discussion ·············································································· 108

4. Conclusion ································································································ 114

References ·····································································································123

Chapter 6 Mesoscale Simulations of PAG Triblock Copolymer Morphology ·············· 125

1. Introduction ·······························································································125

2. Simulation parameters ···············································································131

3. Results and discussion ·············································································· 133

4. Conclusion ································································································ 148

vi

Page 9: Fiber Lubrication: A Molecular Dynamics Simulation Study

References ·····································································································150

Chapter 7 Future Work ·································································································· 152

APPENDICES ················································································································ 155

Appendix 1. The perl script to calculate the interaction energy of two layers························································································································· 155

Appendix 2. The perl script to calculate the interaction energies of PWS system ·············································································································159

vii

Page 10: Fiber Lubrication: A Molecular Dynamics Simulation Study

LIST OF TABLES

Table 1.1 List of block copolymers used in our project ·················································· 10

Table 2.1 CED and solubility parameters of Cellulose and PP layers ····························· 48

Table 2.2 PP and cellulose solubility parameters ···························································· 49

Table 3.1 The δ and δV values used in the calculation of PAG polymers ························ 64

Table 3.2 The list of PAG diblock copolymers equivalent to UCON® products form DOW ················································································································ 68

Table 3.3 The list of PAG tri-block copolymers equivalent to Pluronic® products formBASF················································································································ 69

Table 4.1 PAG block copolymers with parameters to build the confined amorphous cells··················································································································· 87

Table 5.1 The interaction energies between the polymer and the surface affected by water ················································································································113

Table 6.1 Topology parameters for MesoDyn ································································128

Table 6.2 Interaction parameters for MesoDyn ·······························································130

viii

Page 11: Fiber Lubrication: A Molecular Dynamics Simulation Study

LIST OF FIGURES

Figure 1.1 The sketch of Amontons’ friction law. N represents the normal force and Ff denotes the friction. ·················································································· 2

Figure 1.2 The sketch of (a) Coulomb surface roughness model and (b) Vince surface interaction model. N indicates the normal force. v refers to the velocity of moving surfaces. Δh denotes the height of an asperity. ············· 4

Figure 1.3 Boundary lubrication and hydrodynamic lubrication. The upper figure is provided from project collaborator Yan Li, Cornell University. The lower image is drawn by Song. ·············································································· 8

Figure 1.4 The structures of the polymers as lubricants: (a) Diblock copolymers equivalent to UCON® polymers of DOW, (b) Triblock copolymers equivalent to Pluronic® polymers of BASF. ··············································· 9

Figure 1.5 Simulation methods with the time and length scale. This figure is revised from the original source of Nielsen. ···························································· 11

Figure 1.6 The 2-D sketch of periodic boundary conditions. The blue lattice is a simulation model of an atom system. The white lattices are the mirror images of the simulation model. The circled numbers refer to the atoms in the model lattice and mirror images. The red unidirectional arrows denote the motion trajectory of all #2 atoms. The dash circles represent the target positions of all #2 atoms. The yellow, green and purple bidirectional arrows indicate the interaction between #3 and #1 atom. The atom distance of the purple arrows is shortest. ····················································· 18

Figure 1.7 Comparison of the moleculare models in the mesoscale simulation (left) and the MD simulation (right) with the example of dimyristoylphosphatidylcholine (DMPC) which is coarse grained using 13 beads to represent the 118 atoms. This picture was drawn by Nielsen et al. ················································································································· 23

ix

Page 12: Fiber Lubrication: A Molecular Dynamics Simulation Study

Figure 2.1 Repeat units for building polymer (fiber) substrates (a) 1,4-βD glucose, (b) propylene. (Atom colors: carbon atoms—grey, hydrogen—white, and oxygen—red; the head atom is circled by cyan, and the tail atom is circled by magenta). ····················································································· 37

Figure 2.2 Polymer molecules: (a) Cellulose, and (b) Polypropylene (PP). The end-to-end distance of the Cellulose molecule with 8 repeat units is 42.5 Å. The end-to-end distance of the PP molecule with 30 repeat units is 76.7 Å. Atom colors are set the same as Figure 2.1. ···································· 37

Figure 2.3 The procedures of building PP confined layer. (a) The PP original confined layer: 16.7×16.7×27.6 Å. (b) The original xenon-PP-xenon cell: 16.7×16.7×90.6 Å. (c) The xenon-PP-xenon cell after the dynamics procedure. (d) The PP confined unit layer: 16.7×16.7×26.0 Å. Xenon atoms are colored by cyan. Other atom colors are set the same as Figure 2.1. ··············································································································· 39

Figure 2.4 The procedures of building cellulose confined layer. (a) The cellulose original confined layer: 23.6×23.6×46.3 Å. (b) The original xenon-cellulose-xenon cell: 23.6×23.6×63.0 Å. (c) The xenon-cellulose-xenon cell after the dynamics procedure: 16.7×16.7×44.4Å. (d) The cellulose confined unit layer: 16.7×16.7×31.5 Å. Xenon atoms are colored by cyan. Other atom colors are set the same as Figure 2.1. ······················································································································ 40

Figure 2.5 The PP and cellulose confined amorphous layers. (a) The PP super layer (4×unit layers): 33.3×33.3×26.0 Å. (b) The cellulose super layer (4×unit layers): 33.3×33.3×32.3 Å. (c) The PP super layer (8×unit layers): 66.6×66.6×26.0 Å. (d) The cellulose super layer (8×unit layers): 66.6×66.6×32.3 Å. ······················································································· 41

Figure 2.6 The steps of building the round water droplet. There are 500 water molecules involved. (a) The original periodic water cube: 24.6×24.6×24.6Å. (b) The water cube without the periodic lattice. (c) The round water nano-droplet. Its diameter is approximately 32 Å. ······································ 42

Figure 2.7 The contact angle models. (a) A water nano-droplet on the PP surface. (b) A water nano-droplet on the cellulose surface. ··········································· 43

x

Page 13: Fiber Lubrication: A Molecular Dynamics Simulation Study

Figure 2.8 The weight distributions of PP and cellulose confined amorphous layers. · 47

Figure 2.9 Volumetric thermal expansions of PP and cellulose amorphous layers. ······ 51

Figure 2.10 The dynamic process of contact angle models with PP and cellulose surfaces. ······································································································· 52

Figure 2.11 The final distribution sketches of water molecules in the lattice. (a) CA model of PP. (b) CA model of cellulose. The final frames are used. The dash lines are the solid surface positions of PP and cellulose respectively. ····················································································································· 54

Figure 2.12 The final shape of water nano-droplet on the PP and cellulose surfaces. The final frames are used. The dash lines are the solid surface positions of PP and cellulose respectively. The black solid lines are the tangent line of the shape. ····································································································· 55

Figure 3.1 Chemical structures of (a) UCON® and (b) Pluronic® polymers. (R refers to the butyl group) ············································································· 61

Figure 3.2 The hydrogen-suppressed graph of a molecular repeat unit, using PEO as an example. (a) Valence bond structure of a PEO repeat unit. (b) Hydrogen-suppressed graph of a PEO repeat unit. ······································ 64

Figure 3.3 Illustrations calculating the simple connectivity index (δ) and the valence connectivity index (δV). (a) The δ values are shown at the vertices and the β values are shown along the edges. (b) The δV values are shown at the vertices and the βV values are shown along the edges. ······························· 65

Figure 3.4 Repeat units to build PAG block copolymers: (a) butylene, (b) oxyethylene (EO), (c) oxypropylene (PO). (Atom colors: carbon atoms—grey, hydrogen—white, and oxygen—red; the head atom is circled by cyan, and the tail atom is circled by magenta). ··························· 67

Figure 3.5 The densities of PAG block copolymers with the experimental values and the simulation values calculated by the topological method for structure-property correlations. ···································································· 71

Figure 3.6 The zero-shear viscosity of PAG block copolymers, calculated by the topological method for structure-property correlations at 298 K. ··············· 73

xi

Page 14: Fiber Lubrication: A Molecular Dynamics Simulation Study

Figure 3.7 The shear modulus of PAG block copolymers, calculated by the topological method for structure-property correlations at 298 K. ··············· 75

Figure 3.8 The cohesive energies of PAG block copolymers, calculated by the topological method for structure-property correlations at 298 K. ··············· 76

Figure 3.9 The solubility parameters of PAG block copolymers, calculated by the topological method for structure-property correlations at 298 K. ··············· 77

Figure 4.1 The calculation of interaction energy for two components. (a) The original model of two-component layer with the vacuum space. The top layer is the liquid polymers (RP6E8) and the bottom layer is the solid surface (PP). (b) The frame after dynamics processes. (c) The surface layer after the removal of the liquid polymers. (d) The liquid layer after the removal of the bottom layer. Etotal, Esurface and Epolymer are the total potential energy of (b), (c) and (d). All are calculated by the force field of Materials Studio 4.1. ···················································································· 85

Figure 4.2 The interaction energies of melt RPnEm polymers with 50% hydrophobic weight ratio and different molecular weights on the PP and cellulose surfaces. ······································································································· 88

Figure 4.3 The interaction energies of melt RPnEm polymers with variable hydrophobic weight ratios and the similar molecular weights on the PP and cellulose surfaces. ················································································· 91

Figure 4.4 The interaction energies of melt EmPnEm polymers with 50% hydrophobic weight ratios and the similar molecular weights on the PP and cellulose surfaces. ················································································· 92

Figure 5.1 The hypothesis of the EO-PO-EO block copolymer molecule on a PP surface. This figure is revised from the work of Li et al. ···························· 97

xii

Page 15: Fiber Lubrication: A Molecular Dynamics Simulation Study

Figure 5.2 Models of a single PAG molecule affected by water on PP and cellulose surfaces. Colors represents particular: carbon atoms—grey, hydrogen—white, and oxygen—red. The PAG polymers are colored with different segments: butyl group—purple, PPO— green, and PEO—blue. (a) RP6E8-water-PP system, (b) RP6E8-water-cellulose system, (c) E19P29E19-water-PP system, and (d) E19P29E19-water-cellulose system. ········································································································· 101

Figure 5.3 PWS sketch method to calculate the cross terms of interaction energy in the polymer-water-surface system. ······························································ 105

Figure 5.4 The potential energy of each single component. The white plate represents the removal of a component from the PWS system. Only one component is left and its potential energy can be calculated by the force field. ············································································································· 105

Figure 5.5 The potential energy of each two-component sub system. The white plate represents the removal of a component from the PWS system. ·················· 106

Figure 5.6 The sketch of the interaction energy (EPS_interaction) between the polymer and the surface affected by water. ································································ 107

Figure 5.7 The configurations of the RP6E8 and E19P29E19 molecules in vacuum on the PP and cellulose surfaces. (a) RP6E8 on the PP surface, (b) RP6E8 on the cellulose surface, (c) E19P29E19 on the PP surface, (d) E19P29E19 on the cellulose surface. All images are the final frames of the NVT dynamics simulations. The color sets are the same with Figure 5.2. ··············································································································· 115

Figure 5.8 The configurations of the RP6E8 and E19P29E19 molecules in water on the PP and cellulose surfaces. (a) RP6E8 on the PP surface, (b) RP6E8 on the cellulose surface, (c) E19P29E19 on the PP surface, (d) E19P29E19 on the cellulose surface. All images are the final frames of the NVT dynamics simulations. In each top view of the images, the water molecules are invisible for clarity to the PAG polymer molecules. The color sets are the same with Figure 5.2. ······················································· 119

xiii

Page 16: Fiber Lubrication: A Molecular Dynamics Simulation Study

Figure 6.1 Phase morphologies of aqueous E19P29E19 solutions under the bulk condition. Three concentrations: (a series) 10% v/v. (b series) 25% v/v. (c series) 50% v/v. Phase colors: Water—red, PEO—blue, PPO—green. All are the final frames after 20000 time steps. ················································ 134

Figure 6.2 Time evolution of morphology for the aqueous 10% v/v E19P29E19 solution. The time steps from the upper left to the lower right are: 100, 1600, 3000, 4500, 5000, 5700, 10000, 15000. Only PPO phases are displayed for clarity. ···················································································· 137

Figure 6.3 The Order parameter with simulation time steps for 10% E19P29E19. The dash lines are at x=1600 and x=5700. ·················································· 138

Figure 6.4 The order parameter with simulation time steps for 25% E19P29E19. The dash line is at x=750. ··················································································· 138

Figure 6.5 The order parameter with simulation time steps for 50% E19P29E19. The dash line is at x=700. ··················································································· 139

Figure 6.6 Time evolution of morphology for the aqueous 10% v/v E19P29E19 solution confined by cellulose surfaces. The time steps from the upper left to the lower right are: 100, 1600, 2000, 2500, 3900, 5000, 10000, 20000. Only PPO phase is displayed for clarity. ····················································· 140

Figure 6.7 Order parameter with simulation time steps for 10% E19P29E19 solution confined by cellulose surfaces. The dash lines are at x=1600 and x=3900. ········································································································ 141

Figure 6.8 Time evolution of morphology for the aqueous 10% v/v E19P29E19 solution confined by PP surfaces. The time steps from the upper left to the lower right are: 100, 300, 500, 1000, 1500, 2000, 3000, 20000. Water phase is removed for clarity. PPO—green and PEO—blue. ························ 141

Figure 6.9 Order parameter with simulation time steps for 10% E19P29E19 solution confined by PP surfaces. ·············································································· 142

xiv

Page 17: Fiber Lubrication: A Molecular Dynamics Simulation Study

Figure 6.10 The morphology for the 10% v/v E19P29E19 solution sheared by cellulose surfaces. The time steps from the upper left to the lower right are: 100, 1000, 2000, 2500, 3000, 3500, 4000, and 5000. Water phase is removed for clarity. PPO—green and PEO—blue. The yellow arrows denote the shearing directions of surfaces. ·················································· 145

Figure 6.11 Order parameter with simulation time steps for 10% E19P29E19 solution sheared by cellulose surfaces. ······································································ 145

Figure 6.12 Compressibility energy contribution with the time steps for 10% v/v E19P29E19 solution sheared by cellulose surfaces. ···································· 146

Figure 6.13 The morphology for the 10% v/v E19P29E19 solution sheared by PP surfaces. The time steps from the upper left to the lower right are: 100, 1000, and 5000. Water phase is removed for clarity. PPO—green and PEO—blue. The yellow arrows denote the shearing directions of PP surfaces. ······································································································· 147

Figure 6.14 Order parameter with simulation time steps for 10% E19P29E19 solution sheared by PP surfaces. ················································································ 148

xv

Page 18: Fiber Lubrication: A Molecular Dynamics Simulation Study

ABBREVIATIONS

AFM atomic force microscope

AMBER assisted model building with energy minimization

AMU atomic mass unit

CA contact angle

CEC compressibility energy contribution

CED cohesive energy density

CFF consistent force field

CHARMM chemistry at Harvard macromolecular mechanics

COMPASS condensed-phase optimized molecular potentials for atomistic simulation studies

CVFF consistent valence force field

DMA dynamic mechanical analysis

DPD dissipative particle dynamics

EO oxyethylene

EVA polyethylene-vinyl acetate

HEC (hydroxyethyl) cellulose

HPC (hydroxypropy1) cellulose

LFM lateral force microscope

LJ Lennard-Jones

MD molecular dynamics

MesoDyn mesoscopic dynamics

MM molecular mechanics

MW molecular weight

NPH constant atom number, pressure, enthalpy

NPT constant atom number, pressure, temperature

NVE constant atom number, energy, volume

xvi

Page 19: Fiber Lubrication: A Molecular Dynamics Simulation Study

NVT constant atom number, temperature, volume

PA6 polyamide 6

PAG poly(alkylene glycol)

PCFF polymeric consistent force field

PE polyethylene

PEO poly(ethylene oxide)

PET polyethylene tephthalate

PMMA poly(methyl methacrylate)

PO oxypropylene

PP polyproplene

PPO poly(proplene oxide)

PVAC poly(vinyl acetate)

PVC poly-vinyl chloride

PVOH poly(vinyl alcohol)

PWS Polymer-Water-Surface

QCM quartz crystal microbalance

SFA surface force apparatus

SFM scanning force microscope

vdW van der Waals

xvii

Page 20: Fiber Lubrication: A Molecular Dynamics Simulation Study

Chapter 1

Introduction

Abstract The science of friction and lubrication is called tribology. It is attractive to both industrial and academic research. The study scale on this issue varies from macro scale of Amontons’ friction law to micro scale. In the past decades, great interest in the area has been further focused to the nanoscale, with a term called nanotribology. A great deal of experimental and theoretical work has been done at micro and nanoscales with modern instruments. However, atomic details of friction and lubrication are still a mystery because it is technically difficult to observe what happens at a solid-liquid interface during sliding with experimental tools. Molecular dynamics (MD) and microscopic dynamics (MesoDyn) simulation methods are used to study the lubrication of fibers from nanoscale to mesoscale. Following chapters will discuss the use of these methods to study several problems in fiber lubrication.

It has been well-known that friction behaviors have both advantages and

disadvantages in textile processing. When short natural fibers, such as wool, flax and cotton,

are spun into yarns, friction plays a favorable factor to the formation of yarns. If the friction

is too low, the yarn strength decreases and the dimensional stability of fabrics would be

consequently reduced. In addition, friction plays a main role in making roving, usually by

controlling the humidity of cotton or adding fluids on wool and synthetic fibers in roller

drafting. Friction is not desirable when yarns are woven into fabrics. In the past, people

learned to coat yarns with wax or oil to reduce friction. Nowadays, with the industrial trend

of high speeds, friction phenomena have been given more attention in both the textile

1

Page 21: Fiber Lubrication: A Molecular Dynamics Simulation Study

industry and academic research.1

1. A history of the study of friction

The study of friction is based on two classical laws, proposed by da Vinci

(1452-1519), but rediscovered by Amontons in 1699.2, 3 The laws, depicted in Figure 1.1,

state that: (1) the frictional force Ff is proportional to the normal force, N, and the coefficient

of friction, μ=Ff/N, is a constant in a given friction system; (2) the friction is independent of

the area of the contact between two given surfaces. These friction laws were verified by

Coulomb in 1781.2, 3 He pointed out the difference between static friction and kinetic friction:

the static friction force is the force resisting sliding that needs to be conquered in order to

trigger the motion of sliding. He observed that the kinetic friction is independent of the

sliding speed. This discovery is called the third law of friction.2

Ff N

Figure 1.1 The sketch of Amontons’ friction law. N represents the normal

force and Ff denotes the friction.

2

Page 22: Fiber Lubrication: A Molecular Dynamics Simulation Study

Many other theories of friction have been developed. They fall into two main

categories: surface roughness and surface interaction.3 Coulomb claimed that any friction

essentially results in the process of lifting the asperities over the counterparts on the other

surface and that lifting consumes energy. He accordingly proposed that a lubricant could

possibly reduce friction by filling up the gaps between the asperities. During the nineteenth

century, many researchers were also able to repeat Coulomb’s experimental results and

accepted his roughness theory of friction.2 However, there were two drawbacks in the theory.

One was the difficulty to explain how the frictional energy can be produced and dissipated.

When the upper asperities move upward and stride over the counterparts of the surface,

shown in Figure 1.2(a), if a surface is completely free of adhesion interactions, the potential

energy can be measured by E=N·Δh. In the sliding, it is not all of the dissipated energy

converted to heat, which is a drawback of this theory. The other drawback was the difficulty

to determine roughness values before the atomic force microscope (AFM) was invented

which measures roughness at nanoscale.4

3

Page 23: Fiber Lubrication: A Molecular Dynamics Simulation Study

(a) (b)

Figure 1.2 The sketch of (a) Coulomb surface roughness model and (b) Vince surface interaction model. N indicates the normal force. v refers to the velocity of moving surfaces. Δh denotes the height of an asperity.

Vince modified Coulomb’s theory in 1785 by stating that adhesion or cohesion

between the surfaces largely contributed to friction, depicted in Figure 1.2(b).2, 5 The effects

of both adhesion and roughness were observed on some specific surfaces, but only the

adhesion was shown to be the main factor that contributed to frictional force. This

observation could explain that the stationary coefficient of friction is higher than the sliding

coefficient. The molecular adhesion present between the surfaces is caused by short-range

intermolecular forces.2, 5 When surfaces are sliding past one another, the surface molecules

are removed from their equilibrium positions. When that displacement exceeds a certain

distance and then the molecules recover by vibrational motion to their equilibrium positions.

4

Page 24: Fiber Lubrication: A Molecular Dynamics Simulation Study

Vince’s theory can not explain the wear and the material transitions, such as wear, which are

generally observed when naked surfaces slide past one another. In 1936, Bowden, Tabor and

their co-workers suggested that the predominant friction effect typically is a strong adhesion

of two surfaces at real contact points as a result of cold welding.2 When the surfaces are

sliding on a surface, the junctions of real contact points must shear or be sheared. They

assumed the friction force is the primary force required to shear junctions.2 However, their

theory could not explain that substantial friction exists even in cases without wear.

No significant progress was made after Bowden and Tabor’s theory until the 1970’s.

Israelachvili, a student of Tabor, developed a surface force apparatus (SFA), which can

measure friction at an atomic scale. With SFA, he observed the first direct evidence for

wear-free friction.6 In SFA, two cylindrical smooth mica surfaces are positioned in a crossed

cylinder configuration. The two surfaces are brought toward or away from each other using

several mechanical stages to increase sensitivity. An optical technique using multiple beam

interference fringes is used to observe and monitor the contact region of the two surfaces.

Deflections of the sensor connected to the half-cylinders are proportional to frictional forces.

With SFA, Israelachvili et al.6-8 studied surfactant mono-layers and discovered a relationship

between adhesion and friction. Surfactant mono-layers were found to exhibit three states:

solid-like, amorphous, or liquid-like. Moreover, the states of the layers can be changed by

varying the atmosphere, temperature, velocity or related parameters. However, Krim9 pointed

5

Page 25: Fiber Lubrication: A Molecular Dynamics Simulation Study

out that these experiments could not address the explicit physical mechanism that gives rise

to the friction that they were measuring, thus she proposed a new field in friction study.

The new field is called nanotribology, the term firstly coined by Krim10 in 1991. With

the inventation of new techniques that measure surface topography, adhesion, friction, wear,

lubrication, film thickness, and mechanical properties, tribology now can be studied on a

1-1000 nanometer (nm) scale. Nanotribology is concerned with experimental and theoretical

research ranging from atomic and molecular scale to micro-scale. In macrotribology, the

studies are conducted on surfaces with relatively large mass under heavily loaded condition.

Wear is inevitable and the bulk properties of components dominate the fiction phenomena. In

nanotribology, the friction (or shear) systems have light loading forces, where wear is

negligible and the properties of the interface dominate the friction phenomena.11

Besides SFA mentioned above, there are two other important experimental techniques

in nanotribology: lateral force microscope (LFM)12-17 and quartz crystal microbalance (QCM)

18-22. These methods are described in Chapter 2, 5 and 6.

2. Problem definition

Discussions on friction and lubrication have captured the attention of engineers and

scientists for over several hundred years. The main reason that so little is known about either

interaction or friction processes is that we cannot see what happens at the interface during

6

Page 26: Fiber Lubrication: A Molecular Dynamics Simulation Study

sliding. As described above, some instruments such as LFM have been used to perform

friction measurements, characterize contact conditions, and even examine the worn surface.

After the experimental developments, the modeling in nanotribology at the atomic level,

particularly molecular dynamics (MD) simulations, has brought scientists a step closer to

understanding what takes place during sliding contact at an atomic scale.

As sliding conditions become more severe such as with rough surfaces and high

contact pressures, wear becomes a severe problem in the system. Since the fiber-to-fiber or

fiber-to-metal pressure is high, the surfaces contact is inevitable and the boundary lubrication

occurs.23-25 As shown in Figure 1.3, boundary lubrication is a regime in which the lubricant

film becomes too thin to provide total separation. This may be due to excessive loading,

speeds, or a change in the characteristics of fluids. In such a situation, contact between

surface asperities (peaks and valleys) occurs. Friction reduction and wear protection is then

provided through the interfaces of lubricant molecules adhered to surfaces rather than the

bulk properties of the lubricating fluids.

7

Page 27: Fiber Lubrication: A Molecular Dynamics Simulation Study

Figure 1.3 Boundary lubrication and hydrodynamic lubrication. The upper

On rough surfaces, it can be difficult to identify the underlying mechanisms

associated with complex phenomena. The dynamics of each individual asperity in rough

surfaces may be associated with the quantity of contact points, which is time-dependent.

These studies are thus complementary to those on the elasticity, plasticity and rheology of

individual asperities. As for fibers whose surfaces are rough, the deformation of the asperities

is elastic with proper lubrication (see Figure 1.3), which can reduce friction heat as a result of

a low temperature with proper lubricants.

figure is provided from project collaborator Yan Li,12 Cornell University. The lower image is drawn by Song.26

8

Page 28: Fiber Lubrication: A Molecular Dynamics Simulation Study

A lubricant, which is applied on a surface, might create an ultra-thin coating layer

over the surfaces. The ultra-thin lubricant layer strongly adheres to the surface. The film is

often only one or two molecules thick but can provide enough protection to prevent fibers

from being damaged.

Figure 1.4 The structures of the polymers as lubricants: (a) Diblock

copolymers equivalent to UCON® polymers from DOW, (b)

Surfactants are comm . Block

copolym

(a)

(b)

Triblock copolymers equivalent to Pluronic® polymers from BASF.

only used as lubricant due to their amphiphilicity

ers composed of poly(ethylene oxide) (PEO) and poly (propylene oxide) (PPO)

blocks are common nonionic surfactants, which are popular in drug delivery and used in

washing powders and personal care products such as tooth paste. The polymers used in this

work are diblock copolymers (Butyl-POn-EOm) and triblock copolymers (EOm-POn-EOm),

shown in Figure 1.4. PEO is soluble to water while PPO and butylene oxide are insoluble to

water.27 The variations of those block copolymers that we studied in this work is given in

9

Page 29: Fiber Lubrication: A Molecular Dynamics Simulation Study

Table1.1.

Table 1.1 List of block copolymers used in this work.

R represents to the butyl group, P denotes PO, and E refers to EO in our codes here.

in our work name Weight

Density

g/cm3

Code Commercial Molecular at Manufacturer

298K

R 50- 70 P6E8 HB-01 750 1.031 DOW RP10E13 50-HB-0400 1230 DOW 1.041RP13E17 1 50-HB-0660 590 1.051 DOW RP22E29 50-HB-2000 2660 1.056 DOW RP33E44 50-HB-5100 3930 1.056 DOW RP11E42 75-H-1400 2470 1.095 DOW RP12 LB-0165 740 0.983 DOW RP24 LB-0525 1420 0.994 DOW RP40 LB-1715 2490 1.00 DOW E76P29E76 F068 8400 1.06 BASF E103P39E103 F088 11400 1.06 BASF E133P50E133 F108 1 4600 1.06 BASF E19P29E19 P065 3400 1.06 BASF E26P40E26 P085 4600 1.04 BASF E37P56E37 P105 6500 1.05 BASF

. Simulation background

s have been adapted in many research areas of the chemical,

physica

3

Computer simulation

l, material, biological sciences. The rapid promotion in using computer simulations is

primarily due to an enormous recent increase in computer power, effective algorithms, and

successful development of software for modeling a wide range of systems and processes.

10

Page 30: Fiber Lubrication: A Molecular Dynamics Simulation Study

Figure 1.5 depicts the multiple scales available for simulations, and the detecting resolution

at each scale.

Figure 1.5 Simulation methods with the time and length scale. This figure is

redrawn from the original source of Nielsen et al.28

In the current work, to the aim

esoscale, particlularly adhesion, diffusion, morphology at nanoscale and mesoscale. In

general, we assume that there are no electron and chemical reactions involved in fiber

lubrication phenomena, so we do not require quantum mechanics. The scale of our previous

experimental work varies from 1 nm (10 Å) to 1 μm, and the time scale also spans from

picosecond (ps) to microsecond (µs). Therefore, we use molecular dynamics and mesoscale

is to study fiber lubrication phenomena at a nanoscale

and m

Femtosecond (fs)

Picosecond (ps)

Nanosecond (ns)

Microsecond (μs)

Second (s)

hour (hr)

Å nm μm mm m

day (d) QCM, LFM, Nanoindenter

Quantum mechanics

Molecular dynamics

Mesoscale simulation

Finite element analysis

Engineering design (Unit process design)

T

Length Scale

ime Scale

Morphology, interface behavior etc.

Equilibrium structure, interaction energy, etc.

11

Page 31: Fiber Lubrication: A Molecular Dynamics Simulation Study

simulation in this work.

3.1 Molecular dynamics (MD) simulations

on has been intensively developed since 1970s.

MD sim

Molecular dynamics (MD) simulati

ulations are used to track the motions of atoms and molecules as a function of time.

The classical MD method numerically solves Newton’s equations of motion for the particles

that interact to each other in a system (atoms or molecules interacting with pair potentials).

Therefore, the basics of the method are classical mechanics.

i

npoti rrErmF

∂−=

∂=

)( 12 L

ii rt ∂∂ 2 (1.1)

where mi is mass, ri is position of atom pot

.1.1 Force field

lation systems vary from individual atom to multiple-atom molecules,

polyme

i, E is potential energy, and n is the number of

atoms in the system. The potential energy contains the inter-molecular and intra-molecular

interactions.

3

The simu

rs, and biological molecules. Meanwhile, force fields have been developed to vary

from only van der Waals forces to more complicated forces. A force field is the database of

the atom parameters, which describe the potential energy surface of entire classes of

molecules with reasonable accuracy. Therefore, the properties of the system or the molecules

12

Page 32: Fiber Lubrication: A Molecular Dynamics Simulation Study

were simulated by virtual technology.

COMPASS (condensed-phase optimized molecular potentials for atomistic

simulation studies) is the consistent force fields (CFF).29 COMPASS is also an ab initio force

field in which most of parameters are derived from ab initio data. COMPASS has the same

functional forms with its previous version of force fields, but differ mainly in the range of

functional groups that are use to parameterize the system. The total potential energy is

written as a combination of valence terms including cross coupling terms, and non-bond

interaction terms.30 The potential energy of COMPASS is expressed as follows:31

[ ]∑ −+−+−=b

pot bbKbbKbbKE 404

303

202 )()()(

[ ]∑ −+−+−+θ

θθθθθθ 404

303

202 )()()( HHH

[ ] [ ] [ ][ ]∑Φ

Φ−Φ−+−Φ−+Φ−Φ−+ )3cos(1)2cos(1)cos(1 033

022

011 VVV

F'

00' )'')(

Vbb )3cos()2cos()cos()( 3210

3cos()2cos()cos()''(b

VVVbb

Φ

∑+χ

χ χ2K

∑∑ −−+b b

bb bbbb(

∑∑ −−+θ θ

θθ θθθθ'

00' )'')((F

∑∑ −−+b

b bbFθ

θ θθ ))(( 00

[ ]∑∑Φ

Φ+Φ+Φ−+b

VV

[ ]∑∑Φ

Φ+Φ+Φ−+'

3210 )

[ ]∑∑Φ

Φ+Φ+Φ−+θ

θθ )3cos()2cos()cos()( 3210 VVV

Single valence

Cross valence

terms

terms

∑∑∑Φ

Φ −−Φ+θ θ

θθ θθθθ'

00' )'')((cosK

13

Page 33: Fiber Lubrication: A Molecular Dynamics Simulation Study

∑>

+ji ij

ji

rqq

∑> ⎥

⎢⎢

⎟⎟⎠

⎞⎜⎜⎝

⎛−⎟

⎟⎠

⎞⎜⎜⎝

⎛+

ji ij

ij

ij

ijij r

rrr

6090

32ε

(1.2)

n

angle (

clude a Lennard-Jones (LJ) 9-6 function for the van der

Waals

Coulomb term

van der Waals term

The valence terms represent internal coordinates of bond length (b), angle (θ), torsio

Φ), and out-of-plane angle (χ). K, H, V and F are all given as parameters and functions

in COMPASS, according to each equivalent term of bond length, angle, torsion, out-of-plan

angle and their cross terms. q represents the atom centered point charge and rij denotes the

distance between atom i and j. ε is the stress between atoms.31 There are seven cross covalent

terms in Equation 1.2. The cross terms are important to predicting vibrational frequencies

and structural variations associated with conformational changes. Among the cross terms in

the above equation, the bond-bond, bond-angle, and bong-torsion angle are the most

frequently used terms. In order to describe different systems, different models have been

used in the COMPASS force field.

The non-bond interactions in

(vdW) term and a Coulomb function for an electrostatic interaction, shown as the last

two terms in Equation 1.2. The non-bond terms are used for interactions between pairs of

atoms that are separated by two or more internal atoms or those belonging to different

molecules. The LJ-9-6 parameters (ε and r0) are described for the atom pairs with the same

14

Page 34: Fiber Lubrication: A Molecular Dynamics Simulation Study

elements. For the atom pairs of different elements, the off-diagonal parameter is calculated

by a sixth order combination law:31

61

6060 )()( ⎞⎛ + rr0

2 ⎟⎟⎠

⎜⎜⎝

= jiijr

(1.3)

⎟⎟⎞

⎠⎜⎜⎝

+

⋅⋅= 6060

3030

)()()()(

2ji

jijiij rr

rrεεε

(1.4)

In comparison with the comm n

region,

OMPASS extends the range as to include most of the common organic and

inorgan

3.1.2 The integrator and the time step in MD

ith the force field, the MD simulation is that

Newton

on LJ-12-9 function, which is ‘hard’ in the repulsio

the LJ-9-6 function is softer to repulsion but more attractive in the long separation

range.31

C

ic materials. The COMPASS contains more atoms than other second-generation force

fields and works accurately on molecular structures, vibrational frequencies, conformation

energies, dipole moments, liquid structures, crystal structures, equations of state, cohesive

energy densities, etc.

After defining the potential energy w

’s equations of motion are integrated numerically. The integrators assume that the

positions, velocities and accelerations can be approximated by a Taylor series expansion.31

We used Velocity Verlet method as the dynamics algorithm with Materials Studio software.

15

Page 35: Fiber Lubrication: A Molecular Dynamics Simulation Study

Velocity Verlet algorithm is a widely used integrators for the accuracy that is not

achieved by the previous Verlet algorithm.32 The position (ri), velocity (vi) and acceleration

(ai) of atom i at time t+δt (δt is the time step) are all obtained from the equivalents at the time

t in the following way:

2)()(21)( ttattv ii δδ ⋅+⋅+ (1.5) )()( tvttr ii δ =+

ttatvttv iii δδ ⋅+=+ )(21)()

21( (1.6)

ii

n 1)⋅ (1.7) pot

i

ii mr

rrEmF

tta(

)( 1

∂−==+

tδ (1.8) ttattvttv iii δδδ ⋅+++=+ )(21)

21()(

There are four steps and nine variables o

obtain

δt, is essential to success in MD simulations. It should be

at least

to obtain a subsequent state. The first step is t

the next position of one time step. Secondly, the velocity of half step is calculated with

previous acceleration. Then, the acceleration of the next full time step is derived from the

differential of potential energy defined by the force field. The velocity of the next full time

step is the amount of the velocity of half time step and the velocity increment according to

the new acceleration. For each time point, the position, the velocity and the acceleration of a

specified atom are synchronous. Although this integrator is resource-consuming, it is

commonly used in MD integrals.

The choice of the time step,

an order of magnitude less than the typical times of the system that is defined by

phonon frequencies or ratio of velocity to acceleration. If the time step is too large, errors

16

Page 36: Fiber Lubrication: A Molecular Dynamics Simulation Study

will occur in the integration, such as atoms crashing. If the time step is too small, errors will

take place from rounding in the computation.31 The default value of the time steps is 1.0 fs in

MD of Materials Studio 4.1, which was used in most of the current work.

3.1.3 Periodic boundary conditions

are the available approach to simulate a large system by

modeli

e nearest mirror

image

Periodic boundary conditions

ng a small unit cell adjacent to periodic mirror images at each side or facet. For

example, Figure 1.6 shows a sketch of periodic boundary conditions in a 2-D system. The

blue lattice is the unit cell we simulate. Other eight white lattices are the mirror images of the

unit cell, in which each mirror image has atoms with the same states as the simulation lattice.

When an atom moves out of the lattice, an equivalent atom in a mirror image moves in the

lattice simultaneously, such as the atom #2 moving in Figure 1.6. Under periodic boundary

conditions, the number of atoms in the simulate lattice is kept as a constant.

Another principle of periodic boundary conditions is that only th

is considered to calculate interactions. For example in Figure 1.6, there are more

options when we calculate the interaction between atom #3 and #1. Only the purple

interactions are taken into account due to the nearest pairs of the mirror images. This

principle is spontaneous when the edge length of the lattice is longer than the cutoff which is

the maximum distance of calculating interaction of atoms or groups. Atoms or groups beyond

17

Page 37: Fiber Lubrication: A Molecular Dynamics Simulation Study

the cutoff are ignored to a target atom or group when the interaction is being calculated.

Figure 1.6 The 2-D sketch of periodic boundary conditions. The blue lattice

is a simulation model of an atom system. The white lattices are the mirror images of the simulation model. The circled numbers refer to the atoms in the model lattice and mirror images. The red unidirectional arrows denote the motion trajectory of all #2 atoms. The dash circles represent the target positions of all #2 atoms. The yellow, green and purple bidirectional arrows indicate the interaction between #3 and #1 atom. The atom distance of the purple arrows is shortest.

1

5

3 4

1

5

3 4

1

5

3 4

1

5

3 4

1

5

3 4

1

5

3 4

1

5

3 4

2 2 2

1

5

3 4

1

5

3 4

2 2 2

2 2 2

18

Page 38: Fiber Lubrication: A Molecular Dynamics Simulation Study

e build two kinds of models with periodic boundary conditions: confined layers and

periodi

3.1.4 Controlling methods of temperature

only used in MD simulations: constant

atom n

ining

dynami

concept in dynamics simulation, and is the vital state

variabl

W

c cells. The former are 2-D periodic and 1-D confined to simulate ultra-thin substrates

and surfaces. The latter are 3-D periodic to duplicate the bulk materials.

There are four kinds of ensembles comm

umber, pressure, temperature (NPT), constant atom number, energy, volume (NVE),

constant atom number, pressure, enthalpy (NPH), and constant atom number, temperature,

volume (NVT). NVT ensemble is used to reach equilibrium structure under a certain

systemic density, and NPT ensemble to obtain a proper density of a specified system.

As for NVT and NPT ensembles, MD simulation is advantageous in obta

c properties directly because the MD method provides not only particle positions but

also the thermodynamic variables.31

Temperature is an important

e to specify the thermodynamic state of the system. Temperature is calculated from

the kinetic energy:

∑=

⋅== ET 2 N

iii

Bkin

B

vmNkNk 1

2

31

3vv

(1.9)

In the above equation, N denots i the number of particles in the system and mi and v

19

Page 39: Fiber Lubrication: A Molecular Dynamics Simulation Study

represent to the respective mass and velocity of the particle i. kB is Boltzmann’s constant.

There are four temperature-control methods (thermostats) optional in Materia

B

ls

Studio.

equatio

They are the direct velocity scaling, Berendsen, Nosé-Hoover and Andersen methods.

The direct velocity scaling can change the velocities of all atoms by the uniform

n:

system

target2

old

new

TT

vv⎜⎛ =⎟⎟

⎞⎜⎝

(1.10)

The velocities of

the sys

ach velocity

is mult

atoms are changed drastically by adding or subtracting energy from

tem. The target temperature can be exactly reached whatever the system temperature

is. It is because the direct velocity scaling method suppresses the natural fluctuations of a

system. The direct velocity scaling is only used in the equilibration in our work.

Berendsen method is called as the temperature-bath coupling, in which e

iplied by a velocity factor to limit atom velocities as well as the system temperature.

The factor works as follows:

21

01 ⎥⎦

⎤⎢⎣

⎡⎜⎛Δ

−⋅=Ttvv ⎟

⎠⎞

−T

Toldnew τ

(1.11)

where τ is the coupling parameter (0.1

thod33-35 is to extend Newton’s equations of motion by a

-0.4 ps) which determines how tightly the temperature

bath and the system are coupled together, Δt is the time step, T0 is the target temperature, and

T is the current temperature.

The Nosé-Hoover me

20

Page 40: Fiber Lubrication: A Molecular Dynamics Simulation Study

‘friction’ term:36, 37

ii

ii

Fa vmv

vv −= ⋅ζ (1.12)

The friction coefficient ζ(t) fluctuates in time around zero, according to the equation :

⎥⎦

⎤⎢⎡

−= ∑N

TNkvmd 2 31 vζ (1.13) ⎣ =i

BiiQdt 1

where Q is a user-defined constant.

rf ms on tan temperature dynamics that produces the

true can

osen particle by a

predefi

hod is used for NPT and Andersen stochastic

method

3.1.5 MD minimization methods

ed, a frequent activity in molecular simulation is the

optimiz

The Nosé-Hoover method pe or c s t-

onical ensembles in both coordinate space and momentum space.

Andersen stochastic method replaces the velocity of a randomly ch

ned collision period.38 In Materials Studio, the predefined collision period is

proportional to N2/3 where N is the atom number of the system. The velocity replacement

duplicates particle collisions with the wall.

In Materials Studio, Berendsen met

is used for NVT to control temperature in dynamics.

Before dynamics is start

ation or minimization of the system. The optimization or minimization aims to seek

the minimum value of the total energy in the system. There are three commonly used

minimization algorithms: the steepest descents, conjugate gradients, and Newton-Raphson

21

Page 41: Fiber Lubrication: A Molecular Dynamics Simulation Study

method.

In Materials Studio, the system provides four options of the algorithm for

minimi

3.2 Mesoscale simulations

of mesoscale simulation methods commonly used, mesoscopic

dynami

zation: Steepest descents, Conjugate gradient, Newton-Raphson methods, and Smart

method. Smart method is a joint method to combine Steepest descents, Conjugate gradient,

and Newton-Raphson methods in a cascading manner.

There are two kinds

cs (MesoDyn) and dissipative particle dynamics (DPD), which have been developed

in the past two decades. The model of polymer molecules used in both mesoscale simulations

consists of beads with various types of interactions. There are intra-molecular interactions

between beads of a molecule (Gaussian chain) and the mean field potential for all

intermolecular interactions. Each bead has a certain component type to represent covalently

bonded groups of atoms (objects in MD simulations) such as those given by one or a few

structural units of a polymer chain. Figure 1.7 depicted an example of

dimyristoylphosphatidylcholine (DMPC) with the mesoscale and the MD models. Compared

with MD simulation, these mesoscale methods employ soft interaction potentials, as a natural

consequence of the coarse-grained representation, allowing the length and time scales extend

by many orders of magnitude.

22

Page 42: Fiber Lubrication: A Molecular Dynamics Simulation Study

Figure 1.7 Comparison of the moleculare models in the mesoscale

simulation (left) and the MD simulation (right) with the

MesoDyn is based on the dynam

the dyn

example of dimyristoylphosphatidylcholine (DMPC) which is coarse grained using 13 beads to represent the 118 atoms. This picture was drawn by Nielsen et al.28

ic mean field density functional theory to calculate

amics of mesoscale morphology in a variety of block polymer mixtures and solutions,

including charged systems and solid-liquid. The phase separation dynamics are expressed as

the polymer diffusion with Langevin-type equations and the thermal fluctuations with an

additive noise know as the random force. This method was reported with success to the

aqueous solutions of PEO-PPO-PEO triblock copolymers.39-43 In the DPD method, the beads

are driven by conservative, dissipative, and random forces calculated for pairs of neighboring

beads. The time evolution of the system is to numerically integrate Newton’s equations of

motion. This DPD method is able to correctly descript the Navier-Stokes hydrodynamic

behavior.44, 45 MesoDyn was selected as the tools in this work to study the mesoscale

simulations, because there are a number of papers published with this method and we can

23

Page 43: Fiber Lubrication: A Molecular Dynamics Simulation Study

compare our results with other works.28, 39, 40, 42, 43, 46-51

In MesoDyn, the free energy (F) of a multiple-component liquid is a function of the

local de

ρρρ compcoh FF ++ (1.14)

where denotes the ideal free energy,

nergy s can be derived. The mean-field

intrinsi

nsity (ρ).

)(ρ idFF = )()()(

idF cohF represent the cohesive interactions with

Gaussian kernels, and compF refers to the com sibility

From the free e , all thermodynamic function

pres

c chemical potentials (µI) are obtained with the functional differentiation of the free

energy as follows:

)( FrI )()( rIρμ ρ∂ (1.15)

If equilibrium is reach

∂=

ed, )(rIμ =constant, which leads to the self-consistent field

equatio modns for the mean-field bead el. Generally, there are many solutions of these

equations, one of which is with the lowest free energy. The dynamics of the system includes

the diffusions in the component densities and the Gaussian noise in the system, which is

described by a set of functional Langevin equations.40 The following functional Langevin

equations for the diffusive dynamics of the density fields are depicted as follows:

[ ] ημμρρρ

+−∇∇=∂ I Mv ∂ JIJIt

(1.16)

[ ] ημμρρρ

−−∇∇=∂∂

IJJIJ Mv

t (1.17)

24

Page 44: Fiber Lubrication: A Molecular Dynamics Simulation Study

The distribution of the Gauss

fluctuation-dissipation theorem (Equation 1.18) and the time integration of the Langevin

equations resulting in the density fields with Boltzmann distributions (Equation 1.19):

ian noise, η, is restricted by two conditions: the

0),( =trη (1.18)

')'()'()' rJIr rrttt ∇−×∇−−= ρρδδβ

(1.19) 2,'(),( Mvrtr ηη

where M is a bead mobility parameter and v is the avera

coefficient,

ge bead volume. The kinetic

JIMv ρρ , indicates a local exchange mechanism. The Langevin equations are

constrained for an incompressible system with the dynamic condition:

v1trtr JI ),(),( =+ ρρ (1.20)

The parameters in MesoDyn simulations, the Flory-Huggins interaction parameter

relative to )(rIμ and the Gaussian chain topological parameter relative to segment sizes and

bead types, will be discussed in Section 1 of Chapter 6.

4. Summary

The science of friction and lubrication is attractive to both industrial and academic

research generation by generation. The study scale on this issue varies from macro scale of

Amontons’ friction law to micro scale. In the past decades, the great interest in the area has

been further focused to the nanoscale, with a term of so-called nanotribology. A great deal of

experimental and theoretical work has been done at micro and nanoscales with modern

25

Page 45: Fiber Lubrication: A Molecular Dynamics Simulation Study

instruments, such as SFA, QCM, LFM etc. However, friction and lubrication are still of

mystery at atomic levels because it is technically difficult to observe what happens at a

solid-liquid interface during sliding with existing experimental tools. However, a

computational simulation approach opens an opportunity to predict and analyze interfacial

phenomena, such as the adsorption, contact angle, interaction, friction and morphologies of

polymer-surface systems at both nanoscale and mesoscale. This chapter described and

discussed some critical technical issues of simulations, such as focefield, minimizer,

integrator, controlling methods of temperature and periodic boundary conditions. MD

simulations and MesoDyn method are proposed on a comprehensive study on the lubrication

of fibers from nanoscale to mesoscale.

The following chapters focus on lubrication models for fiber surfaces. In Chapter 2,

fiber surface models are successfully built and they are the confined amorphous polymeric

layers with appropriate densities, cohesive energies, solubility parameters, volumetric

thermal expansions and contact angles. Especially, the confined amorphous cellulose layer

has not been found in published works. Chapter 3 predicts several properties of polymeric

lubricants by the topology method. The lubricants are PAG diblock copolymers and triblock

copolymers equivalent to commercial products of DOW and BASF. Chapter 4 calculates

interaction energies between PAG block polymers and fiber surfaces. In Chapter 5,

complicated polymer-water-surface (PWS) systems are constructed where a single PAG

26

Page 46: Fiber Lubrication: A Molecular Dynamics Simulation Study

block copolymer molecule is set in water on hydrophilic or hydrophobic surfaces. A novel

method, PWS interaction calculation, is proposed to calculate the interaction between the

single PAG polymer and the PP or cellulose surface with the presence of water. The results

of interaction energies are in agreement of the simulation configuration and experimental

works from our coworkers. Chapter 6 focuses on mesoscale simulation of PAG triblock

copolymer solution, in which three conditions, bulk, confined layer and shear, are studied.

Finally, future work is discussed in Chapter 7.

27

Page 47: Fiber Lubrication: A Molecular Dynamics Simulation Study

References:

1. Morton, W. E.; Hearle, J. W. S., Physical properties of textile fibres. The Textile Institute: Manchester, UK, 1993; p 725.

2. Dowson, D., History of tribology. 2nd ed.; Professional Engineering Publishing: London, 1998.

3. Howell, H. G., Friction in textiles. In Textile Book Publishers in association with the Textile Institute: New York, 1959; p 263.

4. Gupta, B. S., Friction in textile materials. Crc; Woodhead: Boca Raton, Fla.; London, 2007; p 462.

5. H. G. Howell, K. W. M., D. Tabor, Friction in textiles Textile Book Publishers in association with the Textile Institute: New York, 1959.

6. Israelachvili, J. N., Strength of van der Waals Attraction between Lipid Bilayers. Langmuir 1994, 10, (9), 3369-70.

7. Yoshizawa, H.; Chen, Y. L.; Israelachvili, J., Recent advances in molecular level understanding of adhesion, friction and lubrication. Wear 1993, 168, (1-2), 161-166.

8. Yoshizawa, H.; Israelachvili, J., Fundamental mechanisms of interfacial friction. 2. Stick-slip friction of spherical and chain molecules. J. Phys. Chem. 1993, 97, (43), 11300-13.

9. Krim, J., Friction at the nanoscale. Phys. World 2005, 18, (2), 31-34.

10. Krim, J.; Solina, D. H.; Chiarello, R., Nanotribology of a krypton monolayer: a quartz-crystal microbalance study of atomic-scale friction. Phys. Rev. Lett. 1991, 66, (2), 181-4.

11. Ohmae, N.; Mori, S.; Martin, J. M., Micro and nanotribology. ASME Press: New York, 2005; p 185.

12. Li, Y.; Hinestroza, J., Boundary lubircation phenomena in coated textile surfaces. In Friction in textile materials, Gupta, B. S., Ed. Crc; Woodhead: Woodhead publishing in textiles, 2007; pp 419-447.

13. Bezerra, V. B.; Blagov, E. V.; Klimchitskaya, G. L., Some estimations of the atomic friction coefficient by scanning of the AFM tip along a constant force surface. Surface Review and Letters 1999, 6, (3-4), 341-5.

14. Bhushan, B.; LaTorre, C., Investigation of scale effects and directionality dependence on friction and adhesion of human hair using AFM and macroscale friction test apparatus. Ultramicroscopy 2006, 106, (8-9), 720-34.

28

Page 48: Fiber Lubrication: A Molecular Dynamics Simulation Study

15. Gulbinski, W.; Pailharey, D.; Suszko, T.; Mathey, Y., Study of the influence of adsorbed water on AFM friction measurements on molybdenum trioxide thin films. Surface Science 2001, 475, (1-3), 149-58.

16. Liang, Q.; Hongnian, L.; Yabo, X.; Xudong, X., Friction and adhesion between C60 single crystal surfaces and AFM tips: effects of the orientational phase transition. Journal of Physical Chemistry B 2006, 110, (1), 403-9.

17. Zhenhua, T.; Bhushan, B., Surface modification of AFM Si3N4 probes for adhesion/friction reduction and imaging improvement. Transactions of the ASME. Journal of Tribology 2006, 128, (4), 865-75.

18. Coffey, T. Nanotribology fundamentals : predicting the viscous coefficient of friction. North Carolina State University, Raleigh, 2004.

19. Coffey, T.; Abdelmaksoud, M.; Krim, J., A scanning probe and quartz crystal microbalance study of the impact of C60 on friction at solid-liquid interfaces. Journal of Physics: Condensed Matter 2001, 13, (21), 4991-9.

20. Coffey, T.; Krim, J., Impact of substrate corrugation on the sliding friction levels of adsorbed films. Physical Review Letters 2005, 95, (7), 076101.

21. Krim, J. In QCM studies of phononic and electronic contributions to sliding friction in adsorbed monolayers, Washington, D.C., United States, 2005; American Society of Mechanical Engineers, New York, NY 10016-5990, United States: Washington, D.C., United States, 2005; pp 489-490.

22. Krim, J., QCM tribology studies of thin adsorbed films. Nano Today 2007, 2, (5), 38-43.

23. N. Behary, C. C. C. C. A. P., Using an electronic microbalance technique to study the stick-slip behavior of lubricated polypropylene fibers. Journal of Applied Polymer Science 2003, 89, (3), 645-654.

24. Braun, O. M.; Naumovets, A. G., Nanotribology: Microscopic mechanisms of friction. Surface Science Reports 2006, 60, (6-7), 79.

25. Persson, B. N. J.; Mugele, F., Squeeze-out and wear: fundamental principles and applications. Journal of Physics: Condensed Matter 2004, 16, (10), R295-R355.

26. Song, J., Adsorption of amphoteric and nonionic polymers on model thin films. 2008; p 232.

27. Pedersen, J. S.; Sommer, C., Temperature dependence of the virial coefficients and the chi parameter in semi-dilute solutions of PEG. Prog. Colloid Polym. Sci. 2005, 130, 70-78.

28. Nielsen, S. O.; Lopez, C. F.; Srinivas, G.; Klein, M. L., Coarse grain models and the computer simulation

29

Page 49: Fiber Lubrication: A Molecular Dynamics Simulation Study

of soft materials. Journal of Physics Condensed Matter 2004, 16, (15), 481-512.

29. Sun, H.; Ren, P.; Fried, J. R., COMPASS force field: Parameterization and validation for phosphazenes. Computational and Theoretical Polymer Science 1998, 8, (1-2), 229.

30. Sun, H., COMPASS: an ab initio force-field optimized for condensed-phase applications-overview with details on alkane and benzene compounds. Journal of Physical Chemistry B 1998, 102, (38), 7338.

31. Material Studio, Release 4.1. In Accelrys Software Inc.: San Diego, 2006.

32. Omelyan, I. P.; Mryglod, I. M.; Folk, R., Optimized Verlet-like algorithms for molecular dynamics simulations. Physical Review E - Statistical, Nonlinear, and Soft Matter Physics 2002, 65, (5), 056706.

33. Aoki, K.; Hoover, C. G.; De Groot, S. V.; Hoover, W. G., Time-reversible deterministic thermostats. Physica D: Nonlinear Phenomena 2004, 187, (1-4), 253-267.

34. Hoover, W. G.; Hoover, C. G., Nonequilibrium temperature and thermometry in heat-conducting 4 models. Physical Review E - Statistical, Nonlinear, and Soft Matter Physics 2008, 77, (4), 041104.

35. Posch, H. A.; Hoover, W. G., Time-reversible dissipative attractors in three and four phase-space dimensions. Physical Review E. Statistical Physics, Plasmas, Fluids, and Related Interdisciplinary Topics 1997, 55, (6-A), 6803-6810.

36. Nose, S. In Constant-temperature molecular dynamics, Bristol, Engl, 1990; Publ by Adam Hilger Ltd: Bristol, Engl, 1990; p 115.

37. Nose, S., Molecular dynamics simulations in condensed matter physics. Proceedings of the International Conference and Exhibition on Computer Applications to Materials Science and Engineering 1991, 341.

38. Thijssen, J. M., Computational physics. Cambridge University Press: Cambridge, UK ; New York, 2007; p 620.

39. Fraaije, J. G. E. M., Dynamic density functional theory for microphase separation kinetics of block copolymer melts. J. Chem. Phys. 1993, 99, (11), 9202-12.

40. Fraaije, J. G. E. M.; Van Vlimmeren, B. A. C.; Maurits, N. M.; Postma, M.; Evers, O. A.; Hoffmann, C.; Altevogt, P.; Goldbeck-Wood, G., The dynamic mean-field density functional method and its application to the mesoscopic dynamics of quenched block copolymer melts. J. Chem. Phys. 1997, 106, (10), 4260-4269.

41. Fraaije, J. G. E. M.; Zvelindovsky, A. V.; Sevink, G. J. A., Computational soft nanotechnology with mesodyn. Molecular Simulation 2004, 30, (4), 225-238.

30

Page 50: Fiber Lubrication: A Molecular Dynamics Simulation Study

42. Fraaije, J. G. E. M.; Zvelindovsky, A. V.; Sevink, G. J. A.; Maurits, N. M., Modulated self-organization in complex amphiphilic systems. Molecular Simulation 2000, 25, (3-4), 131-144.

43. Sevink, G. J. A.; Zvelindovsky, A. V.; van Vlimmeren, B. A. C.; Maurits, N. M.; Fraaije, J. G. E. M., Dynamics of surface directed mesophase formation in block copolymer melts. Journal of Chemical Physics 1999, 110, (4), 2250.

44. Litvinov, S.; Ellero, M.; Hu, X.; Adams, N. A., Smoothed dissipative particle dynamics model for polymer molecules in suspension. Physical Review E - Statistical, Nonlinear, and Soft Matter Physics 2008, 77, (6).

45. Liu, M.; Meakin, P.; Huang, H., Dissipative particle dynamics with attractive and repulsive particle-particle interactions. Physics of Fluids 2006, 18, (1).

46. Alexandridis, P.; Olsson, U.; Lindman, B., Phase Behavior of Amphiphilic Block Copolymers in Water-Oil Mixtures: The Pluronic 25R4-Water-p-Xylene System. J. Phys. Chem. 1996, 100, (1), 280-8.

47. Guo, S. L.; Hou, T. J.; Xu, X. J., Simulation of the phase behavior of the (EO)13(PO)30(EO)13(Pluronic L64)/water/p-xylene system using MesoDyn. Journal of Physical Chemistry B 2002, 106, (43), 11397-11403.

48. Jawalkar, S. S.; Nataraj, S. K.; Raghu, A. V.; Aminabhavi, T. M., Molecular dynamics simulations on the blends of poly(vinyl pyrrolidone) and poly(bisphenol-A-ether sulrone). Journal of Applied Polymer Science 2008, 108, (6), 3572-6.

49. Li, Y.; Hou, T.; Guo, S.; Wang, K.; Xu, X., The Mesodyn simulation of pluronic water mixtures using the 'equivalent chain' method. Phys. Chem. Chem. Phys. 2000, 2, (12), 2749-2753.

50. Yang, S.; Yuan, S.; Zhang, X.; Yan, Y., Phase behavior of triblock copolymers in solution: Mesoscopic simulation study. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2008, 322, (1-3), 87-96.

51. Zhao, Y.; Chen, X.; Yang, C.; Zhang, G., Mesoscopic simulation on phase behavior of pluronic P123 aqueous solution. Journal of Physical Chemistry B 2007, 111, (50), 13937-13942.

31

Page 51: Fiber Lubrication: A Molecular Dynamics Simulation Study

Chapter 2

Molecular Modeling of Amorphous Fiber Surfaces

Abstract

Fiber surfaces are soft and deformable. It is interesting and challenging to study interface properties in MD simulations. Polypropylene (PP) and cellulose are two of the most commonly used materials to make fibers. PP and cellulose confined amorphous surface layers are built successfully, and the physical properties are inspected in MD simulations, such as the density, cohesive energy, volumetric thermal expansion, and contact angle of the surfaces. The xenon crystal is a soft confining boundary to compact polymers. When the contact angle is discussed, a water nano-droplet is created on the PP and cellulose surfaces. The shape of the nano-droplet is found spreading on the cellulose surface, and no significant change is observed on the PP surface. The contact angles of PP and cellulose surfaces are 106° and 33°, respectively. The results prove that the PP surface is hydrophobic while the cellulose surface is hydrophilic, which is in agreement with previous experimental values.

1. Introduction

By 1980 there was a great deal of research done on the physical structures and

properties of textile fibers. A semi-crystalline structure (a mixture of crystal and amorphous

structure) was proposed for most textile synthetic fibers, such as cellulose (Lyocell and

Rayon), polypropylene (PP), polyethylene (PE), polyethylene tephthalate (PET), and

polyamide (nylon 6 and nylon 6,6). Although crystallization has been discovered in synthetic

fiber production and is vital to the strength of a spinnable fiber, the amorphous regions in

synthetic fibers contribute to the fiber properties of dying, finishing, softness, and

particularly to surface friction and shear strength, which are very important in textile

processes due to interfacial friction of fiber-to-fiber and fiber-to-metal contacts in textile

processing. Besides industrial manufacturing, the amorphous structure has been found to

32

Page 52: Fiber Lubrication: A Molecular Dynamics Simulation Study

form in film-forming experiments, such as spin coating. Spin coating creates thin films, with

thicknesses ranging from a few nanometers to about a micrometer. The differences in the

interfacial properties of thin films from the properties of the bulk fluid are critical to the layer

formation. Song1 produced PP, PE, PET and cellulose thin films to study absorption and

found that the thin films made by spin coating were in the amorphous state. Studying thin

films created by spin-coating is an alternating approach to probe fibers’ surface properties at

a microscale (1-100 µm) or nanoscale (1-1000 nm). The techniques often used are SFA2-4,

LFM5-10, and QCM11-15. However, little has been done at a molecular level (0.1-10 nm)

toward understanding the behaviors of amorphous surfaces mainly due to the limitation of

experimental apparatuses. MD simulations provide a capability of probing surface properties

of textiles.

In order to apply MD in the study of polymer surface properties, the primary step is to

create a polymer surface. In general, fiber surfaces or polymer films are more or less

amorphous. An amorphous polymer surface used in MD is desirable. The ideal situation is to

use a large number of chains. However, the requested calculation scales with the square of

the atomic number of a system. It is necessary to simulate a larger system with a small

periodic lattice. Several methods have been adapted to create cellulose surfaces. Theodorou

and Suter16 developed a method of building amorphous cells in 1986, which allowed them to

build a 3-D periodic cell but not a surface. They simulated the amorphous state of

polypropylene using single chain conformations that were growing within a periodic box.

Periodic boundary conditions of the unit cell provided a means of forming a space-filling

structure starting from a single chain conformation. The presence of periodic replicates of

33

Page 53: Fiber Lubrication: A Molecular Dynamics Simulation Study

parent chains contributed a simple and elegant way to introduce intermolecular interactions

without increasing the number of atoms necessary to create the solid state.

Kavassalis et al.17 compared results derived from a single chain to those derived from

multiple chains and found the errors from the amount of chains quite acceptable given the

other error sources. In Kavassalis’ work, the models of amorphous liquid state were

represented by a single molecule in a periodic cell. The solubility parameters estimated from

the models were in good agreement with those values derived from models having eight NPE

molecules packed into a cell with the exclusion of the electrostatic interactions, which are the

most sensitive to system size effects. Choi et al.18 continued Kavassalis’ work to build

amorphous layers of (hydroxyethyl) cellulose (HEC) and (hydroxypropy1) cellulose (HPC),

which are the most commonly used celluloses. Kela et al.19 also built (2 0 0) crystalline

cellulose surface with five linear (twenty-glucose-unit-long) glucose arrays. In another paper,

Chauve et al.20 constructed (1 1 0) cellulose Iβ surfaces with the initial monoclinic unit cell of

two chains by Cerius2 software. The advantage of using crystalline cellulose in MD

simulations is the surface flatness, while the surface properties may be discrepant from the

cellulose surfaces made by the spin-coating experiments due to the surface roughness and the

freedom of molecules on upper layers. Chen et al.21, 22 developed a cellulose cell in

amorphous state with Cerius2 package from Accelrys. The model contained only one

cellulose chain consisting of 20 cellobiose repeat units (843 atoms), which were set in a

relaxant state by molecular mechanics. The initial conformation of the cellulose chain in each

model was generated using a Monte Carlo method with a 20×20×20 Å cubic cell. Periodic

boundary conditions were applied to their model, with an initial density of 1.40 g/cm3.

34

Page 54: Fiber Lubrication: A Molecular Dynamics Simulation Study

Derecskei and Derecskei-Kovacs23 also built oligomer cellulose cells. The densities and

solubility parameters were found to change linearly with the degree of polymerization (from

monomers to dodecamers). Although the periodic cell of amorphous cellulose has been

intensively developed and studied, molecular models of cellulose surfaces have not been

found in the literature.

As for other polymers, such as polypropylene (PP), polyethylene (PE) and nylon 6,6,

crystalline substrates are commonly used in the literature. Misra et al.24 stated that the

amorphous polybutadiene thin film was modeled by arbitrarily elongating one of the periodic

boundary conditions to a long coordinate and equilibrate it with the NVT ensemble. With the

same method in MD simulations, Natarajan et al.25 built a thin film with styrene-butadiene

compolymer, He et al.26 built with polyethylene, and Clancy and Mattice27 built with several

variations of polyolefin. With their results, the drawbacks of Misra’s method were obvious at:

(1) high and uncontrollable roughness, (2) non-uniform density normally, (3) the high surface

energy. It could be the reason of these drawbacks that polymers are not confined after

arbitrarily elongating one dimension.

Hirvi et al.28 used the confined condition to build polyethylene (PE) and poly-vinyl

chloride (PVC) amorphous substrates for studying the contact angles. Simulations were

started using the same method for creating initial structures for bulk simulations. Twenty-

four PE molecules or sixteen PVC molecules compose a simulation system with lateral

dimensions of 78 Å. Hirvi et al.’s work did little further study on the physical properties of

amorphous substrates. Prathab et al.29 modeled poly(methyl methacrylate) (PMMA) film

with only one molecule under the confined condition, but little modeling details were

35

Page 55: Fiber Lubrication: A Molecular Dynamics Simulation Study

provided. The essential approach of Hirvi et al.28 and Prathab et al.29 was to add a pair of

visible hard surfaces on the confined dimension. The hard surfaces can arbitrarily rebound

atoms back to the system lattice if these atoms move close to them. This method works fairly

well with the polymers with soft linear chains, such PE, PVC and PMMA.28, 29 However the

cellulose molecule is difficult to compress with hard surfaces because the atoms crash when

glucose rings are folded at the confined border. The key point to solve this failure is to avoid

folding the glucose rings with a low density.

In our work, fiber surfaces were made of polymers by MD simulation using confined

layer condition. This chapter aims to build cellulose as well as PP substrates and calculate the

physical properties of each substrate.

2. Simulation method

The MD modeling was performed by two modules of Materials Studio 4.1 software

from Accelrys, San Diego, CA, USA. The MD calculation module was Discover with the

COMPASS force field. Amorphous Cell was used to build primary layers and run dynamics

and analyses processing. All work was done on the Dell® SC1420 workstation with

Windows® server 2003, double Intel® Xeon™ 2.8GHz CPUs, and 4 GB memory.

2.1 Models of building confined amorphous layer models

2.2.1 Repeat units

The structures of the cellulose and PP repeat units, 1,4-βD glucose and propylene,

were provided by the library of Materials Studio 4.1 software and are depicted in Figure 2.1.

36

Page 56: Fiber Lubrication: A Molecular Dynamics Simulation Study

The atom properties were defined by the COMPASS force filed.

(b)(a)

Figure 2.1. Repeat units for building polymer (fiber) substrates (a) 1,4-βD glucose, (b) propylene. (Atom colors: carbon atoms—grey, hydrogen—white, and oxygen—red; the head atom is circled by cyan, and the tail atom is circled by magenta).

2.2.2 Build molecules

(a)

(b)

Figure 2.2. Polymer molecules: (a) Cellulose, and (b) Polypropylene (PP). The end-to-end distance of the cellulose molecule with 8 repeat units is 42.5 Å. The end-to-end distance of the PP molecule with 30 repeat units is 76.7 Å. Atom colors are set the same as Figure 2.1.

The cellulose molecule was built with 8 repeat units (1,4-βD glucose) under a head-

to-tail orientation and isotactic tacticity (see Figure 2.2(a)). Then, thirty repeat units

(propylene) were employed to construct the PP molecule with the same conditions (see

Figure 2.2(b)). The end-to-end distances of cellulose and PP molecules are 42.5 Å and 76.7

Å in vacuum after 5000 femtosecond (fs) minimization.

37

Page 57: Fiber Lubrication: A Molecular Dynamics Simulation Study

2.2.3 Build the polymeric confined amorphous unit layers

This step aimed to build the unit layer with which we could expand a large layer if

required. With Amorphous Cell®, the polymer molecules were packed into a confined layer

or a periodic cell. The confined layer was used as a cell to build the thin substrate of

polymers. The temperature was defaulted as 298K, and the time step was always 1 fs. The

COMPASS force field was used with the group-based cutoffs.

To build the amorphous polymer surface, it is necessary to find the confining surface.

The ideal confining surface should be ultra-flat and uniform, with small interaction to

polymers, square facet and solid at a target temperature. However there is not such a solid

material in reality. Fortunately, xenon crystals, fixed their atom fractional positions, are ultra-

flat and uniform crystals with inert properties, individual spheres, and fractional positions

fixed as solid at 298 K. This confining reference was a highlight point to build the confined

polymeric amorphous layer. So xenon crystals were the perfect confining reference due to the

inert character, the flat crystal slice, and the solid state at the room temperature.

(1) The PP confined unit layer:

The PP molecular chain is too soft to be compacted, because it does not have a ring,

long side group, or double bond in the chain. Three PP molecules were first built in the

confined layer directly by Amorphous Cell® with a density of 0.822 g/cm3. It is shown in

Figure 2.3(a). The side lengths of X and Y of the PP layer were defined as identical with the

reference of a xenon crystal. Other parameters kept constant as default values of Materials

Studio software. Then the PP unit layer was inserted between two xenon crystal substrates

38

Page 58: Fiber Lubrication: A Molecular Dynamics Simulation Study

which were defined as “Fix Fractional Position”. The fractional XYZ position property in the

fractional space of the lattice was kept constant for the selected atoms. A 50Å-thick vacuum

space was left on the top and covered the rest of layers in order to diminish the z-axis-

direction periodic interruption. The xenon-PP-xenon layer is shown in Figure 2.3(b). Next,

the NVT dynamics ensemble was enforced with equilibrium time of 5 picoseconds (ps) and

dynamics time of 1 nanosecond (ns). The final frame is shown in Figure 2.3(c). Finally, the

vacuum space and two xenon crystals were cropped off by cell building in Materials Studio.

The PP unit confined layer is shown in Figure 2.3(d), which is periodic in the XY plane and

confined in the Z direction.

(a) (b) (c) (d)

Figure 2.3 The procedures of building PP confined layer. (a) The PP

original confined layer: 16.7×16.7×27.6 Å. (b) The original xenon-PP-xenon cell: 16.7×16.7×90.6 Å. (c) The xenon-PP-xenon cell after the dynamics procedure. (d) The PP confined unit layer: 16.7×16.7×26.0 Å. Xenon atoms are colored by cyan. Other atom colors are set the same as Figure 2.1.

(2) The cellulose confined unit layer:

The initial densities of cellulose were defined as 0.5 g/cm3 which specified the target

density in the first step to build layers. Six cellulose molecules were packed to form a

cellulose substrate. The NPT dynamics parameters used in the step have a temperature of 298

K, an equilibration time of 10 ps, a pressure of 0.0001 GPa, a time step of 1 fs and a dynamic

39

Page 59: Fiber Lubrication: A Molecular Dynamics Simulation Study

time of 1ns. The original layer is shown in Figure 2.4(a). Then a sandwich layer, xenon-

cellulose-xenon, was built as shown in Figure 2.4(b). The vacuum space was not necessary

for cellulose because the next step was involved in NPT ensemble and the volume was

variable. There were two dynamics processes in the third step, 1 ns of NPT with 0.0001 GPa

(approximately equal to 1 atm) and 500 ps of NVT dynamics procedures. Additional

parameters were the same as the method used for PP. The final dynamics frame is shown in

Figure 2.4(c). The last step was the same as the pervious method of building the PP layer,

and the cellulose confined unit layer is shown in Figure 2.4(d).

(a) (b) (c) (d)

Figure 2.4 The procedures of building cellulose confined layer. (a) The

cellulose original confined layer: 23.6×23.6×46.3 Å. (b) The original xenon-cellulose-xenon cell: 23.6×23.6×63.0 Å. (c) The xenon-cellulose-xenon cell after the dynamics procedure: 16.7×16.7×44.4 Å. (d) The cellulose confined unit layer: 16.7×16.7×31.5 Å. Xenon atoms are colored by cyan. Other atom colors are set the same as Figure 2.1.

2.2.4 Super layer construction

Due to the periodic border in the XY plane, the unit confined layer was expanded to a

super layer in the X and Y directions. The super layer had X and Y lattice vectors which

were integral multiples of their equivalents in the original unit. The Z direction was confined

and not expandable.

40

Page 60: Fiber Lubrication: A Molecular Dynamics Simulation Study

(a) (b)

(c) (d)

Figure 2.5 The PP and cellulose confined amorphous layers. (a) The PP super layer (4×unit layers): 33.3×33.3×26.0 Å. (b) The cellulose super layer (4×unit layers): 33.3×33.3×32.3 Å. (c) The PP super layer (8×unit layers): 66.6×66.6×26.0 Å. (d) The cellulose super layer (8×unit layers): 66.6×66.6×32.3 Å.

The expansions were implemented by “Build—Symmetry—Supercell” in Materials

Studio 4.1. The expansion multiples of OA (X) and OB (Y) were selected as 2 and OC (Z) is

kept as 1. Then new super layers were created to have a surface area four times larger than

41

Page 61: Fiber Lubrication: A Molecular Dynamics Simulation Study

the unit layers. These new super layers are shown in Figure 2.5(a) and (b) and can be used in

the property analysis, except for contact angle (CA). The lager surfaces were required for

creating contact angle models. The larger CA layers were constructed by repeating above

expansions on the last super layers whose surface areas were eight times of the unit layers, as

shown in 2.5(c) and (d).

2.2.5 Contact angle (CA) models

Firstly, 500 water molecules were built into a periodic cube with the density at 1

g/cm3 using Amorphous Cell®, shown in Figure 2.6(a). The cubic lattice was then removed

to cancel the periodic condition and the cube was separated from bulk (see Figure 2.6(b)).

500 ps NVT dynamics procedure was executed and a water cube became a water nano-

droplet in the ball shape shown in Figure 2.6(c).

(a) (b) (c)

Figure 2.6 The steps of building the round water droplet. There are 500 water molecules involved. (a) The original periodic water cube: 24.6×24.6×24.6 Å. (b) The water cube without the periodic lattice. (c) The round water nano-droplet. Its diameter is approximately 32 Å.

42

Page 62: Fiber Lubrication: A Molecular Dynamics Simulation Study

The contact angle models were built by adding a 100Å-thick vacuum space on the

confined layers shown in Figure 2.6(c), followed by setting of the above water nano-droplet

on the confined layers by moving tools of Materials Studio 4.1, as shown in Figure 2.7.

(a) (b)

Figure 2.7 The contact angle models. (a) A water nano-droplet on the PP

surface. (b) A water nano-droplet on the cellulose surface.

2.2 Model property calculations

2.2.1 Cohesive energy density (CED) and solubility parameter

CED indicates the energy that breaks all intermolecular physical links in a unit

volume of a material. In polymer materials the physical links includes van der Waals force,

Coulomb force and hydrogen bonding. The earliest understanding of CED was involved in

the vaporization of liquid as seen in the following equation:30

mVRTHCED −Δ

= (2.1)

where ΔH represents the heat of vaporization, R denotes the gas constant, T refers to

temperature, and Vm is the molar volume.

43

Page 63: Fiber Lubrication: A Molecular Dynamics Simulation Study

When CED is calculated for a simulated polymer, the total energy, the intermolecular

energy and the intramolecular energy can be defined by the COMPASS force field. Eichinger

et al.31 compared the CED simulation results with the experimental values and claimed that

the simulation value was usually higher than the experimental ones. Due to the cutoff when

calculating interaction between atoms or groups, the truncation effect of the non-bond

potential by using periodic boundaries has to be taken into account. Hence, the cohesive

energy is given by adding tails, such as an ensemble average of the quantity. The CED

equation of MD simulation can be expressed as follows:

m

tailintratotal

VEEE

CED)(δ+−

= (2.2)

where Etotal represents the total energy, Eintra denotes the intramolecular energy, Etail refers to

the tail energy of the cutoff truncation, and Vm is the molar volume.

CED is such a fundamental property of a polymer that many properties can be

predicted from CED. CED is most often used to calculate the solubility parameter. In 1919

Joel H. Hildebrand32 proposed the concept of solubility indicating the solvency behavior of a

specific solvent. The solubility parameter (δ) is derived from CED as follows:

CED=δ (2.3)

In addition to CED, the solubility parameter (δ) indicates the solvent-solute

interactions that are strong enough to separate solvent-solvent molecules and solute-solute

molecules in order to minimize the Gibbs free energy of the entire system.

All CED calculations were carried on with the analysis function of Materials Studio

4.1 from Accelrys Software Inc. The PP and cellulose layers shown in Figure 2.5 (a) and (b)

44

Page 64: Fiber Lubrication: A Molecular Dynamics Simulation Study

were used in CED analysis.

2.2.2 Stepwise temperature procedure (NPT)

The stepwise temperature protocol is implemented by the Temperature Cycle protocol,

a function of Amorphous Cell®. The procedure performs a stepwise, constant pressure

molecular dynamics heating and/or cooling process to investigate thermal expansion,

constant pressure heat capacity, and the so-called 'molecular dynamics glass transition'.

The parameters of stepwise temperature protocol used in this work are as follows:

pressure—0.0001 GPa, lower temperature—200 K, upper temperature—800 K, number of

temperature stage—5, stage duration—50000 fs, starting temperature—298 K, average

period—2000 fs and number of half cycles—1. Other dynamics parameters are the default

values as above.

2.2.3 Contact angle

Contact angle is the angle at which a liquid/vapor interface meets a solid surface.33

The contact angle is specific for any given system and is determined by the interactions

across the three interfaces34, 35. The contact angle is usually used to characterize the

hydrophilic/hydrophobic properties of solid surfaces.

In the contact angle models built previously, NVT dynamics time was 1ns for both

models. The temperature and the time step were the defaults.

45

Page 65: Fiber Lubrication: A Molecular Dynamics Simulation Study

3. Results and discussion

3.1 Density and weight distribution

The final density of the PP and cellulose confined amorphous layers were 0.873

g/cm3 and 1.446 g/cm3 respectively. The value of 0.873 g/cm3 for PP is exactly within the

range between 0.855 g/cm3 of amorphous and 0.946 g/cm3 of crystalline36. Therefore, the PP

layer built at the density of 0.873 g/cm3 is quite close to the experimental values. The density

of the cellulose amorphous layers was the result of the dynamic procedure with the

COMPASS force field. It was lower than the density of crystal cellulose, 1.599 g/cm3 from

Sun’s work37. Sun37 also stated that the true density of cellulose fell in the range of 1.40 and

1.47 g/cm3. Mazeau and Heux38 simulated crystalline and amorphous cellulose layers and

obtained a average density of 1.3762 g/cm3 using the PCFF force field. Chen et al.21, 22 found

the density of their amorphous periodic cellulose cell was 1.385 g/cm3. Derecskei and

Derecskei-Kovacs23 addressed that their oligomer cellulose model had the density of 1.42

g/cm3. Our value of the amorphous cellulose layer was valid when compared to the values in

the literature.

46

Page 66: Fiber Lubrication: A Molecular Dynamics Simulation Study

0 5 10 15 20 25 30 350.0

0.5

1.0

1.5

2.0 Cellulose at X axis Cellulose at Y axis Cellulose at Z axis

Wei

ght d

istri

butio

n (%

)

Distance (Angstrom)

0 5 10 15 20 25 30 350.0

0.5

1.0

1.5

2.0 PP at X axis PP at Y axis PP at Z axis

Wei

ght d

istri

butio

n (%

)

Distance (Angstrom)

Figure 2.8 The weight distribution of PP and cellulose confined amorphous

layers.

The weight distribution of PP and cellulose confined amorphous layers are shown in

Figure 2.8. There are 100 slices in each X, Y and Z direction. The weight distributions of

both PP and cellulose layers are slightly oscillating but still quite stable at 1% in the X and Y

directions. Therefore, the PP and cellulose confined layers are uniform in the XY plane. The

first and final values at the Z axis are equal to zero which corresponds to the weight at the top

and bottom of the surface. They display boundary confined conditions for the surfaces. The

rough ranges of the confined surfaces are only of about 1-2 Å and so the surfaces are ultra

47

Page 67: Fiber Lubrication: A Molecular Dynamics Simulation Study

flat. The uniformity of the amorphous layers is one of advantages of using the xenon crystal

in MD.

3.2 Cohesive energy density (CED)

The CEDs of the PP and cellulose layers are listed in Table 2.1. The average CEDs of

the PP and cellulose confined layers are 2.1E+08 and 4.9E+08 J/m3 respectively. Apparently,

the cellulose has a stronger intermolecular force than the PP in a specified volume. As for the

PP confined layers, the van der Waals CED is much higher than the electrostatic CED.

Table 2.1 CED and solubility parameters of cellulose and PP layers

PP Cellulose Average Std. Dev Average Std. Dev Cohesive Energy Density (J/m3) 2.110E+08 1.863E+06 4.905E+08 7.386E+06 - Electrostatic (J/m3) 1.484E+05 3.795E+04 4.463E+08 5.503E+06 - van der Waals (J/m3) 2.108E+08 1.876E+06 4.428E+07 7.028E+06 Solubility Parameter ((J/cm3)0.5) 14.52 0.064 22.15 0.167

- Electrostatic ((J/cm3)0.5) 0.38 0.054 21.12 0.130 - van der Waals ((J/cm3)0.5) 14.52 0.065 6.63 0.535

Therefore, the van der Waals interactions contribute most to PP intermolecular

interactions while the electrostatic interaction does little. On the other hand, there is a

different trend for the cellulose layers. The van der Waals interaction is nearly one tenth of

the electrostatic interaction. The electrostatic intermolecular interaction (including hydrogen

bonding interaction) plays a primary role in the cellulose layers.

It is previously discussed that the solubility parameter is a function of CED (see

Equation 2.3). However, the solubility parameter has been studied more in the literatures23, 30.

The solubility parameters from references are listed in Table 2.2. The solubility parameters

48

Page 68: Fiber Lubrication: A Molecular Dynamics Simulation Study

from the literature are for bulk polymers. The solubility parameters derived from our work

are usually lower than that of bulk polymers, whether obtained from the experimental or the

simulation. The reason for the difference might be caused by two factors. The first is referred

to as the confined condition. The materials of the confined layers are not continuous in the

confined direction. Therefore, the CED of the confined layer is lower than the bulk material.

A low CED means the low solubility parameter. The other factor might be because the

amorphous layers are built from the oligomers due to the limitation of MD simulation. The

PP and cellulose layers used in experiments are semi-crystalline and made of

macromolecules with high molecular weight. Oligomers have higher freedom than

macromolecules and so the solubility parameter of oligomers is lower than that of

macromolecules. The solubility parameters of confined layers are normal if the former factor

plays a larger role.

Table 2.2 PP and cellulose solubility parameters

Solubility Parameter ((J/cm3)0.5) PP Cellulose

Our value Reference Our value Reference 14.52 16.8-18.8 10 22.15 23.5-28 8

16.8, 16.1 10 31.7, 34.8 10

16.19-19.26 39 34.81 ♠ 16.06 ♠

♠ The values are calculated by the topology method which is explained in Chapter 3.

However, this assumption can not be proven, therefore it is not sure that the latter

factor should not be omitted. Otherwise, the latter factor could particularly influence other

properties derived from the solubility parameter and CED, even though our values of PP and

cellulose solubility parameters are slightly lower than the references. The freedom of the

49

Page 69: Fiber Lubrication: A Molecular Dynamics Simulation Study

layer atoms can be limited by fixing their fractional position, a method which was also used

by Kulmi and Basu40, Chen41, Prathab29, Olgun and Kalyon42. The surface properties of the

confined amorphous layers under fixed atom positions will be discussed by studying contact

angle in section 3.4.

3.3 Volumetric thermal expansion of PP and cellulose amorphous layers

Figure 2.9 displays volumetric thermal expansion of PP and cellulose amorphous

layers obtained from the stepwise temperature procedures. In order to investigate a wide

range of the thermal expansion, we defined the upper temperature at 800 K which is much

higher than the thermal decomposition temperature. The thermal decompositions involve the

valence bond breaking which is beyond MD simulation and so are not shown in Figure 2.7.

As for the PP layer, the thermal volume is nearly linear with increasing temperature. The last

temperature is quite high so that is ignored. The fitting line for volumetric thermal expansion

in the case of the PP is y=13.91x+25757.7, with a coefficient of volumetric thermal

expansion of PP amorphous layers of 13.91 Å3/K. The linearity of the volumetric thermal

expansion indicates that the PP confined layer is amorphous.

50

Page 70: Fiber Lubrication: A Molecular Dynamics Simulation Study

200 300 400 500 600 700 80028000

30000

32000

34000

36000

38000

40000

42000

PP Volume (BatchAvg) Cellulose Volume (BatchAvg)

Vol

ume

(Bat

chA

vg) (

Ang

stro

m3 )

Temp (BatchAvg) (K)

y2=12.77x+31400.1 y1=4.68x+34788.4

(419, 36749)

y=13.91x+25757.7

Figure 2.9 Volumetric thermal expansions of PP and cellulose amorphous layers.

The volumetric thermal expansion for cellulose surfaces is not linear and the fit line is

divided into two parts with different slopes as shown in Figure 2.9. The intersection point is

at (419, 36749), which indicates a phase transition at 419K. The phase transition is most

likely the glass transition. The coefficient of volumetric thermal expansion is 4.68 Å3/K

below 419K and 12.77 Å3/K above 419K, respectively. The cellulose confined layer was

built with the 1,4-βD glucose oligomers and there was no crystal in the structure. The reason

proposed for the transition is that some of hydrogen bonds between cellulose chains are

broken at 419 K.

51

Page 71: Fiber Lubrication: A Molecular Dynamics Simulation Study

3.4 Contact angles

Contact angle is essentially sensitive to the surface energy of a material, and can aid

in the study of other surface properties of the confined amorphous layers. Figure 2.10 shows

the evolution of contact angle on the PP and cellulose confined amorphous surfaces during a

dynamic process.

100ps 200ps 300ps 400ps 500ps

600ps 700ps 800ps 900ps 1ns

(a) The dynamic process of water nano-droplet on the PP surface

100ps 200ps 300ps 400ps 500ps

600ps 700ps 800ps 900ps 1ns

(b) The dynamic process of water nano-droplet on the cellulose surface

Figure 2.10 The dynamic process of contact angle models with PP and

cellulose surfaces.

52

Page 72: Fiber Lubrication: A Molecular Dynamics Simulation Study

The water nano-droplet on a PP surface reaches systemic equilibrium in around 400

ps and the shape dose not significantly change. In contrast, the shape of the water nano-

droplet on a cellulose surface keeps spreading out and does not become stable until 700 ps.

The good wetability of the cellulose is in a good agreement with the fact that cellulose is

hydrophilic. However, there is no wetting phenomenon on the PP surface due to its high

hydrophobicity. Although it takes a period of time (400 ps) to reach equilibrium on the PP

surface, the change in shape is negligible. The droplet on the cellulose surface takes a longer

amount of time to reach equilibrium during the wetting. Also some water molecules can

diffuse into the cellulose surface, possibly entering intermolecular spaces within the

amorphous structure.

The distribution of water molecules on the PP and cellulose surfaces are shown in

Figure 2.11. The water nano-droplet on the PP surface is fairly symmetric in the XYZ system.

In the Z axis direction, most of water molecules are located to the right side of the dashed

line which is defined as the boundary of the PP layer. Very few water molecules diffuse into

the PP amorphous surface. The distribution of the water nano-droplet in the X and Y

direction on the cellulose surface are asymmetric. This is possibly because the intermolecular

space is randomly distributed in the amorphous layer and consequently, the water molecules

randomly penetrate the layer. However, the symmetry in the Z direction corresponds to the

shape of the droplet on the cellulose surface. In the Z axis direction, the water molecules are

distributed well close to the cellulose surface, due to its hydrophilicity. The distribution

curves clearly demonstrate the shape of a water nano-droplet on either the PP or cellulose

surface.

53

Page 73: Fiber Lubrication: A Molecular Dynamics Simulation Study

0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 1500

2

4

6

8 Water at X axis Water at Y axis Water at Z axis

Dis

tribu

tion

on c

ellu

lose

(%)

Distance (Angstrom)

0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 1500

1

2

3

4 Water at X axis Water at Y axis Water at Z axis

Dis

tribu

tion

on P

P (%

)

Distance (Angstrom)

z=26

(a)

z=32(b)

Figure 2.11 The final distribution sketches of water molecules in the lattice. (a) CA model of PP. (b) CA model of cellulose. The final frames are used. The dash lines are the solid surface positions of PP and cellulose respectively.

However, the distribution can not provide the contact angle directly. The following

equations are employed to obtain the information of the water droplet shape:

zrVm Δ∗∗∗=∗=Δ 2πρρ (2.4)

and zmrΔΔ

∗∗

=πρ

1 (2.5)

where r is the radius of the water droplet, ρ is the water density assumed as 1 g/cm3, and

54

Page 74: Fiber Lubrication: A Molecular Dynamics Simulation Study

Δm/Δz is the weight distribution on the Z axis. The shape of the water nano-droplet can be

drawn with r and z, as shown in Figure 2.12.

0 10 20 30 40 50 60 70 80 900

10

20

30

40

50

60

70

80

90

100

110

120

130

140

150 Water droplet on PP

Z ax

is (A

ngst

rom

)

Radius (Angstrom)0 10 20 30 40 50 60 70 80 90

0

10

20

30

40

50

60

70

80

90

100

110

120

130

140

150 Water droplet on cellulose

Z ax

is (A

ngst

rom

)

Radius (Angstrom)

106±4° 33±3°

Figure 2.12 The final shape of water nano-droplet on the PP and cellulose

surfaces. The final frames are used. The dash lines are the solid surface positions of PP and cellulose respectively. The black solid lines are the tangent line of the shape.

Contact angle values were determined by the tangent line of the shapes. The CA

values of PP and cellulose surfaces are 106±4° and 33±3°, respectively. They are in a good

agreement with the experimental results of 102.9±1.7° and 28.6±3.4° for PP and cellulose

thin film, respectively, made from the spin-coating method by Song.1 The agreement between

55

Page 75: Fiber Lubrication: A Molecular Dynamics Simulation Study

the experimental and modeled values first shows that the technique of fixing fractional

position in MD is feasible, and also validates the simulation of the solid surfaces with

oligomers.

4. Conclusion

The amorphous materials were successfully modeled in 1986, but little work involved

in amorphous surfaces. Most of the surfaces reported in the literature are crystal or metal,

which are hard surfaces. However, fiber surfaces are soft and deformable. It is interesting and

challenging to study interface properties in MD simulations. Polypropylene (PP) and

cellulose are two of the most commonly used materials to make fibers. PP and cellulose

confined amorphous surface layers are built successfully. And the physical properties are

inspected using MD simulations, such as the density, cohesive energy, volumetric thermal

expansion, and contact angle of the surfaces. The xenon crystal is the key to compact

polymers. A water nano-droplet is created on the PP and cellulose surfaces to investigate the

contact angles. The shape of the nano-droplet spreads on the cellulose surfaces, and there is

not a significant change in the dimension of the water nano-droplet on the PP surface. The

contact angles of PP and cellulose surfaces are 106° and 33°, respectively. The results prove

that the PP surface is hydrophobic while the cellulose surface is hydrophilic, which is

agreement with previous experimental values.

56

Page 76: Fiber Lubrication: A Molecular Dynamics Simulation Study

References:

1. Song, J. Adsorption of amphoteric and nonionic polymers on model thin films. North Carolina State University, Raleigh, 2008.

2. Israelachvili, J. N., Strength of van der Waals Attraction between Lipid Bilayers. Langmuir 1994, 10, (9), 3369-70.

3. Yoshizawa, H.; Chen, Y. L.; Israelachvili, J., Recent advances in molecular level understanding of adhesion, friction and lubrication. Wear 1993, 168, (1-2), 161-166.

4. Yoshizawa, H.; Israelachvili, J., Fundamental mechanisms of interfacial friction. 2. Stick-slip friction of spherical and chain molecules. J. Phys. Chem. 1993, 97, (43), 11300-13.

5. Li, Y.; Hinestroza, J., Boundary lubircation phenomena in coated textile surfaces. In Friction in textile materials, Gupta, B. S., Ed. Crc; Woodhead: Woodhead publishing in textiles, 2007; pp 419-447.

6. Bezerra, V. B.; Blagov, E. V.; Klimchitskaya, G. L., Some estimations of the atomic friction coefficient by scanning of the AFM tip along a constant force surface. Surface Review and Letters 1999, 6, (3-4), 341-5.

7. Bhushan, B.; LaTorre, C., Investigation of scale effects and directionality dependence on friction and adhesion of human hair using AFM and macroscale friction test apparatus. Ultramicroscopy 2006, 106, (8-9), 720-34.

8. Gulbinski, W.; Pailharey, D.; Suszko, T.; Mathey, Y., Study of the influence of adsorbed water on AFM friction measurements on molybdenum trioxide thin films. Surface Science 2001, 475, (1-3), 149-58.

9. Liang, Q.; Hongnian, L.; Yabo, X.; Xudong, X., Friction and adhesion between C60 single crystal surfaces and AFM tips: effects of the orientational phase transition. Journal of Physical Chemistry B 2006, 110, (1), 403-9.

10. Zhenhua, T.; Bhushan, B., Surface modification of AFM Si3N4 probes for adhesion/friction reduction and imaging improvement. Transactions of the ASME. Journal of Tribology 2006, 128, (4), 865-75.

11. Coffey, T. Nanotribology fundamentals : predicting the viscous coefficient of friction. North Carolina State University, Raleigh, 2004.

12. Coffey, T.; Abdelmaksoud, M.; Krim, J., A scanning probe and quartz crystal microbalance study of the impact of C60 on friction at solid-liquid interfaces. Journal of Physics: Condensed Matter 2001, 13, (21), 4991-9.

13. Coffey, T.; Krim, J., Impact of substrate corrugation on the sliding friction levels of adsorbed films. Physical Review Letters 2005, 95, (7), 076101.

14. Krim, J. In QCM studies of phononic and electronic contributions to sliding friction in adsorbed monolayers, Washington, D.C., United States, 2005; American Society of Mechanical Engineers, New York, NY 10016-5990, United States: Washington, D.C., United States, 2005; pp 489-490.

15. Krim, J., QCM tribology studies of thin adsorbed films. Nano Today 2007, 2, (5), 38-43.

16. Theodorou, D. N.; Suter, U. W., Atomistic modeling of mechanical properties of polymeric glasses.

57

Page 77: Fiber Lubrication: A Molecular Dynamics Simulation Study

Macromolecules 1986, 19, (1), 139-54.

17. Kavassalis, T. A.; Choi, P.; Rundin, A., Calculation of 3D solubility parameters using molecular models. Molecular Simulation 1993, 11, (2-4), 229-241.

18. Choi, P.; Kavassalis, T. A.; Rudin, A., Estimation of Hansen solubility parameters for (hydroxyethyl)- and (hydroxypropyl)cellulose through molecular simulation. Industrial & Engineering Chemistry Research 1994, 33, (12), 3154-3159.

19. Kela, L.; Knuutinen, J.; Linnanto, J.; Suontamo, R.; Peltonen, S.; Kataja, K., Interactions between cationic amylose derivatives and a pulp fiber model surface studied by molecular modelling. Journal of Molecular Structure: THEOCHEM 2007, 819, (1-3), 1-12.

20. Chauve, G.; Heux, L.; Arouini, R.; Mazeau, K., Cellulose poly(ethylene-co-vinyl acetate) nanocomposites studied by molecular modelling and mechanical spectroscopy. Biomacromolecules 2005, 6, (4), 2025-2031.

21. Chen, W.; Lickfield, G. C.; Yang, C. Q., Molecular modeling of cellulose in amorphous state part II: Effects of rigid and flexible crosslinks on cellulose. Polymer 2004, 45, (21), 7357-7365.

22. Chen, W.; Lickfield, G. C.; Yang, C. Q., Molecular modeling of cellulose in amorphous state. Part I: Model building and plastic deformation study. Polymer 2004, 45, (3), 1063-1071.

23. Derecskei, B.; Derecskei-Kovacs, A., Molecular dynamic studies of the compatibility of some cellulose derivatives with selected ionic liquids. Molecular Simulation 2006, 32, (2), 109-115.

24. Misra, S.; Fleming, P. D.; Mattice, W. L., Structure and energy of thin films of poly-(1,4-cis-butadiene): A new atomistic approach. Journal of Computer-Aided Materials Design 1995, 2, (2), 101-112.

25. Natarajan, U.; Tanaka, G.; Mattice, W. L., Atomistic simulations of the surfaces of thin films of random copolymers. Journal of Computer-Aided Materials Design 1998, 4, (3), 193-205.

26. He, D.; Reneker, D. H.; Mattice, W. L., Fully atomistic models of the surface of amorphous polyethylene. Computational and Theoretical Polymer Science 1997, 7, (1), 19-24.

27. Clancy, T. C.; Mattice, W. L., Computer simulation of polyolefin interfaces. Computational and Theoretical Polymer Science 1999, 9, (3-4), 261-270.

28. Hirvi, J. T.; Pakkanen, T. A., Molecular dynamics simulations of water droplets on polymer surfaces. Journal of Chemical Physics 2006, 125, (14), 144712.

29. Prathab, B.; Subramanian, V.; Aminabhavi, T. M., Molecular dynamics simulations to investigate polymer-polymer and polymer-metal oxide interactions. Polymer 2007, 48, (1), 409.

30. Bicerano, J., Prediction of polymer properties. Marcel Dekker: New York, 2002; p 756.

31. Eichinger, B. E.; Rigby, D. R.; Muir, M. H., Computational Chemistry Applied to Materials Design - Contact Lenses. Computational Polymer Science 1995, 5, (4), 147.

32. Hildebrand, J. H., Solubility. Relative values of internal pressures and their practical application. American Ceramic Society -- Journal 1919, 41, (7), 1067-1080.

58

Page 78: Fiber Lubrication: A Molecular Dynamics Simulation Study

33. Martines, E.; Seunarine, K.; Morgan, H.; Gadegaard, N.; Wilkinson, C. D. W.; Riehle, M. O., Superhydrophobicity and superhydrophilicity of regular nanopatterns. Nano Letters 2005, 5, (10), 2097-2103.

34. Lee, H. J.; Michielsen, S., Preparation of a superhydrophobic rough surface. Journal of Polymer Science, Part B: Polymer Physics 2007, 45, (3), 253-261.

35. Michielsen, S.; Lee, H. J., Design of a superhydrophobic surface using woven structures. Langmuir 2007, 23, (11), 6004-6010.

36. http://en.wikipedia.org/wiki/Polypropylene

37. Sun, C. C., Quantifying errors in tableting data analysis using the Ryshkewitch equation due to inaccurate true density. Journal of Pharmaceutical Sciences 2005, 94, (9), 2061-2068.

38. Mazeau, K.; Heux, L., Molecular Dynamics Simulations of Bulk Native Crystalline and Amorphous Structures of Cellulose. The Journal of Physical Chemistry B 2003, 107, (10), 2394-2403.

39. Alan S. Michaels, W. R. V. H. H. A., The solubility parameter of polypropylene. Journal of Applied Polymer Science 1968, 12, (7), 1621-1624.

40. Kulmi, U.; Basu, S., A molecular dynamics study of the failure modes of a glassy polymer confined between rigid walls. Modell. Simul. Mater. Sci. Eng. FIELD Full Journal Title:Modelling and Simulation in Materials Science and Engineering 2006, 14, (6), 1071-1093.

41. Chen, B.; Shi, X.; Gao, H., Apparent fracture/adhesion energy of interfaces with periodic cohesive interactions. Proc. R. Soc. A FIELD Full Journal Title:Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 2008, 464, (2091), 657-671.

42. Olgun, U.; Kalyon, D. M., Use of molecular dynamics to investigate polymer melt-metal wall interactions. Polymer FIELD Full Journal Title:Polymer 2005, 46, (22), 9423-9433.

59

Page 79: Fiber Lubrication: A Molecular Dynamics Simulation Study

Chapter 3

The Structure-property Correlations of PAG Block Copolymer Properties Predicted by Topological Method

Abstract

The topological method of studying structure-property correlations is a simple, quick and valid simulation approach, which can predict physical and chemical properties of polymers from their molecular structures. Two series of PAG block copolymers are selected to be used as lubricants on surfaces: RPnEm diblock copolymers and EmPnEm triblock copolymers. The calculated values of density are higher than the experimental values provided by manufacturers. The zero shear viscosities of PAG block copolymers are linear as a function of molecular weight for those copolymers containing the same hydrophobic weight ratio. The shear modulus of RPnEm polymers decreases nonlinearly with the molecular weight due to the presence of butyl group, while EmPnEm polymers have the nearly identical shear modulus for the copolymers containing the same hydrophobic segment ratio. The cohesive energies of PAG block copolymers increase linearly with molecular weight as well. The solubility of the copolymers is a function of the cohesive energy divided by the molar weight. There is not a significant difference in the solubility parameter of the copolymers with the same hydrophobic segment ratio. The PAG block copolymers with 25% hydrophobic segments show the highest solubility parameter and the RPn polymers illustrate the lowest solubility parameter.

1. Introduction

Poly(alkylene glycol) (PAG)-based block copolymers are widely used as detergents,

emulsifiers, dispersing agents and lubricants in medicine, cosmetic, food, metal and textile

proc.1-5 The poly(ethylene oxide) (PEO) sections are hydrophilic6, 7 while the poly(proplene

oxide) (PPO) sections are hydrophobic. Therefore the PAG-based block copolymers are

usually used as amphiphilic non-ionic surfactants, which make PAG-based surfactants

valuable in the textile industry. Many textile processes, such as dyeing, finishing, and sizing,

60

Page 80: Fiber Lubrication: A Molecular Dynamics Simulation Study

are sensitive to the pH value, and the excess cationic or anionic surfactants could possibly

change processing parameters outside of ideal ranges. In addition there is great influence of

excess ionic surfactants on machinery parts. Non-ionic surfactants can not only overcome the

drawbacks addressed previously, but they are safe to human skin. It is significant to both

manufacturing and end end-users in terms of environmental, health, and safety regulations.

Also, PAG can be manufactured from natural gas with low cost, which is helpful to cut costs

in the textile supply chain. There are two main brands of PAG-based surfactants in the

chemical market: UCON® surfactants manufactured by DOW and Pluronic® surfactants

manufactured by BASF. UCON® is a PPO-PEO diblock copolymers and Pluronic® is a

PEO-PPO-PEO triblock copolymers, shown in Figure 3.1.

(a) (b)

Figure 3.1 Chemical structures of (a) UCON® and (b) Pluronic® polymers. (R refers to the butyl group)

A great deal of experimental and simulation studies have been carried out on these

two copolymers. Although the chemical structures of PEO and PPO with the same repeat

units are similar, except for a methyl group on PPO, their physical properties (e.g. viscosity)

are quite different due to chain length effects.8 Pereira et al.9 studied the UCON® 50-HB-

5100 in a salt solution and stated that the phase configuration was dependent on temperature,

and the phases could be inversed by increasing the temperature to change the densities of the

phases. Olsson et al.10 also studied a mixture of UCON®/dextran/water and agreed that the

molecular weights of copolymers were key factors that influenced the phase separation.

61

Page 81: Fiber Lubrication: A Molecular Dynamics Simulation Study

Pluronic® copolymers were also studied in an aqueous solution. Halacheva et al.11 measured

important parameters, such as the cloud point of Pluronic® copolymers in aqueous solutions.

The most important discovery of their study was that the interactions of the PPO and

Poly(glycidol) blocks with water changed in opposite manners of transmittance vs

temperature curve patterns, which was confirmed with 1H NMR spectroscopy. The

rheological properties of these copolymer aqueous solutions were also investigated by

Halacheva.12 The zero shear viscosity was found to decrease with an increase of temperature

due to the particle size reduction. The storage modulus and loss modulus showed significant

frequency dependence. A series of Pluronic® co-polymers were investigated for their phase

morphology by mesoscale simulation methods: EO13PO30EO13 by Fraaije et al.13,

EO13PO30EO13 and EO19PO33EO19 by van Vlimmeren et al.14, EO26PO40EO26 and

EO99PO65EO99 by Lam and Goldbeck-Wood15, EO13PO30EO13 by Guo et al.16,

EO6PO34EO6, EO13PO30EO13 and EO37PO58EO37 by Li et al.17, EO20PO70EO20 by

Zhao et al.18, EO17PO60EO17, EO19PO43EO19 and EO19PO29EO19 by Yang et al.19.

PAG-based block copolymers are so popular that many application parameters are well

defined. However, little has been reported about the fundamental physical properties in a

comprehensive manner, even from the product specifics of the manufacturers. This chapter

predicted fundamental physical properties of two PAG-based block copolymers equivalent to

the commercial products UCON® and Pluronic®, using the topological method to seek

structure-property correlations.

The topological method has provided structure-property correlations for bulk,

amorphous homopolymers and copolymers in view of thermodynamic, mechanical, and

62

Page 82: Fiber Lubrication: A Molecular Dynamics Simulation Study

transport properties. As opposed to group contributions, which are dependent on the database

of group parameters, the topological method for structure-property correlations employs

connectivity indices in the atom correlations and the database of group contributions is not

required. So far the topological method is developed to predict the properties of the polymers

that are composed of any combination of the following nine elements: carbon, hydrogen,

nitrogen, oxygen, silicon, sulfur, fluorine, chlorine, bromine. The method is suitable and

powerful for predicting the physical properties interested, for the target polymers in this

paper, PAG-based block copolymers.

2. Simulation method and modeling

2.1 Topological method for structure-property correlations

Based on Kier and Hall’s theory of molecular connectivity,20 the topological method

for structure-property correlations was developed by Bicerano21 in 2002.

Building the hydrogen-suppressed graph of the molecule is the first step in the

graphical treatment of molecular properties. A repeat unit is used to represent a molecular

chain in a consistent manner and also to avoid introducing truncation errors. This procedure

is illustrated with the example of the repeat unit of PEO, shown in Figure 3.2(a), together

with the hydrogen-suppressed graph that is obtained in the next step.

63

Page 83: Fiber Lubrication: A Molecular Dynamics Simulation Study

C C

H

O

H

H

H

C C

O

(a) (b) Figure 3.2 The hydrogen-suppressed graph of a molecular repeat unit, using

PEO as an example. (a) Valence bond structure of a PEO repeat unit. (b) Hydrogen-suppressed graph of a PEO repeat unit.

The bonding and electronic environment of each non-hydrogen atom can be

represented by two atomic indices, the simple connectivity index (δ) and the valence

connectivity index (δV). δ is equal to the number of non-hydrogen atoms to which an atom is

bonded. And δV is defined as:21

1−−−

= VH

VV

ZZNZδ (3.1)

where ZV is the number of valence electrons of the atom, NH is the number of hydrogen

bonded to it, and Z is its atomic number. In this paper, two elements, O and C, are involved.

The δ and δV values of O and C are shown in Table 3.1.

Table 3.1 The δ and δV values used in the calculation of PAG polymers 21

Atom Hybridization NH δ δV

3 1 1 2 2 2 1 3 3

sp3

0 4 4 2 1 2 1 2 3

sp2

0 3 4 1 1 3

C

sp 0 2 4 1 1 5 sp3

0 2 6 O

sp2 0 1 6

64

Page 84: Fiber Lubrication: A Molecular Dynamics Simulation Study

Then, bond indices can be derived from the atomic indices. Bond indices (β) and (βV)

are defined as follows.21

jiij δδβ ⋅= (3.2)

And

Vj

Vi

Vij δδβ ⋅= (3.3)

The values of the atomic and bond indices are obtained as shown in the following

example using an EO repeat unit (see Figure 3.3). The simple indices δ and β are shown in

Figure 3.3(a), and the valence indices δV and βV are shown in Figure 3.3(b).

Figure 3.3 Illustrations calculating the simple connectivity index (δ)

and the valence connectivity index (δV). (a) The δ values are shown at the vertices and the β values are shown along the edges. (b) The δV values are shown at the vertices and the βV values are shown along the edges.

All connectivity indices for the polymer chain are obtained and are employed to

calculate the polymer property parameters.

Zeroth-order (atomic) connectivity indices 0χ and 0χV for are defined as follows:21

∑= δχ 10 (3.4)

And

∑= V

V

δχ 10 (3.5)

2 2

2

2 2

4 4 4 4

5

4 10 10 10

(a) (b)

65

Page 85: Fiber Lubrication: A Molecular Dynamics Simulation Study

First-order (bond) connectivity indices 1χ and 1χV are defined as follow:21

∑= βχ 11 (3.6)

And

∑= VV

βχ 11 (3.7)

Extensive properties are correlated directly with the values of χ and depend upon the

size of the system and increase in direct proportion to the amount of material present, such as

molar volume, molar heat capacity, and cohesive energy. Intensive properties are essentially

independent of the amount of materials present. These properties are correlated with values

of χ scaled by the number (N) of non-hydrogen atoms in the system. These scaled values (ξ)

are defined as

Nχξ = (3.8)

Thus, two general forms of correlation are used:21

(Extensive Property)= ∑ + (extensive structural parameters and atomic or group

correction terms) (3.9)

χa

(Intensive Property)= ∑ + (intensive structural parameters and atomic or group

correction terms) + constant (3.10)

ξb

The linear regression coefficients (a and b), correction terms and constant in

Equations 3.9 and 3.10 are adjustable parameters provided by the simulation module,

Synthia®, of Materials Studio 4.1.

All simulation work was performed by two modules of Materials Studio 4.1 software

from Accelrys, San Diego, CA, USA. The MD calculation module used was Discover with

66

Page 86: Fiber Lubrication: A Molecular Dynamics Simulation Study

COMPASS (Condensed-phase Optimized Molecular Potentials for Atomistic Simulation

Studies) as the force field. The group additive method was executed by the Synthia® module.

All calculations were carried out on the Dell® SC1420 workstation with Windows®

server 2003, double Intel® Xeon™ 2.8GHz CPUs, and 4 GB memory.

2.2 Repeat unit of polymers

There are three repeat units used to build PAG block copolymers, butylene,

oxyethylene (EO), and oxypropylene (PO) shown in Figure 3.4. Their atomic parameters are

typed by the COMPASS force field, presented by Sun et al.22, 23 in 1998.

(a) (c)(b)

Figure 3.4 Repeat units to build PAG block copolymers: (a) butylene, (b) oxyethylene (EO), (c) oxypropylene (PO). (Atom colors: carbon atoms—grey, hydrogen—white, and oxygen—red; the head atom is circled by cyan, and the tail atom is circled by magenta).

2.3 Polymer formula and physical properties

There are three kinds of PAG diblock copolymers with the same molecular formula

shown in Figure 3.1(a). They have a different hydrophobic/hydrophilic weight ratio

according to the specifics from the manufacturer. Both butyl groups and PPO are considered

as the hydrophobic segment. The hydrophobic weight ratios (f) are listed in Table 3.2. With

67

Page 87: Fiber Lubrication: A Molecular Dynamics Simulation Study

the given MW of each polymer, the repeat unit numbers, integers m and n, can be obtained as

follow:

roundingunitrepeatEO

given

WMWf

m__

)1( ×−= (3.11)

roundingunitrepeatPO

butylgiven

WWMWf

n__

−×= (3.12)

The repeat unit numbers (m and n) of diblock copolymers which are equivalent to

UCON® products are shown in Table 3.2.

Table 3.2 The list of PAG diblock copolymers equivalent to UCON® products form DOW

Polymers MW♠

(given) MW♣

(calculated) m n Density♥(293K)

Hydrophobic weight ratio♦

f RP6E8 750 775.021 8 6 1.031 50%RP10E13 1230 1227.61 13 10 1.041 50%RP13E17 1590 1578.06 17 13 1.051 50%RP22E29 2660 2629.42 29 22 1.056 50%RP33E44 3930 3929.09 44 33 1.056 50%RP11E42 2470 2507.12 42 11 1.095 25%RP12 740 771.083 0 12 0.983 100%RP24 1420 1468.04 0 24 0.994 100%RP40 2490 2397.32 0 40 1 100%

♠♥ Information from DOW; ♣ The simulation values of the polymers built with m, n; ♦ The MW of butyl group and PO segments over the calculated MW.

As for the PAG triblock copolymers equivalent to Pluronic® polymers, the repeat

unit numbers can be calculated as follow:

roundingunitrepeatEO

given

WMWf

m__2

)1(×

×−= (3.13)

68

Page 88: Fiber Lubrication: A Molecular Dynamics Simulation Study

roundingunitrepeatPO

given

WMWf

n__

×= (3.14)

The repeat unit numbers (m and n) of each triblock copolymer equivalent to

Pluronic® products are shown in Table 3.3.

Table 3.3 The list of PAG triblock copolymers equivalent to Pluronic® products form BASF

Polymers MW♠

(given) MW♣

(calculated) m n Density♥(298K)

Hydrophobic weight ratio♦

f E76P29E76 8400 8398.39 76(2) 29 1.06 25%E103P39E103 11400 11358.1 103(2) 39 1.06 25%E133P50E133 14600 14640.1 133(2) 50 1.06 25%E19P29E19 3400 3376.35 19(2) 29 1.06 50%E26P40E26 4600 4631.97 26(2) 40 1.04 50%E37P56E37 6500 6530.42 37(2) 56 1.05 50%

♠♥ Information from BASF; ♣ The simulation values of the polymers built with m, n; ♦ The MW of butyl group and PO segments over the calculated MW.

3. Results and discussion

3.1 Density of melt PAG block copolymers

An essential density is equal to molecular weight in the molar volume of a molecule.

The molecular weight per polymer is determined exactly from the molecular structure

formula.21 It is important to predict the molar volume to obtain the density. The volume is a

function of temperature. Changes in volume could be caused by molecular configuration. The

topological method for structure-property correlations has three components of the molar

volume: the space truly occupied by atoms, the amount of additional “empty” space, and the

expansion volume. The space truly occupied by atoms is also called van der Waals volume,

which is impenetrable to other molecules with normal thermal energies corresponding to

69

Page 89: Fiber Lubrication: A Molecular Dynamics Simulation Study

ordinary temperatures.24 The amount of additional “empty” space is called packing volume,

which is equal to the difference between the molar volume at absolute zero temperature and

the van der Waals volume. The expansion volume is the volume of thermal motions, which is

equal to the difference between the molar volume at the objective temperature and at absolute

zero. Bicerano21 stated that the standard deviation between the experimental value and the

calculated value was 0.0354 g/cm3 and the correlation coefficient was 0.9872.

Both computational and experimental values of the PAG block copolymers are shown

in Figure 3.5. The calculated values are always higher than the experimental values provided

from the manufacturers. The difference between the experimental and the calculated density

of RP11E42 is the smallest, at 0.0008 g/cm3. The E76P29E76, E103P39E103, and

E133P50E133 have a difference in density of more than 0.44 g/cm3, which is greater than the

standard deviation, 0.0354 g/cm3, obtained by Bicerano.21 The RP6E8, RP10E13, RP13E17,

RP22E29, and RP33E44 have the same hydrophobic weight ratio. Their densities increase

with increasing molecular weight. The increase speed of the experimental values is higher

than that of the calculated. As for RP12, RP24 and RP40 with the 100% hydrophobic weight

ratio, the calculated densities change significantly as a function of molecular weights, while

the experimental densities increase with an increase of molecular weight. There is not a

significant difference in the densities with the variation of molecular weight of EmPnEm

polymers, except for the experimental values of the group of E19P29E19, E26P40E26 and

E37P56E37.

70

Page 90: Fiber Lubrication: A Molecular Dynamics Simulation Study

RP6E8

RP10E13

RP13E17

RP22E29

RP33E44

RP11E42

RP12RP24

RP40

E76P29

E76

E103P

39E10

3

E133P

50E13

3

E19P29

E19

E26P40

E26

E37P56

E37

0.90

0.95

1.00

1.05

1.10

1.15

at 298K (Synthia) at 295K for RPnEm and at 298K for EmPnEm

from manufacturers Molecular weight

Polymers

Den

sity

(g/c

m3 )

0

5000

10000

15000

20000

Mol

ecul

ar w

eigh

t (AM

U)

Figure 3.5 The densities of PAG block copolymers with the experimental

values and the simulation values calculated by the topological method for structure-property correlations.

The polymer density is strongly correlated with the hydrophobic weight ratio of PAG

block copolymers researched in this study. For RPnEm polymers, RP11E42 with 25%

hydrophobic weight ratio has the highest calculated density, 1.0958 g/cm3, while RP12, RP24

and RP40 with a 100% hydrophobic segment has the lowest calculated value at about 1.022

g/cm3. A similar trend is found as for EmPnEm polymers. The calculated densities of

E19P29E19, E26P40E26 and E37P56E37 with 50% PEO segments are lower than those of

71

Page 91: Fiber Lubrication: A Molecular Dynamics Simulation Study

E76P29E76, E103P39E103 and E133P50E133 with 75% PEO segments. It is possibly

because PEO does not have a side group, therefore, has a requests smaller packing volume

than the PPO. Additionally, there are hydrogen bonds in the PEO phases causing the

expansion volume of the PEO to be smaller than that of PPO.

Compared to experimental numbers, the standard deviation of the density calculation

is still high and the correlation coefficient is low. The density from the topological method

for structure-property correlations is derived from the connectivity index of one molecule.

The intermolecular factors are not taken into account. Therefore, the results of calculated

densities are not accurate to predict the density of the PAG block copolymers. It is assumed

that there are self-assembly phenomena25-27 in the block copolymers and the molecular

configurations could expand the molar volume in the bulk. Although the topological method

has limitations, it is still valuable to probe the polymer properties and to establish a

relationship with molecular weight.

3.2 Zero shear viscosity and shear modulus of melt PAG block copolymers

Zero shear viscosity is defined as the melt viscosity of a polymer at the low shear rate

limit. When the temperature is given, the zero shear viscosity is a function of molecular

weight. The results are shown in Figure 3.6, in which the PAG block copolymers are grouped

by the molecular structure and the hydrophobic weight ratio.

In Figure 3.6, the zero shear viscosities are highly linear with molecular weight,

except EmPnEm with a 25% hydrophobic segment. A high molecular weight correlates with

a high zero shear viscosity, according to Figure 3.6. As for the RPnEm polymers, the linear

72

Page 92: Fiber Lubrication: A Molecular Dynamics Simulation Study

slope of RPn polymers is higher than that with a 50% hydrophobic segment. RP11E42 with

25% hydrophobic segments has a lower zero shear viscosity, compared to other polymers

with the similar molecular weight. The linear slope of EmPnEm polymers with 50%

hydrophobic segment is almost equal to that of RPnEm with the same hydrophobic ratio,

which might be caused by the same hydrophobic weight ratio. EmPnEm polymers with 25%

hydrophobic weight ratio have higher zero shear viscosities. In fact, E103P39E103 and

E133P50E133 are not of melt at 298 K, which might result in nonlinearity of the zero shear

viscosity as a function of molecular weight in the case of EmPnEm with 25% hydrophobic

segments.

0 2000 4000 6000 8000 10000 12000 14000 160000

10

20

30

40

50

60

70

80

Zero

she

ar v

isco

sity

(N*s

/m2 )

Molecular weight

RPnEm with 50% hydrophobic segment RPn RP11E42 EmPnEm with 25% hydrophobic segment EmPnEm with 50% hydrophobic segment

Figure 3.6 The zero shear viscosity of PAG block copolymers, calculated

by the topological method for structure-property correlations at 298 K.

73

Page 93: Fiber Lubrication: A Molecular Dynamics Simulation Study

Shear modulus is defined as the ratio of the shear stress to the shear strain. It is a

function of temperature and molecular weight of homopolymers.21 The calculated values of

shear modulus of PAG block copolymers are shown in Figure 3.7. As for RPnEm polymers,

RP11E42 with 25% hydrophobic segment has the highest shear modulus, and the shear

modulus of RPnEm polymers with 50% hydrophobic segment is higher than that of RPn

polymers. In each of groups with different hydrophobic weight ratios, there is a nonlinear

relationship where the shear modulus decreases with increasing molecular weight. This may

be due to the presence of a butyl group present in the hydrophobic segment. The weight ratio

of the butyl group to the molecular weight is not constant, which does not provide a linear

relationship between the modulus and the molecular weight. As for EmPnEm copolymers,

the group with 50% hydrophobic segments has a lower shear modulus than that with 25%.

This may be caused by the different ratios of hydrophobic/hydrophilic segments or/and the

different molecular weights of two groups. There is not a significant difference in the shear

modulus between the polymers with different molecular weights in each EmPnEm polymer

group. The same PPO/PEO weight ratio in groups might contribute to the similar shear

modulus, and the molecular weights of these EmPnEm polymers are constant as a function of

shear modulus.

74

Page 94: Fiber Lubrication: A Molecular Dynamics Simulation Study

0 2000 4000 6000 8000 10000 12000 14000 160000.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

Shea

r mod

ulus

(MPa

)

Molecular weight (AMU)

RPnEm with 50% hydrophobic segment RPn RP11E42 EmPnEm with 25% hydrophobic segment EmPnEm with 50% hydrophobic segment

Figure 3.7 The shear modulus of PAG block copolymers, calculated by the

topological method for structure-property correlations at 298 K.

3.3 Cohesive energy and solubility parameter of melt PAG block copolymers

Cohesive energy is equal to the increase in the internal energy per mole of material,

necessary to eliminate all of the intermolecular forces. The cohesive energy includes three

contributions from dispersion, polar and hydrogen bonding interactions.21 Using Synthia to

calculate cohesive energy, the standard deviation (van Krevelen) is present as low as 1.647

kJ/mol. The correlation coefficient between the calculated and the experimental values is

observed as high as 0.9988.21 The cohesive energies of PAG block copolymers are shown in

Figure 3.8. The cohesive energies of the polymers are grouped by the molecular formula and

the hydrophobic weight ratio linearly increases as a function of molecular weight. There is

75

Page 95: Fiber Lubrication: A Molecular Dynamics Simulation Study

not a significant difference between the slope of RPnEm with 50% hydrophobic segments,

and EmPnEm with 25% and 50% hydrophobic segment. Overall RPn polymers have the

largest slope.

0 2000 4000 6000 8000 10000 12000 14000 160000

200

400

600

800

1000

1200

1400

1600

1800

2000C

ohes

ive

ener

gy (v

an K

reve

len)

at 2

98K

(kJ/

mol

)

Molecular weight (AMU)

RPnEm with 50% hydrophobic segment RPn RP11E42 EmPnEm with 25% hydrophobic segment EmPnEm with 50% hydrophobic segment

Figure 3.8 The cohesive energies of PAG block copolymers, calculated by

the topological method for structure-property correlations at 298 K.

The most important application of cohesive energy is for calculating the solubility

parameter, which quantifies the interactions between polymers and solvents. The solubility of

a polymer is affected by crosslinking, crystallinity, and molecular weight.21 Quantitatively,

the solubility parameter is equal to the square root of the ratio of the cohesive energy to the

molar volume. The calculated results of the cohesive energies are shown in Figure 3.8.

76

Page 96: Fiber Lubrication: A Molecular Dynamics Simulation Study

0 2000 4000 6000 8000 10000 12000 14000 1600015

16

17

18

19

20

Solu

bilit

y pa

ram

eter

(van

Kre

vele

n) a

t 298

K (J

/cm

3 )^0.

5

Molecular weight (AMU)

RPnEm with 50% hydrophobic segment RPn RP11E42 EmPnEm with 25% hydrophobic segment EmPnEm with 50% hydrophobic segment

Figure 3.9 The solubility parameters of PAG block copolymers, calculated

by the topological method for structure-property correlations at 298 K.

Figure 3.9 indicates that the solubility parameter is not sensitive to the molecular

weight. However, the solubility parameter can be divided into three groups with different

hydrophobic weight ratio. RP11E42 has the solubility parameters of 18.68 (J/cm3)0.5, which

is similar to 18.79 (J/cm3)0.5, the value for EmPnEm with 25% hydrophobic segments. Both

RPnEm and EmPnEm with 50% hydrophobic segments have the solubility parameter of

about 18.3(J/cm3)0.5. The solubility parameters of RPn are reduced to about 17.6 (J/cm3)0.5. In

general, solubility parameter is a valid indicator of the solubility in non-aquatic solvents,

such as olefin and benzene. However, it is not valid for water as a solvent, due to the large

quantity of hydrogen bonds between the hydroxyls of the solvents and the PEO segments.

77

Page 97: Fiber Lubrication: A Molecular Dynamics Simulation Study

4. Conclusion

The topological method for structure-property correlations is a simple, quick and

valid simulation approach to predict physical and chemical properties of polymers from their

molecular structures.

The calculated values of PAG block copolymer densities are higher than the

experimental values provided by the manufacturers. This difference might be caused by the

self-assembly phenomena at the mesoscale between different segments, which is not taken

into account in the topological method for structure-property correlations. The zero shear

viscosities of PAG block copolymers are linear as a function of molecular weight within

groups of the same hydrophobic weight ratio, except for EmPnEm with 25% hydrophobic

segments. The shear modulus of RPnEm polymers decreases nonlinearly with the molecular

weight due to the presence of a butyl group, while EmPnEm polymers have a nearly identical

shear modulus within each group of the same hydrophobic segment ratio. The cohesive

energies of PAG block copolymers increase linearly with molecular weight. The solubility is

a function of the cohesive energy over the molar value. The copolymers with the same

hydrophobic segment ratio have the similar solubility parameter. The PAG block copolymers

with 25% hydrophobic segment have the highest solubility parameter, and RPn polymers

have the lowest solubility parameter. But the solubility parameter can not indicate the

solubility of PAG block copolymers in aqueous solvents, due to the large quantity of

hydrogen bonding between the PEO and the hydroxyls of the solvent.

78

Page 98: Fiber Lubrication: A Molecular Dynamics Simulation Study

5. Supplement

It is valuable to obtain the simulation results of a polymeric candidate or objective

before it is planned to be used in industrial or academic research. There are 15 PAG block

copolymers that are candidates. In order to study fiber lubrication with hydrophobic and

hydrophilic surfaces, a polymer chain carrying both hydrophobic and hydrophilic segments is

important to research. Therefore, RPn polymers are not studied further. In addition, the

molecular weight of the E76P29E76, E103P39E103 and E133P50E133 are extremely high

and are not intended to use as fiber lubricants, so they are eliminated from the list as well.

79

Page 99: Fiber Lubrication: A Molecular Dynamics Simulation Study

References:

1. Anon, Some notes on Ucon lubricants. Scientific Lubrication 1966, 18, (5), 31-32.

2. Li, J. T.; Caldwell, K. D.; Rapoport, N., Surface properties of pluronic-coated polymeric colloids. Langmuir 1994, 10, (12), 4475-4482.

3. Oh, S. H.; Kim, J. H.; Kim, J. M.; Lee, J. H., Asymmetrically porous PLGA/Pluronic F127 membrane for effective guided bone regeneration. Journal of Biomaterials Science, Polymer Edition 2006, 17, (12), 1375-1387.

4. Xiong, X. Y.; Tam, K. C.; Gan, L. H., Polymeric nanostructures for drug delivery applications based on pluronic copolymer systems. Journal of Nanoscience and Nanotechnology 2006, 6, (9-10), 2638-2650.

5. Zhao, H.; Totten, G. E.; Webster, G. M.; Tom, F., Poly (alkylene glycol) quenchants for aluminum heat treating. Cailiao Gongcheng/Journal of Materials Engineering 1997, (1), 14-16.

6. Nagahama, K.; Inomata, H.; Saito, S., Measurement of osmotic pressure in aqueous solutions of poly (ethylene glycol) and poly (N-isopropylacrylamide). Fluid Phase Equilibria 1994, 96, 203-214.

7. Ozdemir, C.; Guner, A., Solution thermodynamics of poly(ethylene glycol)water systems. Journal of Applied Polymer Science 2006, 101, (1), 203-216.

8. Forciniti, D.; Hall, C. K.; Kula, M. R., Influence of polymer molecular weight and temperature on phase composition in aqueous two-phase systems. Fluid Phase Equilibria 1991, 61, (3), 243-262.

9. Pereira, M.; Wu, Y.-T.; Madeira, P.; Venancio, A.; Macedo, E.; Teixeira, J., Liquid-liquid equilibrium phase diagrams of new aqueous two-phase systems: Ucon 50-HB5100 + ammonium sulfate + water, Ucon 50-HB5100 + poly(vinyl alcohol) + water, Ucon 50-HB5100 + hydroxypropyl starch + water, and poly(ethylene glycol) 8000 + poly(vinyl alcohol) + water. Journal of Chemical and Engineering Data 2004, 49, (1), 43-47.

10. Olsson, M.; Joabsson, F.; Piculell, L., Particle-induced phase separation in mixed polymer solutions. Langmuir 2005, 21, (4), 1560-1567.

11. Halacheva, S.; Rangelov, S.; Tsvetanov, C., Poly(glycidol)-based analogues to pluronic block copolymers. Synthesis and aqueous solution properties. Macromolecules 2006, 39, (20), 6845-6852.

12. Halacheva, S.; Rangelov, S.; Tsvetanov, C., Rheology of aqueous solutions of polyglycidol-based analogues to pluronic block copolymers. Journal of Physical Chemistry B 2008, 112, (7), 1899-1905.

13. Fraaije, J. G. E. M.; Zvelindovsky, A. V.; Sevink, G. J. A., Computational soft nanotechnology with mesodyn. Molecular Simulation 2004, 30, (4), 225-238.

14. van Vlimmeren, B. A. C.; Maurits, N. M.; Zvelindovsky, A. V.; Sevink, G. J. A.; Fraaije, J. G. E. M., Simulation of 3D mesoscale structure formation in concentrated aqueous solution of the triblock polymer surfactants (ethylene oxide)13(propylene oxide)30(ethylene oxide)13 and (propylene oxide)19(ethylene oxide)33(propylene oxide)19. Application of dynamic mean-field density functional theory. Macromolecules 1999, 32, (3), 646-656.

15. Lam, Y.-M.; Goldbeck-Wood, G., Mesoscale simulation of block copolymers in aqueous solution:

80

Page 100: Fiber Lubrication: A Molecular Dynamics Simulation Study

Parameterisation, micelle growth kinetics and the effect of temperature and concentration morphology. Polymer 2003, 44, (12), 3593-3605.

16. Guo, S. L.; Hou, T. J.; Xu, X. J., Simulation of the phase behavior of the (EO)13(PO)30(EO)13(Pluronic L64)/water/p-xylene system using MesoDyn. Journal of Physical Chemistry B 2002, 106, (43), 11397-11403.

17. Li, Y.; Hou, T.; Guo, S.; Wang, K.; Xu, X., The Mesodyn simulation of pluronic water mixtures using the 'equivalent chain' method. Phys. Chem. Chem. Phys. 2000, 2, (12), 2749-2753.

18. Zhao, Y.; Chen, X.; Yang, C.; Zhang, G., Mesoscopic simulation on phase behavior of pluronic P123 aqueous solution. Journal of Physical Chemistry B 2007, 111, (50), 13937-13942.

19. Yang, S.; Yuan, S.; Zhang, X.; Yan, Y., Phase behavior of triblock copolymers in solution: Mesoscopic simulation study. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2008, 322, (1-3), 87-96.

20. Kier, L. B.; Hall, L. H., Molecular connectivity in structure-activity analysis. Research Studies Press; Wiley: Letchworth, Hertfordshire, England; New York, 1986; p 262.

21. Bicerano, J., Prediction of polymer properties. Marcel Dekker: New York, 2002; p 756.

22. Sun, H.; Ren, P.; Fried, J. R., COMPASS force field: Parameterization and validation for phosphazenes. Computational and Theoretical Polymer Science 1998, 8, (1-2), 229.

23. Sun, H., COMPASS: an ab initio force-field optimized for condensed-phase applications-overview with details on alkane and benzene compounds. Journal of Physical Chemistry B 1998, 102, (38), 7338.

24. van Krevelen, D. W., Properties of polymers : their correlation with chemical structure, their numerical estimation and prediction from additive group contributions. Elsevier: Amsterdam ; New York, 1990; p 875.

25. Schmalz, H.; Knoll, A.; Muller, A. J.; Abetz, V., Synthesis and characterization of ABC triblock copolymers with two different crystalline end blocks: Influence of confinement on crystallization behavior and morphology. Macromolecules 2002, 35, (27), 10004-10013.

26. Tsarkova, L.; Knoll, A.; Krausch, G.; Magerle, R., Substrate-induced phase transitions in thin films of cylinder-forming diblock copolymer melts. Macromolecules 2006, 39, (10), 3608-3615.

27. Tsarkova, L.; Knoll, A.; Magerle, R., Rapid transitions between defect configurations in a block copolymer melt. Nano Letters 2006, 6, (7), 1574-1577.

81

Page 101: Fiber Lubrication: A Molecular Dynamics Simulation Study

Chapter 4

Interaction Energies of PAG Block Copolymer Melts with Polymeric Surfaces in Vacuum

Abstract

Molecular dynamics simulations are used to study the interaction energy of PAG block copolymer melts with PP and cellulose surfaces. The interaction energy is defined as the ratio of interfacial interaction energy to the contact area. Both PPO and PEO segments have lipophilic properties. The molecular weight and the head group weight ratio are found to improve adhesion. Hydrogen bonds may be formed between the PEO and cellulose, resulting in more PEO segments on the cellulose surface than on the PP surface. The interaction energy of PEO on the cellulose surfaces is high. The PPO/PEO ratio and the molecular weight can impact the interaction energy on both PP and cellulose surfaces.

1. Introduction

Interfacial interactions at polymer surfaces are critical in many applications, including

adhesives, coatings, contact lenses, composites, prosthetic devices, films, lubricants, paints,

and printing inks. Interaction energy caused by non-bonding forces between two objects is of

interest in both academic and industrial research.1-5 There are three main components of

intermolecular interaction energy: dispersion interaction of the temporary random dipole,

polar interaction of the permanent dipole, and the hydrogen bonding interaction.6 If the

interaction energy is negative, there is an attractive force between the objects, such as

adhesion. If it is positive, there is a repulsive force. In the majority of engineering polymers,

interfacial interactions at polymer surfaces are rarely measured by experimental methods,

82

Page 102: Fiber Lubrication: A Molecular Dynamics Simulation Study

because usually both the surfaces and polymers involved are insoluble. The surface materials

have high glass transition temperatures, and in some cases it is difficult to precisely

characterize the polymers at the interface.7-9 Therefore, the application of molecular

modeling strategies would be helpful in situations that are difficult to study by experimental

means.10

Researchers are interested in serval critical polymer properties including surface

tension, wetting, adhesion, molecular structure at the interface or inter-phase, and the

structure difference from the bulk. Prathab et al.10 reported the study on interactions of

polymer-to-polymer and polymer-to-metal oxide. They studied poly(methyl methacrylate)

(PMMA) with 10-mers as the matrix, and several metal oxides (Al2O3, Fe2O3, SiO2 and TiO2).

Other engineering polymers were used as the surface materials. The paper provided good

insight into the interactions of PMMA with the polymeric surfaces. Chauve et al.11

established an atomic method to examine the reinforcing effects of nano-composites made of

cellulose whiskers on polyethylene-vinyl acetate (EVA) matrixes. With MD simulations, they

found the differences between the ethylene homopolymer and the polymers with acetate

groups. Dynamic mechanical analysis (DMA) was performed to confirm the results of

adhesion. However, there is little reported work on amorphous cellulose surfaces, which are

commonly seen in the textile, paper and painting industries.

In this chapter, the PP and cellulose amorphous surfaces previously built in Chapter 2

are used. The melt polymers will be studied with PAG block copolymers which are

equivalent to UCON® and Pluronic® products.

83

Page 103: Fiber Lubrication: A Molecular Dynamics Simulation Study

2. Interaction energy and simulation method

2.1 Interaction energy calculation

A general method to calculate the interaction of two materials is shown as Figure 4.1.

First, each component is built as a confined layer with the same length and width at the top

and bottom sides. It is 2-D periodic in the x and y direction. Secondly, the layers are stacked

to form a composite layer. Usually, a vacuum space is adopted on the top of the composite

layer in order to diminish the interaction between the top and the bottom due to the periodic

boundary condition of the composite layer. Then MD simulations are run in NVT ensemble

at a given temperature. The simulation time should be long enough to allow the dynamics

process to reach equilibrium. The last step is that the energy of the entire system as well as

each of components can be obtained with the force field.

Dynamics

(NVT)

(a) (b) (c) (d)

E interaction = E total - ( E surface + E polymer)

Figure 4.1 The calculation of interaction energy for two components. (a) The original model of two-component layer with the vacuum space. The top layer is the liquid polymers (RP6E8) and the bottom layer is the solid surface (PP). (b) The frame after dynamics processes. (c) The surface layer after the removal of the liquid polymers. (d) The liquid layer after the removal of the bottom layer. Etotal, Esurface and Epolymer are the total potential energy of (b), (c) and (d). All are calculated by the force field of Materials Studio 4.1.

84

Page 104: Fiber Lubrication: A Molecular Dynamics Simulation Study

The interaction energy (Einteraction) is calculated by subtracting the energy of each

separated material (Esurface, Epolymer) without interacting surfaces, from that of the total system

(Etotal) with interacting surfaces. In order to obtain the interaction energy per unit, the

modified equation is used as follow:10

VEEE

E polymersurfacetotaladhesion

)( +−= (4.1)

where V is the molar volume of the polymer. Here the term of volume in the equation is not

appropriate because the interaction is considered relative to the area of the interface, instead

of the volume of the polymer. In the Accelrys MD software, the unit of energy is kcal/mol,

where “mol” means a mole number of system molecules in the periodic condition. Therefore,

the area of the interface is a mole number of interface in a single system and the unit of

interaction becomes cal/cm2 (energy over area). In Prathab’s paper,10 there was only one

polymer. If the original polymer confined layer was built in a cube, the area of the interface

could be obtained by the equation3 2VS = . Olgun et al.12 also used the same equation to

calculate the interaction with more than one polymer in their system. However, the molar

volume used in Equation 4.1 can not be used reliably to calculate the interaction energy,

because the interaction energy is only a function of the area of interface, not the volume of

the polymer.

In Chauve’s work, the adhesion at a molecular level corresponds to the energy

difference of the total energy from the addition of surfaces and polymers energies.11

)_2/()( areasurfaceEEEE polymersurfacetotalninteractio ×−−−= (4.2)

Chauve’s interaction equation improves upon Prathab’s work and the unit of interaction is

85

Page 105: Fiber Lubrication: A Molecular Dynamics Simulation Study

mJ/m2 which is energy over area. The minus sign is not be necessary in the equation, because

the simulation system usually provides the positive or negative sign with the value of energy.

Therefore, the following equation is used to calculate the interaction energy in our further

molecular dynamics simulations:

)_/()( areasurfacenEEEE polymersurfacetotalninteractio ×−−= (4.3)

where n is equal to 1 or 2 when the polymer is respectively confined by one or two solid

surfaces. This equation works quite well in a two-component system where there is a

polymer (liquid) and a surface (solid).

2.2 Simulation method

The software and hardware used in the simulations are the same as Chapter 2.

2.1 Model building

The PP and cellulose amorphous surface layers built in Chapter 2 are used in this

work. The dimensions of the PP and cellulose layers are 33×33×26 Å and 33×33×32 Å,

respectively.

Both RPnEm and EmPnEm polymers with 50% hydrophobic weight ratio and

RPnEm with 25% ratio are the target polymers in order to compare results with the results in

Chapter 3. Three additional RPnEm polymers are added in this chapter, RE58, RP32E15 and

RP48, to study the polymeric properties only as a function of hydrophobic weight ratio.

The PAG polymeric confined layers are constructed by Amorphous Cell®, a module

of Materials Studio software. The number of molecules and the layer density are shown in

Table 4.1. The temperature is at 298 K. And the group-based cutoff is adopted when building

86

Page 106: Fiber Lubrication: A Molecular Dynamics Simulation Study

the confined layer. The length and width of the layers are kept constant for all surfaces.

Table 4.1 PAG block copolymers with parameters to build the confined amorphous cells

Polymer MW m_EO n_PO Hydrophobic weight ratio

Number of molecules in one unit

layer

Layer densityg/cm3

RP6E8 775.021 8 6 50% 22 1.03RP10E13 1227.61 13 10 50% 14 1.04RP13E17 1578.06 17 13 50% 11 1.05RP22E29 2629.42 29 22 50% 7 1.06RP33E44 3929.09 44 33 50% 5 1.06RP11E42 2507.12 42 11 25% 7 1.09RE58 2585.14 58 0 2% 7 1.06RP32E15 2593.48 15 32 75% 7 1.06RP48 2803.88 0 48 100% 7 1.06E19P29E19 3376.35 19(2) 29 50% 6 1.06E26P40E26 4631.97 26(2) 40 50% 4 1.06E37P56E37 6530.42 37(2) 56 50% 3 1.06

* New polymers are shadowed by grey.

The original models, similar to Figure 4.1(a), are completed by stacking the polymer

layer on the amorphous surface layer. This is performed by using “building layers” function

of Materials Studio 4.1. A 50Å-thick vacuum layer is inserted on top of the original models to

avoid the border interruption due to the periodic border condition. Therefore, only one

surface is involved and n is equal to 1 in Equation 4.3.

There are three dynamics processes carried out on each model. 5000 fs minimization

is first executed with the medium convergence level and the Smart Minimizer method of

Materials Studio 4.1 software. Then, a 500 ps NVT ensemble of MD simulation at 298 K

with 5 fs equilibrium time is run for each minimized model. Finally, 100 ps NVT ensemble is

performed on the last frame of each model, meanwhile the frames are saved every 10 ps so

that they can be used in the next interaction energy calculation.

87

Page 107: Fiber Lubrication: A Molecular Dynamics Simulation Study

The Accelrys program was modified and customized to calculate the interaction

energy with Equation 4.3, which is attached in Appendix 1.

3. Results and discussion

3.1 RPnEm copolymers

Figure 4.2 shows the interaction energy between RPnEm polymers and the PP and

cellulose surfaces, as well as the molecular weight of RPnEm polymers with 50%

hydrophobic weight ratio. The black cubes with the error bars represent the interaction

energy with the cellulose surfaces, and the red denotes those with the PP surfaces. The

molecular weights are shown with the x axis at the bottom of the figure.

500 1000 1500 2000 2500 3000 3500 4000

-160

-140

-120

-100

-80

-60

-40

Inte

ract

ion

ener

gy (J

/m2 )

Molecular weight (AMU)

Polymers on PP surfaces

Polymers on Cellulose surfaces

Figure 4.2 The interaction energies of melt RPnEm polymers with 50%

hydrophobic weight ratio and different molecular weights on the PP and cellulose surfaces.

88

Page 108: Fiber Lubrication: A Molecular Dynamics Simulation Study

All of the reported interaction energies in Figure 4.2 are negative. Therefore, the

interaction force between RPnEm polymers and the PP and cellulose surfaces is attractive. In

general, an interaction energy with a large absolute value quantifies the strong adhesion.

Cellulose is lipophilic in vacuum and has affinity to both PPO and PEO segments. RP6R8

has the maximum adhesion with both the cellulose and PP surfaces, because the weight ratio

of the butyl groups is 7%, which is higher than any other polymer. The butyl group is

strongly hydrophobic. In addition, there are more hydroxyl groups in the same volume of

RP6E8 because of the low molecular weight and the similar density with other RPnEm

polymers in Figure 4.2. The hydroxyls build strong hydrogen bonds with the cellulose

surface, so the adhesion of RP6E8 with cellulose is stronger than that with PP.

The interaction energy with the cellulose surface has the highest value at RP13E17

while the PP surfaces’ highest value is with RP22E29. The highest interaction energy

correlates to the lowest adhesion value. In Chapter 3, it is shown that RPnEm polymers with

a high molecular weight accordingly have a high viscosity and density, which may contribute

to low adhesion. Therefore, there are two factors which enhance the adhesion, or reduce the

negative value of the interaction energy, of RPnEm with the same hydrophobic weight ratio:

the high weight ratio of head groups (butyl and hydroxyl) or the high molecular weight.

However, these two factors are inversely related to each other. A high weight ratio of head

groups leads to a lower molecular weight and vice versa. Consequently, the RPnEm polymers

with either high head group weight ratio or high molecular weight have strong adhesion to PP

and cellulose surfaces.

Figure 4.3 shows the interaction energies of RPnEm polymers with variable

89

Page 109: Fiber Lubrication: A Molecular Dynamics Simulation Study

hydrophobic weight ratios and the similar molecular weight, which excludes the influence of

the molecular weight, and the weight ratio of butyl and hydroxyl groups. Only the PPO/PEO

ratio is considered.

On the cellulose surfaces, the interaction energy does not change significantly from

RE58 to RP11E42. From RP11E42 to RP48, the interaction energy increases with the

hydrophobic weight ratio. The solubility parameter of the PPO is equal to that of the PEO

when no hydrogen bonding interactions are taken into account. There are both a small

segment of butyl and PEO in the RE58 and RP11E42 samples. The PEO segments play the

main role in contributing to the interaction energy (or adhesion) with cellulose surfaces.

Hydrogen bonding interaction is proportional to the quantity of PEO segments on the

cellulose surfaces. Therefore, RPnEm copolymers with low hydrophobic weight ratio show

good adhesion to the cellulose surfaces.

On the PP surfaces, the effects of the weight ratio of end groups and the molecular

weight are still necessary to consider, but hydrogen bonding interactions are not a factor for

RPnEm polymers. The variable of the PPO/PEO ratio is also important. Although PPO and

PEO have similar solubility parameters, the PPO segment is closer to the surface than the

PEO. RE58 and RP11E42 are mostly comprised of PEO segments, while RP32E15 and RP48

are mostly comprised of PPO segments. Therefore, the interaction energies of RE58 and

RP11E42 are slightly higher than RP32E15 and RP48. RP22E29, with a 50% hydrophobic

weight ratio, has the least adhesion to the PP surface. The reason might be that a 50/50

PEO/PPO segment ratio leads to a nearly symmetric structure, which corresponds to a

characteristic point of low affinity to the PP surfaces.

90

Page 110: Fiber Lubrication: A Molecular Dynamics Simulation Study

0 25 50 75 100-100

-90

-80

-70

-60

-50

-40

-30

-20

Inte

ract

ion

ener

gy (J

/m2 )

Hydrophobicity (%)

Polymers on Cellulose surface

Polymers on PP surface

Figure 4.3 The interaction energies of melt RPnEm polymers with variable

hydrophobic weight ratios and the similar molecular weights on the PP and cellulose surfaces.

3.2 EmPnEm triblock copolymers

The interaction energies of EmPnEm polymers on PP and cellulose surfaces are

shown in Figure 4.4.

91

Page 111: Fiber Lubrication: A Molecular Dynamics Simulation Study

E19P29E19 E26P40E26 E37P56E37-0.050

-0.045

-0.040

-0.035

-0.030

-0.025

-0.020

-0.015

-0.010

-0.005

0.000

Interaction energy with the Cellulose surface Interaction energy with the PP surface Molecular weight

Polymers

Inte

ract

ion

ener

gy (c

al/m

2 )

0

1000

2000

3000

4000

5000

6000

7000

8000

9000

10000

Mol

ecul

ar w

eigh

t (AM

U)

Figure 4.4 The interaction energies of melt EmPnEm polymers with 50%

hydrophobic weight ratios and the similar molecular weights on the PP and cellulose surfaces.

The molecular weights of EmPnEm polymers are relatively high compared to the

RPnEm copolymers previously discussed. The head groups are two hydroxyls that occupy a

small weight fraction in the liquid state. Therefore, the extra influence of the head groups on

the interaction energy can be omitted. The interaction energies of E19P29E19 and

E26P40E26 on PP are slightly higher than those on cellulose surfaces. However, the standard

error is quite large. This indicates the formation of unstable and weak bonds, such as

hydrogen bonds in the system. The interaction energy of the E37P56E37 and the PP surface

is not significantly different from that of the cellulose surface. On the other hand, the

92

Page 112: Fiber Lubrication: A Molecular Dynamics Simulation Study

interaction energy on the PP surface remains nearly constant, while the molecular weight

changes. In the case of the cellulose surface, the lowest interaction energy is found in

E26P40E26, and the highest is seen in E37P56E37. Again, there is a high standard deviation

due to large standard error, which weakens the comparison of the interaction energies of

EmPnEm polymers on cellulose. The high standard deviation is most likely caused by the

small size of MD simulation.

4. Summary

Interaction energy is a valid parameter to study adhesion between the polymer melt

and the solid surface. A program based on the calculation equation 4.3 is used to calculate the

interaction energies of the RPnEm and EmPnEm copolymers on polymeric surfaces. The

results are compared to unveil the impact of molecular weight, hydrophobic weight ratio,

head group weight ratio, and surface nature on the interaction energy. The molecular weight

and the head group weight ratio are found to improve adhesion. Hydrogen bonding can form

between PEO and cellulose, causing more PEO segments to absorb on cellulose surfaces, in

turn increasing the interaction energy.

93

Page 113: Fiber Lubrication: A Molecular Dynamics Simulation Study

References:

1. Costa, A. C.; Composto, R. J.; Vlcek, P.; Geoghegan, M., Block copolymer adsorption from a homopolymer melt to an amine-terminated surface. European Physical Journal E 2005, 18, (2), 159-166.

2. Feng, J.; Ruckenstein, E., Morphology transitions of AB diblock copolymer melts confined in nanocylindrical tubes. Journal of Chemical Physics 2006, 125, (16), 164911.

3. Green, P. F.; Russell, T. P., Adsorption of copolymer chains from a melt onto a flat surface. Macromolecules 1992, 25, (2), 783-787.

4. Kudryavtsev, Y. V.; Govorun, E. N.; Litmanovich, A. D.; Fischer, H. R., Polymer melt intercalation in clay modified by diblock copolymers. Macromolecular Theory and Simulations 2004, 13, (5), 392-399.

5. Kyrylyuk, A. V.; Fraaije, J. G. E. M., Electric field versus surface alignment in confined films of a diblock copolymer melt. Journal of Chemical Physics 2006, 125, (16), 164716.

6. Bicerano, J., Prediction of polymer properties. Marcel Dekker: New York, 2002; p 756.

7. Oslanec, R.; Brown, H. R., Random copolymer adsorption at the polymer melt/substrate interface: Effect of substrate type. Macromolecules 2001, 34, (26), 9074-9079.

8. Oslanec, R.; Vlcek, P.; Hamilton, W. A.; Composto, R. J., Crossover of a block copolymer brush in a polymer melt from a stretched to collapsed conformation. Physical Review E. Statistical Physics, Plasmas, Fluids, and Related Interdisciplinary Topics 1997, 56, (3-A pt A), 2383-2386.

9. Potemkin, I. I.; Kramarenko, E. Y.; Khokhlov, A. R.; Winkler, R. G.; Reineker, P.; Eibeck, P.; Spatz, J. P.; Moeller, M., Nanopattern of diblock copolymers selectively adsorbed on a plane surface. Langmuir 1999, 15, (21), 7290-7298.

10. Prathab, B.; Subramanian, V.; Aminabhavi, T. M., Molecular dynamics simulations to investigate polymer-polymer and polymer-metal oxide interactions. Polymer 2007, 48, (1), 409-416.

11. Chauve, G.; Heux, L.; Arouini, R.; Mazeau, K., Cellulose poly(ethylene-co-vinyl acetate) nanocomposites studied by molecular modelling and mechanical spectroscopy. Biomacromolecules 2005, 6, (4), 2025-2031.

12. Olgun, U.; Kalyon, D. M., Use of molecular dynamics to investigate polymer melt-metal wall interactions. Polymer FIELD Full Journal Title:Polymer 2005, 46, (22), 9423-9433.

94

Page 114: Fiber Lubrication: A Molecular Dynamics Simulation Study

Chapter 5

Configuration and Interaction Energies of a Single PAG Copolymer Molecule in Aqueous Solutions on Fiber Surfaces

Abstract

Poly(alkylene glycol) (PAG) triblock copolymers are widely used as lubricants, surfactants and additives in many industries. In aqueous solutions of PAG copolymers, the interaction between the PAG copolymers and the solid (fibers) surfaces becomes complicated with the presence of water due to the cross interaction of multiple components. The polymer-water-surface (PWS) calculation method is proposed in this chapter to calculate the interaction energy for this complex system. The RP6E8 and E19P29E19 are studied at the water-surface interfaces. Water results in different interaction energy values for the two PAG block copolymers. In contrast with vacuum condition, the presence of water increases the attractive interaction energy of RP6E8 on the cellulose surface, compared with that on the PP surface. However, water decreases the interaction energy of E19P29E19 on the cellulose surface, compared with the PP surface. Additionally, the simulation results are in agreement with the experimental work.

1. Introduction

PAG copolymers are widely used as lubricants, surfactants and additives in many

industries, such as textiles, paper, cosmetics, beverage, food, and painting. PPO and PEO

block copolymers are nonionic, so their pH value in solution has little effect on the solution

properties. This is important in most industrial processes and also beneficial to various end

uses as well. In addition, since PPO segments are hydrophobic and PEO segments are

hydrophilic, they are referred to as amphiphilic polymers. Therefore, water is one of the

commonly used solvent for PAG copolymers. Due to the hydrogen bonding in an aqueous

PAG solution, PEO segments play an important role in the solubility and have drawn a great

95

Page 115: Fiber Lubrication: A Molecular Dynamics Simulation Study

deal of theoretical interest using classic large-scale methods. Using modified small-angle X-

ray equipment Pedersen and Sommer1 applied a random phase approximation to obtain the

virial coefficients and Flory-Huggins parameters of PEO segments in an aqueous solution.

The solubility properties of PAG block copolymers were experimentally studied by Ahmed et

al.2, Bromberg et al.3, Liang et al.4, Oh et al.5, Qiu et al.6, Xiong et al.7 etc. and simulation

methods were employed by Agarwal et al.8, Chen et al.9, 10, Pedersen and Gerstenberg11,

Wijmans et al.12, Sommer et al.13, etc.

When a polymer solution comes into contact with a solid surface, a layer of the liquid

located at the interface behaves differently from the bulk liquid. Although it has been argued

that there are characteristic interactions and structures created by the polymer in solution at

the interface, the interfacial properties are still not fully understood. However, the chemical

and physical properties of a solid surface are factors which determine the interfacial

properties. Equally, the solubility and the self-assembly properties of PAG copolymer

solutions are obviously impacted by the hydrophobicity or hydrophilicity of the surfaces.

Knoll and Magerle14 studied thin films of cylinder-forming polystyrene-polybutadiene-

polystyrene triblock copolymers on polished silicon substrates by scanning probe microscopy.

They found that the structure of thin films depends on the local thickness of the thin film, as

well as the polymer concentrations as well. Tsarkova et al.15 continued Knoll and Magerle’s

work on the carbon substrates and presented similar results. Liang16 used a nanoindenter

instrument to study the lubricity of PAG block copolymers (1% volume) on PP, PE and

cellulose surfaces through nanoscratch tests. Song17 investigated the adsorption of PAG block

copolymers on PP, PET, PE and cellulose surfaces by QCM technique. Li et al.18 adopted a

96

Page 116: Fiber Lubrication: A Molecular Dynamics Simulation Study

lateral force microscopy to measure lateral forces when a tip moved on PP or cellulose

surfaces in an aqueous solution of PAG block copolymers. The previous work shows good

agreement that both PEO-PPO-PEO and butyl-PPO-PEO block copolymers were good

lubricants on PP surfaces, and there is a little adsorption of PAG copolymers in aqueous

solution on cellulose surfaces. Molecular weight, as well as the EO/PO ratio, was found to

play an important role in the polymer-surface interactions. Li et al.18 obtained lateral force

images with different surfaces and proposed a self-assembly hypothesis for PP surfaces

shown in Figure 5.1. PEO segments are assumed to be highly soluble in, and have a good

affinity to water, because of the hydrogen bonding between the oxygen in the ether groups of

PEO and the hydrogen of water. Meanwhile PPO segments are adsorbed to the PP surface

because they are hydrophobic and repellent from water. Due to the limitation of experimental

methods, these hypothesis could not further verified by Liang16, Song17 and Li et al.18

PolypropylenePolypropylene

PEO

PPO

PEO

Water

Figure 5.1 The hypothesis of the EO-PO-EO block copolymer molecule on a PP surface. This figure is revised from the work of Li et al.18

Simulation methods are valid to investigate polymer interfacial phenomena at the

atom scale. Prathab et al.19 and Chauve et al.20 studied the interaction energy between liquid

97

Page 117: Fiber Lubrication: A Molecular Dynamics Simulation Study

homopolymers and solid polymeric surfaces. The liquids in our study are aqueous solutions

of PAG block copolymer. A complicated solution system is created with PPO, PEO and water

on polymer surfaces. Zhange et al.21 adopted a M3B model in molecular dynamics (MD)

simulation to investigate dextran in water on the surface of a bse matrix. Dextran is an α-D-

(1-6)-linked glucan with side chains which are mostly one to two glucose units long. The

system was built with 30×30×100 Å with 30,000 water molecules. All molecule chains were

represented with coarse-grain beads, and the force field only has two terms for the valence

force and one term for the non-bond force. Therefore, Zhang’s model is simplified. We will

use COMPASS as the force field to characterize all atoms in the system. Hower et al.22

studied a complex system of lysozyme above a mannitol self-assembled monolayer surface in

water and other additives. Their monolayer was built with Metropolis Monte Carlo method

and all surface atoms were fixed in the xy plane. Wongkoblap et al.23 focused on the

adsorption of a fluid, such as water, in carbon nanotube bundles. Additional research has

been carried out which studies polymers in water on solid surfaces.24-27 Most of the

previously listed work obtained configurations of macromolecules on solid surfaces by

simulation methods when water was present as the solvent at the atom scale, however little

has been done to calculate the interaction energy between an objective polymer and a surface,

particularly with the presence of water.

In this paper, two PAG block copolymers were placed in water and then confined

with amorphous polymer surfaces of hydrophobic PP and hydrophilic cellulose. The

interaction energy between the polymer and a surface was calculated.

98

Page 118: Fiber Lubrication: A Molecular Dynamics Simulation Study

2. Modeling and methods

Simulation work was performed using two modules of Materials Studio 4.1 software

from Accelrys, San Diego, CA, USA. The MD calculation module was Discover with

COMPASS (Condensed-phase Optimized Molecular Potentials for Atomistic Simulation

Studies) as the force field. A Dell® SC1420 workstation with Window® server 2003, double

Intel® Xeon™ 2.8GHz CPUs, and 4 GB memory was used.

2.1 Modeling

Two block copolymers, RP6E8 and E19P29E19 were used, which are shown in Table

1.1. Their end-to-end distances calculated using a distance measuring tool in Materials Studio

4.1 are 40.7 Å and 164.5 Å after 5000fs minimization.

PP and cellulose surfaces employed here were expanded from the surface layers built

in Chapter 2. Two cells were combined with OA (x axis) orientation by “SuperCell” for

RP6E8, while E19P29E19 was composed of eight cells. The XY dimensions of all cells for

RP6E8 and E19P29E19 were 66.6×33.3 Å and 199.9×33.3 Å, respectively. The thicknesses

(the Z dimension) of the PP and cellulose surfaces were 26.0 Å and 32.3 Å.

Water was defined by the COMPASS force field in Materials Studio 4.1 and designed

as one charge group for each water molecule. A primary layer of water was composed of

1500 molecules, with a dimension of 33.3×33.3×40.4 Å and a density of 1.00 g/cm3. The

confined layer of water used to envelop the RP6E8 molecule was constructed with two

primary layers with the OA (x axis) orientation. Eight primary layers with the same

orientation were used for E19P29E19. The dimensions were 66.6×33.3×40.4 Å for RP6E8

99

Page 119: Fiber Lubrication: A Molecular Dynamics Simulation Study

and 199.9×33.3×40.4 Å for E19P29E19, matching the dimensions of the PP and cellulose

layers created previously in Chapter 2.

A single polymer molecule of RP6R8 or E19P29E19 was placed in confined water

layers with the Perl script from Accelrys Software Inc.28 The geometry center of the polymer

molecule in the water is (0.5, 0.5, 0.3) and the polymer molecules were oriented with the OA

(the X axis).

The original systems of polymer-water-surface were completed by “Build→Build

Layer”, a function in the main menu of Materials Studio 4.1. A vacuum space of 50Å was

added in the original system to avoid the interaction interference due to the periodic

condition. Water was deleted from each system in order to build the single polymer molecule

in a vacuum on the surfaces, and therefore the effect of water on the interactions was

eliminated. The original systems in the chapter are shown in Figure 5.2.

100

Page 120: Fiber Lubrication: A Molecular Dynamics Simulation Study

Figure 5.2 Models of a single PAG molecule affected by water on PP and

cellulose surfaces. Colors represents particular: carbon atoms—grey, hydrogen—white, and oxygen—red. The PAG polymers are colored with different segments: butyl group—purple, PPO— green, and PEO—blue. (a) RP6E8-water-PP system, (b) RP6E8-water-cellulose system, (c) E19P29E19-water-PP system, and (d) E19P29E19-water-cellulose system.

101

Page 121: Fiber Lubrication: A Molecular Dynamics Simulation Study

(a)

PP

(b)

Cellulose

102

Page 122: Fiber Lubrication: A Molecular Dynamics Simulation Study

(c)

PP

(d)

Cellulose

103

Page 123: Fiber Lubrication: A Molecular Dynamics Simulation Study

2.2 Dynamics processes

RP6E8 in a vacuum: 5000fs minimization, 5ps equilibration time, 500ps NVT

dynamics production time. One circle is carried out.

RP6R8 in water: 5000fs minimization, 5ps equilibration time, 500ps NVT dynamics

production time. Two circles are carried out.

E19P29E19 in a vacuum: 5000fs minimization, 10ps equilibration time, 500ps NVT

dynamics production time. Two circles are carried out.

E19P29E19 in water: 5000fs minimization, 10ps equilibration time, 100ps NVT

dynamics production time. Ten circles are carried out.

The frames were saved every 1ps and the last 10 frames were employed to calculate

the average of interaction energy.

2.3 Calculation method of Polymer-Water-Surface (PWS) interaction energy

A polymer-water-surface (PWS) system has a polymer, water and a surfaces as

elements of the system. In the system there can be four cross terms, three double-cross terms

and one triple-cross term. The double-cross terms are EPW, EPS and EWS, and the triple-cross

term is EPWS, as shown in Figure 5.3.

104

Page 124: Fiber Lubrication: A Molecular Dynamics Simulation Study

EW

EP ES

EPW

EPS

EWS

EPWS

Figure 5.3 PWS sketch method to calculate the cross terms of interaction energy in the polymer-water-surface system.

If each term in the PWS system is independent of one another, the single terms (EP,

EW and ES) can be calculated directly with the force field by deleting two other components

from the system, shown in Figure 5.4.

ES

EP

EW

(a) EW (b) EP (c) ES

Figure 5.4 The potential energy of each single component. The white plate represents the removal of a component from the PWS system. Only one component is left and its potential energy can be calculated by the force field.

The double-cross terms (EPW, EWS, EPS) can be easily obtained by deleting one

component and calculating the total energy of each two-component sub system shown in

Figure 5.5.

105

Page 125: Fiber Lubrication: A Molecular Dynamics Simulation Study

EP ES EPS

EW

EP

EPW

EW

ES

EWS

(a) EP+PW+W (b) EW+WS+S (c) EP+PS+S

Figure 5.5 The potential energy of each two-component sub system. The white plate represents the removal of a component from the PWS system.

And the calculation equations of each two-component sub system are as follows:

EPW=EP+PW+W-EP-EW (5.1)

EWS=EW+WS+S-EW-ES (5.2)

EPS=EP+PS+S-EP-ES (5.3)

Finally, the triple-cross term EPWS can be calculated by subtracting each single term

and double-cross terms from the total energy of the whole system (ET):

EPWS=ET-EP-EW-ES-EPW-EWS-EPS

= ET -EP+PW+W -EW+WS+S -EP+PS+S+EP+EW+ES (5.4)

The interaction energy between the polymer and the surface when influenced by

water can be obtained:

EPS_interaction =EPS+EPWS = ET -EP+PW+W -EW+WS+S+EW (5.5)

106

Page 126: Fiber Lubrication: A Molecular Dynamics Simulation Study

Figure 5.6 The sketch of the interaction energy (EPS_interaction)between the

polymer and the surface affected by water.

Equation 5.5 is valid under the assumption that all interaction energy terms are

independent of one another. However, they are dependent of each other in reality. When one

or two components are removed from the system, a new non-bond interaction could result in

the remaining molecules, causing the single terms (EP, EW and ES) and the two-component-

system total energies (EP+PW+W, EW+WS+S and EP+PS+W) to be higher than the actual values in

the PWS system. Due to the cutoff applied when calculating the energy of the system, the

newly created non-bond interactions, such as van der Waals and the dipole-dipole forces, can

be omitted when the atom distance between atoms is longer than 9.5Å. However, the newly

created hydrogen-bonding interactions can not trail off with the cutoff. For example, after

deleting the polymer from the PWS system, two water molecules beside the polymer

molecule may undergo new non-bond interactions, specifically the hydrogen bonding, which

has two orders of magnitude higher than van der Waals force.

If ξ(E) represents the contribution of the new non-bond interactions created when

deleting one or two components, ξ(E) is positive and unknown. Equation 5.5 can be rewritten:

107

Page 127: Fiber Lubrication: A Molecular Dynamics Simulation Study

EPS_interaction = ET -EP+PW+W -EW+WS+S+EW-ξ(EP+PW+W)-ξ(EW+WS+S)+ξ(EW) (5.6)

Equation 5.6 shows the difficulty when obtaining the polymer-surface interaction

energy affected by water. However, there are two inspiring discoveries when ξ(E) was

analyzed: ξ(EP+PW+W) → 0 and ξ(EW+WS+S) ≈ ξ(EW), because the solid polymer surfaces are

nearly flat, new hydrogen bonds are rarely created between water molecules when the solid

surface is deleted from the PWS system, and the removal of polymer causes the majority of

new hydrogen bonds between water molecules.

Therefore, Equation 5.5 is effective for calculating calculate the polymer-surface

interaction energy affected by water. The script has been written to calculate the interaction

of the PWS system and can be seen in Appendix 2.

3. Results and discussion

3.1 Polymer configurations

3.1.1 In a vacuum system

First, we investigated the configuration of the PAG block copolymer molecules in a

vacuum on both the PP and the cellulose surfaces, which are shown in Figure 5.7. The entire

molecule of RP6E8 is very close to both the PP and the cellulose surfaces (see Figure 5.7(a)

and (b)). The butyl groups of RP6E8 even dive into in the PP and cellulose surfaces, and the

PEO and PPO segments display a good affinity to both surfaces. PP belongs to the olefin

family and is obviously lipophilic. Cellulose is a carbohydrate, which is well known as a

hydrophilic material and a good hydrogen-bonding donor. Cellulose is also lipophilic in the

absence of water, which is commonly seen when cotton shirts are easily stained by food oils.

108

Page 128: Fiber Lubrication: A Molecular Dynamics Simulation Study

However, the molecular configurations of RP6E8 on the PP and cellulose surfaces are much

different. The RP6E8 molecule on the PP surface is tangled together, whereas it has a much

more outspreading structure, with a longer end-to-end chain distance on the cellulose surface.

In the top view of Figure 5.7(b), we can see the zigzag shape of the PPO segment of RP6E8

on the cellulose surface, which results in the methyl groups inclining towards the cellulose

surface. The PEO segment also sticks to the cellulose surface and stretches out, which is due

to the hydrophilic affinity, as well as hydrogen bonding between PEO and cellulose.

Therefore, the interaction between RP6E8 and the cellulose surface is stronger than that

between RP6E8 and the PP surface, which is in agreement with the results for RP6E8 in

Chapter 4.

E19P29E19 has almost the same zigzag configuration on both the PP and cellulose

surfaces (see Figure 5.7 (c) and (d)). Although there are a quite few of bends in the PEO

segments, the molecule remains close to the surfaces, showing that the E19P29E19 molecule

has a strong affinity with both fiber surfaces. This is because the hydroxyl sticks to the

cellulose surface due to the hydrogen bonding. The configurations also imply that the

interaction between E19P29E19 and the cellulose surface is somewhat higher than that

between E19P29E19 and the PP surface. However, the interaction increase might be

inconspicuous, since there are only two hydroxyls in one molecule and the molecular weight

of E19P29E19 is distinctly high.

In a vacuum, both the PP and the cellulose surfaces show good adsorption of RP6E8

and E19P29E19. The adsorption is stronger on the cellulose surface when compared to the

PP surfaces, due to hydrogen bonding. Also the surface tension of cellulose is 75.23 dyn/cm,

109

Page 129: Fiber Lubrication: A Molecular Dynamics Simulation Study

whereas the value is only higher than 30.39 dyn/cm for PP, which might contribute to the

affinity of PPO and PEO, which are calculated in Chapter 3 with Synthia® software.

3.1.2 In water

Water is a common solvent for materials with polar groups. PAG block copolymers

are non-ionic surfactants with polar groups and non-polar tails, in which the PPO segment is

the non-polar region and the PEO segment is the polar region. Also there are hydrogen bonds

between PEO and water molecules, even though PEO is lipophilic without water present. The

configurations of RP6E8 and E19P29E19 on PP and cellulose surfaces are impacted by the

strong hydrogen bonding with water.

With water present, the RP6E8 molecule on the PP surface is configured like a

comma, where butyl groups and PPO segments tangle together, and the PEO segment spreads

into the water-PP interface (see Figure 5.8 (a)). The butyl group sticks tightly to the PP

surface and the PPO segment also shows hydrophobic properties and affinities to PP. There is

not water between the hydrophobic regions and the PP surface. However, the hydrophobic

regions contribute greatly to the adsorption of RP6E8 on to PP. The hydroxyl of RP6E8 is

comfortable in water and the PEO segment close to this hydroxyl also expands in water.

Therefore, water molecules can be found between the PEO segments and the PP surface,

causing less affinity between the two components. Under the adsorption competition of water

and the PP surface, PEO definitely prefers staying with water rather than with PP. The butyl

and PPO hydrophobic segments have the potential to escape from water, and increase the

adsorption onto the PP surface.

110

Page 130: Fiber Lubrication: A Molecular Dynamics Simulation Study

The configuration of the RP6E8 molecule in water on the cellulose surface is shown

in Figure 5.5(b), which has several similarities to Figure 5.8(a). The butyl group and PPO

hydrophobic segments curl together and are close to the cellulose surface. Water is not

located between this hydrophobic segment and the cellulose surface. However, the butyl and

PPO segments have an affinity with the cellulose surface. Also the cellulose surface is

hydrophilic in water and lipophilic without water. When water, the butyl group and PPO

segments are present at the same time, the cellulose surface is passive to be captured by them.

The hydrophobicity of the butyl and PPO segment overcomes the hydrophilicity of cellulose

and the cellulose surface has an affinity to the hydrophobic region of RP6E8. Therefore, the

hydrophobic regions have similar configurations on the PP and cellulose surfaces, but the

competition roles of the hydrophobic component and the hydrophilic component are

completely different with the different surfaces. The PEO segments of RP6E8 on the

cellulose surface remain at the interface and are comfortable with both water and the

cellulose surface. Also, there is not an affinity competition between water and cellulose with

PEO. The PEO segments also contribute to the adsorption of water onto the cellulose surface.

The butyl group of RP6E8 has a critical impact on the hydrophobic-hydrophilic

competition of the passive regions. There is not a butyl group in E19P29E19, which is highly

hydrophobic, and the hydroxyls do not play a large role since they occupy only a small ratio

of the molecular weight of E19P29E19. Therefore, the configuration of E19P29E19 is

primarily impacted by the PEO and PPO segments, and the function of the head groups

should be removed.

The zigzag shape of the PPO segment is observed in Figure 5.8(c), in which the

111

Page 131: Fiber Lubrication: A Molecular Dynamics Simulation Study

E19P29P19 is on the PP surface and is affected by water. Some PPO fractions also immerge

with the PP surface, but there is not a water molecule between the polymer and the surface. It

is obvious that there is a strong affinity between PPO and PP when affected by water, due to

the hydrophobic attraction. In the top view of Figure 5.8(c), the straight PPO chain proves to

have a strong affinity with the PP surface again, but the PEO segments curl on the PP surface.

Some PEO fractions ramble into the water from the PP surface, and there is the hydrogen

bonding present between water and the PEO segment. According to the configuration for

E19P29E19, the PPO segment is highly absorbed onto the PP surface, whereas little

absorption of the PEO segments is observed.

The configuration of E19P29E19 on the cellulose surface with water is deceptive (see

Figure 5.8 (d)). The PPO segment is straight with partial zigzag shapes, but quite a few

water molecules are found between the PPO segments and the cellulose surface. The PPO

segment is further away from the surface and folds together with small fractions. It is obvious

that the PPO segment fails to compete with the water molecules to adhere to the cellulose

surface. The PEO segments can be divided into two parts, the dissociative segments in water

and the affinitive segments on the cellulose surface. The former segments are located at the

end which stretches into water, and the later segments are found between the former segment

and the PPO segment. Little water is observed between the affinitive PEO segments and the

cellulose surface. The affinitive PEO segments are the only contributor to the adsorption with

the cellulose surface, and are almost the half of the length of the PEO segment. Also, water

attracts this affinitive segment, so the adsorption of E19P29E19 on the cellulose surface is

not strong, when compared with the three above configurations.

112

Page 132: Fiber Lubrication: A Molecular Dynamics Simulation Study

3.2 The interaction energy of the PWS system

The interaction energies obtained with the PWS interaction energy calculation

method are shown in Table 5.1. The negative interaction energy between two objects is an

attractive interaction force.

Table 5.1 The interaction energies between the polymer and the surface affected by water Polymer RP6E8 E19P29E19 Medium vacuum water vacuum water Surface PP Cellulose PP Cellulose PP Cellulose PP CellulosePS_int

(kcal/mol) -31.3389 -54.8726 -38.0548 -65.234 -145.978 -154.664 -154.847 -101.158

Std.Dev 4.94 11.69 4.76 13.43 10.31 29.24 n/a n/a In a vacuum system, both polymers have a higher attractive energy on the cellulose

surface than on the PP surface, due to the hydrogen bonding between the PEO and the

cellulose hydroxyls. The butyl group leads to a 70% higher attractive energy with cellulose as

opposed to PP, because cellulose is lipophilic without water. Due to the limited number and

MW ratio with E19P29E19, head groups and hydrogen bonding do not greatly contribute to

the attractive interaction energy. Therefore, the interaction energy of E19P29E19 on cellulose

is not more than 10% higher than that on PP. These results reasonably support the

explanation in Section 3.1.

In water, RP6E8 and E19P29E19 have a different trend. Water induces a higher

attractive interaction energy between RP6E8 and the cellulose surface than on PP, but a lower

interaction energy between E19E29E19 and the cellulose surface than on PP. Upon analysis

of the configurations in 3.2, the butyl group is more competitive than the water molecules on

the cellulose surface. Also, the PEO segment is comfortable with the cellulose surface.

113

Page 133: Fiber Lubrication: A Molecular Dynamics Simulation Study

However, E13P29E19 does not have a high hydrophobic head group and the PPO segment

fails to compete with water to adsorb onto the cellulose surface. On the contrary, water

increases the affinity between the PPO segment and the PP surface. These results reasonably

support the explanation in Section 3.2.

4. Conclusion

The new PWS interaction energy calculation method was proposed in this chapter.

The calculation results are in good agreement with the configuration results, proving that this

calculation method is effective and can be applied to complicated PWS systems.

In a vacuum, the interaction energies of RP6E8 and E19P29E19 are attractive, and

they are stronger on the cellulose surface than that on the PP surface, due to the hydrogen

bonding. The butyl group and the PEO segment play a critical role in the hydrophobic-

hydrophilic competition of water and the surfaces. E19P29E19 is not greatly affected by the

head groups, and the interaction energy of the cellulose surface is just a small amount larger

than that of the PP surface.

Water results in different interaction energies of both PAG block copolymers. The

presence of water increases the attractive interaction energy of RP6E8 on the cellulose

surface, compared with that on the PP surface. However, water decreases the interaction

energy of E19P29E19 on the cellulose surface, compared with that on the PP surface. These

results were also observed by Liang16 and Song17.

114

Page 134: Fiber Lubrication: A Molecular Dynamics Simulation Study

Figure 5.7 The configurations of the RP6E8 and E19P29E19 molecules in vacuum on the PP and cellulose surfaces. (a) RP6E8 on the PP surface, (b) RP6E8 on the cellulose surface, (c) E19P29E19 on the PP surface, (d) E19P29E19 on the cellulose surface. All images are the final frames of the NVT dynamics simulations. The color sets are the same with Figure 5.2.

115

Page 135: Fiber Lubrication: A Molecular Dynamics Simulation Study

(a) (b)

116

Page 136: Fiber Lubrication: A Molecular Dynamics Simulation Study

(c)

117

Page 137: Fiber Lubrication: A Molecular Dynamics Simulation Study

(d)

118

Page 138: Fiber Lubrication: A Molecular Dynamics Simulation Study

Figure 5.8 The configurations of the RP6E8 and E19P29E19 molecules in water on the PP and cellulose surfaces. (a) RP6E8 on the PP surface, (b) RP6E8 on the cellulose surface, (c) E19P29E19 on the PP surface, (d) E19P29E19 on the cellulose surface. All images are the final frames of the NVT dynamics simulations. In each top view of the images, the water molecules are invisible for clarity to the PAG polymer molecules. The color sets are the same with Figure 5.2.

119

Page 139: Fiber Lubrication: A Molecular Dynamics Simulation Study

(a)

(b)

120

Page 140: Fiber Lubrication: A Molecular Dynamics Simulation Study

(c)

121

Page 141: Fiber Lubrication: A Molecular Dynamics Simulation Study

(d)

122

Page 142: Fiber Lubrication: A Molecular Dynamics Simulation Study

References:

1. Pedersen, J. S.; Sommer, C., Temperature dependence of the virial coefficients and the chi parameter in semi-dilute solutions of PEG. Prog. Colloid Polym. Sci. 2005, 130, 70-78.

2. Ahmed, F.; Alexandridis, P.; Neelamegham, S., Pluronic block copolymers inhibit platelet aggregation: Role of critical micelle concentration & side chain length. Annual International Conference of the IEEE Engineering in Medicine and Biology - Proceedings 1999, 2, 723.

3. Bromberg, L.; Alakhov, V. Y.; Hatton, T. A., Self-assembling Pluronic [registered trademark] -modified polycations in gene delivery. Current Opinion in Colloid and Interface Science 2006, 11, (4), 217-223.

4. Liang, X.; Mao, G.; Ng, K. Y. S., Effect of chain lengths of PEO-PPO-PEO on small unilamellar liposome morphology and stability: An AFM investigation. Journal of Colloid and Interface Science 2005, 285, (1), 360-372.

5. Oh, S. H.; Kim, J. H.; Kim, J. M.; Lee, J. H., Asymmetrically porous PLGA/Pluronic F127 membrane for effective guided bone regeneration. Journal of Biomaterials Science, Polymer Edition 2006, 17, (12), 1375-1387.

6. Qiu, Y.-R.; Rahman, N. A.; Matsuyama, H., Preparation of hydrophilic poly(vinyl butyral)/Pluronic F127 blend hollow fiber membrane via thermally induced phase separation. Separation and Purification Technology 2008, 61, (1), 1-8.

7. Xiong, X. Y.; Tam, K. C.; Gan, L. H., Synthesis and thermally responsive properties of novel pluronic F87/polycaprolactone (PCL) block copolymers with short pcl blocks. Journal of Applied Polymer Science 2006, 100, (5), 4163-4172.

8. Agarwal, A.; Unfer, R.; Mallapragada, S. K., Investigation of in vitro biocompatibility of novel pentablock copolymers for gene delivery. Journal of Biomedical Materials Research - Part A 2007, 81, (1), 24-39.

9. Chen, S.; Guo, C.; Hu, G.-H.; Liu, H.-Z.; Liang, X.-F.; Wang, J.; Ma, J.-H.; Zheng, L., Dissipative particle dynamics simulation of gold nanoparticles stabilization by PEO-PPO-PEO block copolymer micelles. Colloid and Polymer Science 2007, 285, (14), 1543-1552.

10. Chen, S.; Hu, G.-H.; Guo, C.; Liu, H.-Z., Experimental study and dissipative particle dynamics simulation of the formation and stabilization of gold nanoparticles in PEO-PPO-PEO block copolymer micelles. Chemical Engineering Science 2007, 62, (18-20 SPEC ISS), 5251-5256.

11. Pedersen, J. S.; Gerstenberg, M. C., The structure of P85 Pluronic block copolymer micelles determined by small-angle neutron scattering. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2003, 213, (2-3), 175-187.

12. Wijmans, C. M.; Eiser, E.; Frenkel, D., Simulation study of intra- and intermicellar ordering in triblock-copolymer systems. Journal of Chemical Physics 2004, 120, (12), 5839-5848.

13. Sommer, C.; Pedersen, J. S.; Stein, P. C., Apparent specific volume measurements of poly(ethylene oxide), poly(butylene oxide), poly(propylene oxide), and octadecyl chains in the micellar state as a function of temperature. Journal of Physical Chemistry B 2004, 108, (20), 6242-6249.

123

Page 143: Fiber Lubrication: A Molecular Dynamics Simulation Study

14. Knoll, A.; Magerle, R.; Krausch, G., Phase behavior in thin films of cylinder-forming ABA block copolymers: Experiments. Journal of Chemical Physics 2004, 120, (2), 1105-1116.

15. Tsarkova, L.; Knoll, A.; Krausch, G.; Magerle, R., Substrate-Induced Phase Transitions in Thin Films of Cylinder-Forming Diblock Copolymer Melts. Macromolecules 2006, 39, (10), 3608-3615.

16. Liang, J. Investigation of synthetic and natural lubricants. North Carolina State University, Raleigh, 2008.

17. Song, J. Adsorption of amphoteric and nonionic polymers on model thin films. North Carolina State University, Raleigh, 2008.

18. Li, Y.; Hinestroza, J., Boundary lubircation phenomena in coated textile surfaces. In Friction in textile materials, Gupta, B. S., Ed. Crc; Woodhead: Woodhead publishing in textiles, 2007; pp 419-447.

19. Prathab, B.; Subramanian, V.; Aminabhavi, T. M., Molecular dynamics simulations to investigate polymer-polymer and polymer-metal oxide interactions. Polymer 2007, 48, (1), 409-416.

20. Chauve, G.; Heux, L.; Arouini, R.; Mazeau, K., Cellulose poly(ethylene-co-vinyl acetate) nanocomposites studied by molecular modelling and mechanical spectroscopy. Biomacromolecules 2005, 6, (4), 2025-2031.

21. Zhang, X.; Wang, J. C.; Lacki, K. M.; Liapis, A. I., Construction by molecular dynamics modeling and simulations of the porous structures formed by dextran polymer chains attached on the surface of the pores of a base matrix: Characterization of porous structures. Journal of Physical Chemistry B 2005, 109, (44), 21028-21039.

22. Hower, J. C.; He, Y.; Jiang, S., A molecular simulation study of methylated and hydroxyl sugar-based self-assembled monolayers: Surface hydration and resistance to protein adsorption. Journal of Chemical Physics 2008, 129, (21), 215101.

23. Wongkoblap, A.; Do, D. D.; Wang, K., Adsorption of polar and non-polar fluids in carbon nanotube bundles: Computer simulation and experimental studies. Journal of Colloid and Interface Science 2009, 331, (1), 65-76.

24. Imai, T.; Hiraoka, R.; Kovalenko, A.; Hirata, F. In Solvation structure at the protein-water interface studied by a statistical-mechanical theory of solutions, Tokyo, 104-0042, Japan, 2005; Society of Polymer Science: Tokyo, 104-0042, Japan, 2005; pp 3881-3882.

25. Fritz, L.; Hofmann, D., Molecular dynamics simulations of the transport of water-ethanol mixtures through polydimethylsiloxane membranes. Polymer 1997, 38, (5), 1035-1045.

26. Fritz, L.; Hofmann, D., Behaviour of water/ethanol mixtures in the interfacial region of different polysiloxane membranes - a molecular dynamics simulation study. Polymer 1998, 39, (12), 2531-2536.

27. Gambogi, J. E.; Blum, F. D., Effects of water on the interface in a model polymer composite system: A nuclear magnetic resonance study. Materials Science & Engineering A: Structural Materials: Properties, Microstructure and Processing 1993, A162, (1-2), 249-256.

28. Accelrys http://forums.accelrys.org/eve/forums/a/tpc/f/3091072081/m/8661038581

124

Page 144: Fiber Lubrication: A Molecular Dynamics Simulation Study

Chapter 6

Mesoscale Simulations of PAG Triblock Copolymer Morphology

Abstract

The MesoDyn method is adopted to investigate the morphology of a PAG triblock copolymer, E19P29E19, in bulk solutions and on the solid-liquid interfaces. Two conditions on the solid-liquid interface are considered for triblock copolymers: the confined and the confined with shearing. The E19P29E19 chains are formed to unique self-assembled structures which are dependent on the solution concentration. In a bulk aqueous solution, when the concentration is 10% v/v, micelles exhibit in a fashion of PPO blocks inside and PEO blocks outside. While the concentration increases to 25% and 50%, the worm-like micelles and irregular cylinders are observed in the solutions, respectively. The core radius calculated for the micelles with 10% E19P29E19 is 31.1Å, and the aggregation number of N is around 46. The formation of micelles occurs during the 1600 time step to the 5700 time step. The micelles are also found in aqueous solutions when the solutions are confined by cellulose surfaces. The micelle formation under the confined condition is faster than that in the bulk solution. At a shear rate of 0.001ns-1, the micelles are deformed and eventually broken due to the shearing, resulting in the formation of a wall-like micelle. The wall-like micelle structure increases the viscosity of the solution. During the shearing of the PP surfaces, the E19P29E19 solutions become adsorbed onto the surfaces, and are able to prevent the surfaces from damage due to frictional forces.

1. Introduction

There are two commonly used mesoscale simulation methods: mesoscopic dynamics

(MesoDyn) and dissipative particle dynamics (DPD), which have been developed in the past

two decades. Both approaches treat the polymer chains at the coarse-grained level by

grouping atoms together up to the equivalent length of the polymers. This treatment can

extend the length of the polymer chains and also increase the time scale of simulation by

many orders of magnitude in companion to the atomistic simulations. MesoDyn is based on

the dynamic mean field density functional theory.1, 2 The phase separation dynamics are

125

Page 145: Fiber Lubrication: A Molecular Dynamics Simulation Study

described by Langevin-type equations for polymer diffusion, and the thermal fluctuations are

added as noise (random force). MesoDyn has been successfully applied to concentrated

aqueous solutions of PEO-PPO-PEO triblock copolymers. Meanwhile, in the DPD method,

particles are subject to conservative, dissipative, and random forces. These forces are

calculated from pairs of neighboring particles.3 In contrast to classical MD, the DPD method

employs soft interaction potentials, as a natural consequence of the coarse-grained

representation, allowing large time scale simulations.1 The time evolution of the system is

obtained by solving Newton’s equations of motion. The method leads to a correct description

of Navier-Stokes hydrodynamic behavior.

MesoDyn was originally developed by Akzo Nobel in the early 1990s, in an attempt

to solve a stability problem in waterborne coatings4. Since then, it has been greatly expanded

and applied to many different systems and external conditions, specially in two large

industrial European research projects, CAESAR and MesoDyn, led by BASF.5 Many

copolymers produced by BASF were investigated, EO13PO30EO13 by Fraaije et al.5,

EO13PO30EO13 and EO19PO33EO19 by van Vlimmeren et al.6, EO26PO40EO26 and

EO99PO65EO99 by Lam and Goldbeck-Wood7, EO13PO30EO13 by Guo et al.8, EO6PO34EO6,

EO13PO30EO13 and EO37PO58EO37 by Li et al.9, EO20PO70EO20 by Zhao et al.10,

EO17PO60EO17, EO19PO43EO19 and EO19PO29EO19 by Yang et al.11. Other polymers were also

studied, such as polyamide 6 (PA6) with poly(vinyl alcohol) (PVOH), poly(vinyl acetate)

(PVAC), and partially hydrolyzed PVAC12, poly (vinyl alcohol) with poly (methyl

methacrylate)13, poly(vinyl pyrrolidone) with poly(bisphenol-A-ether sulrone)14.

The methodology of mesoscale simulation softwares can be divided into two

126

Page 146: Fiber Lubrication: A Molecular Dynamics Simulation Study

categories. The first category consists of modeling problems, i.e. finding a mathematical

model that represents the most important physical features of the mesoscale structures. The

second category consists of algorithmic and implementation related issues, i.e. given a

certain model, finding the algorithms that solve the discrete problems accurately and

efficiently.15 Our focus is on the first category in our work.

Finally, two parameters are critical to model the physical features of the mesoscale

properties: the chain topology and interaction parameters.

1.1 Chain topology

The chain topology depends on the degree of coarseness of the original system. For

each system, there is a specific characteristic length that is related to the Gaussian chain

length of a component, which is similar to the Kuhn length.7 One constrain is only one bead

size, i.e. the same length of sections. The other constrain is that the end-to-end distance

squared of the Gaussian chain and atomistic chain should be the same, which is called the

equivalent chain method.9

For the ExPyEx type of triblock copolymers, the end distribution can be calculated

with the characteristic ratios as

)2()2(2 22222COCC

POCOCC

EO llyCllxCr +++= ∞∞ (6.1)

and the length of the backbone is

(6.2) )2)(2( COCCi llyxl ++=∑ The reported Gaussian chains of triblock copolymer were shown in the following

table.

127

Page 147: Fiber Lubrication: A Molecular Dynamics Simulation Study

Table 6.1 Topology parameters for MesoDyn Molecular formula Gaussian Chain EO6PO34EO6 E3P18E3 9 EO13PO30EO13 E7P16E7 8, 9 E3P9E3 5, 6 EO17PO60EO17 E4P18E4 11 EO19PO29EO19 E4P9E4 11 EO19PO33EO19 E6P8E6 6 EO19PO43EO19 E4P13E4 11 EO20PO70EO20 E5P21E5 10 EO26PO40EO26 E6P12E6 7 EO37PO56EO37 E20P31E20 9 EO99PO65EO99 E24P20E24 7

There are two different Gaussian chains reported for EO13PO30EO13. We found that

Fraaije and van Vlimmeren defined the beads to be twice larger than that of Gou and Li, so

that the molecular volume should be the same.

1.2 Interaction parameters of PEO, PPO and water

Another critical parameter of MesoDyn is the interaction parameter representing the

pairwise interactions of beads similarly to what is defined in the Flory-Huggins model. For

non-polar polymers, the interaction parameter can be simply obtained from the component

solubility parameters as

RTVspssp 2)( δδχ −= (6.3)

where δs, δp are Scatchard-Hildebrand solubility parameters of the solvent and polymer,

respectively, and Vs is the molar volume of the solvent. Equation 6.3 has been adopted by

Yuan et al.16 , Li et al.9, Zhang et al.17, Yang et al.18, Lem et al.7 etc. They calculated the

interaction using only one pair of components and omitted the cross interaction of multiple

components as addressed in Chapter 5. However, this approach is not valid for polar and

hydrogen bonding systems, such as a water-PEO system. Therefore, a different approach is

128

Page 148: Fiber Lubrication: A Molecular Dynamics Simulation Study

proposed to calculate the interaction parameter from vapor pressure data.

⎭⎬⎫

⎩⎨⎧

−−−−= − φφφχ )11()1ln(ln 02

Npp

IJ (6.4)

where p is the vapor pressure, φ is the polymer volume fraction, and N is the chain length,

which is the number of monomers per bead.

The interaction parameters of PEO-water and PPO-water were previously reported in

experimental data as 1.4 and 1.7, respectively.2, 5-7, 9 Fraaije et al.2 point out that the PEO-

water interaction parameter was surprisingly high. Guo et al.8 also claimed χEW=1.4 was

apparently not in agreement with the experimental results. Pedersen and Sommer19 published

their new experimental data of the chi parameter, χEW=0.35, in semi-dilute solutions of PEO

at 20˚C. Guo et al. used χEW=0.3,8 which was in closer agreement with Pedersen and

Sommer’s experimental results.19 Regarding the interaction parameter of PEO-PPO, the

previous works found that all experimental data were unreliable. Fraaije et al.5 remarked that

PEO was a unique polymer because it is soluble in water even though it is a member of an

insoluble homologous series. Recent comparison of molecular dynamics simulation

discovered a strong influence in hydrogen bond formation, due to the competition between

water-water and PEO-water non-bonds, making it challenging to calculate the Flory-Huggins

parameters from the first principle. The interaction parameter of PEO-PPO was roughly

estimated as a value between 3 and 5 by group contribution method.2, 8-11 The interaction

parameters reported are shown in the following table.

129

Page 149: Fiber Lubrication: A Molecular Dynamics Simulation Study

Table 6.2 Interaction parameters for MesoDyn References χEW χPW χEP

Fraaije 5 1.4 1.7 3.0 van Vlimmeren 6 1.4 1.7 3.0 Zhao 10 1.4 1.7 3.0 Yang 11 1.4 1.7 3.0 Lam 7 1.35 1.64 5.0 Li 9 1.4 1.7 4.0 Guo 8 0.3 2.1 3.0

The interaction parameter can be obtained from atomic simulations, except for

receiving the interaction parameter values from experimental and empirical solubility

parameters. Spyriouni and Vergelati12 studied the binary blend compatibility of PA6, PVAC

and partially hydrolyzed PVAC by atomistic and mesoscopic simulation. An equation was

provided in the atomistic method to obtain the Flory-Huggins parameter as

monmix V

RTE

⎟⎠⎞

⎜⎝⎛ Δ=χ (6.5)

where Vmon is a monomer unit volume which is set equal to the monomer volume for each

blend. ΔEmix for a binary mixture means a variation of the cohesive energy before and after

mixture. It is defined as

( ) ( ) ( )mixBBAAmix CEDCEDCEDE −+=Δ ϕϕ (6.6)

where φA and φB are the monomer volume fractions of A and B. A 200-250 ps NVT ensemble

was employed to equilibrate the system, resulting in a proper density. The reported

interaction parameter results were in agreement with values obtained in similar studies.

However, this method has not been used in an aqueous liquid system because there are strong

hydrogen bonds. Even the developers of MesoDyn method, Fraaije et al. , used only the

empirical values of the Flory-Huggins parameters on the pairs of PEO-PPO, PEO-water and

B

5

130

Page 150: Fiber Lubrication: A Molecular Dynamics Simulation Study

PPO-water. Equations 6.5 and 6.6 were used in an aqueous system by Yuan et al. , Li et al. ,

Zhang et al. , Yang et al. etc. They calculated the interaction by only one pair of

components and neglected the influence of the aqueous solvent.

16 9

17 18

In order to create a confined liquid between two solid layers in the MesoDyn model,

we need to first determine the Flory-Huggins parameters between the solid layer and each

component Knoll et al.20 experimentally investigated the morphology of a polystyrene-

polybutadiene-polystyrene triblock copolymer in chloroform, on polished silicon substrates

by scanning force microscope (SFM). In addition, they used MesoDyn to simulate the

triblock polymer system and found the simulation results in good agreement with the

experimental results.21 Hovat et al.22 and Lyakhova et al.23 also proved Knoll’s simulation

results with MesoDyn simualtions. The Flory-Huggins interaction parameters in their system

were cited directly from Fraaije’s work2, where the Flory-Huggins parameters were

calculated by Equation 6.3. The system of polystyrene-polybutadiene-polystyrene triblock

copolymer in chloroform on silicon was not aqueous and so Equation 6.3 worked well.

Similar research on aqueous systems has not been conducted, probably because of the

difficulty acquiring interaction parameters between multiple components in the aqueous

systems with the presence of water in an aqueous system.

2. Simulation parameters

In this chapter, we focus on E19P29E19, a PAG triblock copolymer with the 50/50

weight ratio of PPO and PEO segments. The chain topology is selected as E4P9E4, the same

as used by Yang et al.11 Yang et al. used the Flory-Huggins interaction parameters that were

131

Page 151: Fiber Lubrication: A Molecular Dynamics Simulation Study

also adopted by Fraaije et al. 1, 2, 5, van Vlimmeren et al.6, Zhao et al.10 and Li et al.9

However, they did not observe the formation of a micelle structure until the E19P29E19

concentration was increased to 43% v/v, which did not agree with the experimental value.

Usually micelles are formed below 2% w/v at 40 °C24 (The densities of water and

E19P29E19 are very close to 1 g/cm3, so v/v is close to w/v). We adjusted the interaction

parameters properly in order to match the experimental data. The interaction parameters of

PEO-water and PPO-PEO, χEW=1.4 and χEP=4.0, are selected from Table 6.2, which are the

same as that of Li et al.9 The interaction parameter of PPO-water is not consistent with the

previous reports. In fact, PPO is well known to be hydrophobic. The PEO and PPO are

compatible in the absence of water. Therefore, we consider that PPO has a greater affinity to

PEO than to water. Accordingly, the interaction parameter of PPO-water was selected as

χPW=5.0.

The interaction parameters between surfaces and polymers are difficult to obtain due

to the cross interaction between multiple components (discussed in Chapter 5). PWS systems

were built in Chapter 5, and the interaction energy of PAG block copolymers and two

surfaces with the presence of water were studied. We adopted the interaction parameters as

follows:

PP surface: χP_PP < χE_PP<χEW<χP_W<=χW_PP then χP_PP=0.3, χE_PP=1.2 and χW_PP=5;

Cellulose surface: χW_cellulose<χE_cellulose< χEW< χP_cellulose <χPW, then χE_cellulose=0.5, χP_cellulose=3,

and χW_cellulose=0.2.

For all MesoDyn simulations in this chapter, a dimensionless time step of 50 ns, a

noise scaling parameter of 100, and a compressibility parameter of 10 are chosen. The grid

132

Page 152: Fiber Lubrication: A Molecular Dynamics Simulation Study

parameter and the bead diffusion coefficient are 1.1543 and 1.0E-7 cm2/s, respectively. The

dimension of the simulation lattice is 32×32×32 nm. In the simulation lattice, the solid wall

thickness is set to 32 nm when a pair of surfaces is present to confine the simulation lattice.

20000 time steps of dynamics and then 5000 time steps of shear were run for the MesoDyn

models. Three volume concentrations were investigated: 10%, 25% and 50%. Only the 10%

volume concentration is studied in the confined model with PP and cellulose surfaces. All

models are simulated at 298 K.

3. Results and discussion

3.1 Phase morphology of the aqueous E19P29E29 solutions under the bulk condition

Figure 6.1 shows the morphology of E19P29E19 in the bulk solution with different

volume concentrations. Figure 6.1 (a1), (b1) and (c1) are the MesoDyn models, where water

is defined as red, PEO as blue and PPO as green. In order to show the polymer structures

clearly, the water phase (red) is removed, shown in Figure 6.1 (a2), (b2) and (c2). Figure 6.1

(a3), (b3) and (c3) only shows the morphology of the PPO phase. The concentrations of the

E19P29E19 solutions are 10%, 25% and 50% varied from (a), (b) to (c).

133

Page 153: Fiber Lubrication: A Molecular Dynamics Simulation Study

(a1) (a2) (a3)

(b1) (b2) (b3)

(c1) (c2) (c3)

Figure 6.1 Phase morphology of aqueous E19P29E19 solutions under the bulk condition. Three concentrations: (a series) 10% v/v. (b series) 25% v/v. (c series) 50% v/v. Phase colors: Water—red, PEO—blue, PPO—green. All are the final frames after 20000 time steps. (a1), (b1) and (c1) are the MesoDyn models defining water as red, PEO as blue and PPO as green. In order to show the polymer structures clearly, the water phase (red) is removed, shown in (a2), (b2) and (c2). (a3), (b3) and (c3) only the morphology of the PPO phase is shown.

134

Page 154: Fiber Lubrication: A Molecular Dynamics Simulation Study

The morphology at a concentration of 10% is clearly found as spherical micelles,

shown in Figure 6.1 (a1-a3). The PPO blocks aggregate in the core of the spherical micelles,

while the PEO blocks stay outside and form a shell. The spherical micelles are surrounded by

the water phase. PPO is strongly hydrophobic; therefore the PPO blocks reside inside the

PEO shells to avoid contact with water. In a 32×32×32 nm cubic lattice, there are

approximately 13 micelles. The concentration of E19P29E19 in the aqueous solution is

approximately 10% and the PPO blocks comprise 50% of the weight segment of one

E19P29E19 molecule. The average core radius of PPO blocks obtained is 31.1 Å. The

aggregation number (N) can then be estimated by

3

34

cPO RVnN ⋅=⋅⋅ π (6.7)

where n denotes the repeat unit number, which is 29 here, and VPO is the volume of PO

monomer, about 95.4 Å3.11 Consequently, the aggregation number of N is approximately 46.

Figure 6.1 (b1), (b2) and (b3) show the morphology of a 25% E19P29E19 solution.

Instead of spherical micelles, worm-like micelles are formed. However, the PPO blocks are

still in the core, constructing a continuous worm-like shape. The shape of the PPO and PEO

phases appears spindle-like. This result is similar with the morphology of a 30% P84

polymer solution derived from Yang’s simulation work.11 As the concentration increases, the

morphology changes from micelles to bi-component structures.

When the concentration is increased to 50%, there are more E19P29E19 chains

present, which contribute to the irregular cylinder shape shown in Figure 6.1 (c1), (c2) and

(c3). The high concentration increases the contact between molecules and more cylinders are

assembled. Since the amount of water in the system is decreased, the water phase which is

135

Page 155: Fiber Lubrication: A Molecular Dynamics Simulation Study

shown is quite small.

In the dynamics processes, order parameters can be supervised, indicating a change in

molecular structures. The order parameter, P, is defined as:11

[ ]V

drrP

V∫ ∑ −≡ 1

201

21 )()( θθ

(6.8)

where θ refers to the polymer volume fraction and V is the volume of cells. P denotes the

mean-squared deviation from homogeneity in the system, which captures both the effects of

phase separation and compressibility.

The morphology revolution with the time steps is shown in Figure 6.2. At the time

step 1000, the system is disordered. PPO aggregates into small blocks at the time step 1600

and larger blocks at the time step 3000. There are some micelles formed at the time step 4500.

Most of the PPO segments aggregate into the spherical cores at the time step 5000. After the

time step 5700, all PPO segments form stable spherical cores.

Figure 6.3 shows the order parameter of the 10% E19P29E19 throughout the time

steps, which exhibits three evolving morphological stages (I, II and III). The first stage (I) is

the formation of pre-micelles, in which the order parameter and aggregate morphology

change very little. In the stage II, the order parameter increases from 1600 time step to 5700

time step, which indicates the rapid formation of large micelles. The stage (II) is critical in

determining the primary morphology of molecular structures in the solution. The stage III

allows the micelles time to adjust by overcoming the defects formed in the stage II, in order

to reach an equilibrium state.

At higher concentrations of 25% and 50%, changes in the order parameter are

136

Page 156: Fiber Lubrication: A Molecular Dynamics Simulation Study

observed and shown in Figure 6.4 and 6.5. The stage (II) and (III) are still observed, in which

the worm-like and irregular cylinder micelles are developed until in an equilibrium state in

reached. However, a stage (I) is not seen on the curves, because the rapid formation of

micelles due to increased number of polymer chains and the increased contacts among the

chains. The simulation results are not able to show the stage (I). The stages (II) are finished at

the 750 (for 25%) and 700 (for 50%) time steps. respectively.

100 1600 3000 4500

5000 5700 10000 15000

Figure 6.2 Time evolution of morphology for the aqueous 10% v/v

E19P29E19 solution. The time steps from the upper left to the lower right are: 100, 1600, 3000, 4500, 5000, 5700, 10000, 15000. Only PPO phase is displayed for clarity.

137

Page 157: Fiber Lubrication: A Molecular Dynamics Simulation Study

0 5000 10000 15000 200000.00

0.01

0.02

0.03III

Ord

er p

aram

eter

Time step

PEO PPO WaterI II

Figure 6.3 The order parameter with simulation time steps for 10%

E19P29E19 solution. The dash lines are at x=1600 and x=5700.

0 1000 2000 3000 4000 15000 200000.00

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08IIIII

Ord

er p

aram

eter

Time step

PEO PPO Water

Figure 6.4 The order parameter with simulation time steps for 25%

E19P29E19 solution. The dash line is at x=750.

138

Page 158: Fiber Lubrication: A Molecular Dynamics Simulation Study

0 1000 2000 3000 4000 15000 200000.00

0.01

0.02

0.03

0.04

0.05

0.06IIIII

Ord

er p

aram

eter

Time step

PEO PPO Water

Figure 6.5 The order parameter with simulation time steps for 50%

E19P29E19 solution. The dash line is at x=700.

3.2 Phase morphology of the aqueous E19P29E19 solutions in a confined condition

Figure 6.6 shows the morphology evolution of the aqueous 10%v/v E10P29E19

solution confined by cellulose surfaces as a function of time step. At the time step 100, the

PPO phase is random, in the middle of the confined lattice, and the morphology is different

from that of the bulk condition, as shown in Figure 6.2. In the thin slices (3~5 nm) of the

solutions that interface with the cellulose surfaces, there is not a continuous water phase. The

PPO segments aggregate to form large blocks at the time step 1600, resulting in the

appearance of micelles in the confined solution. The formation of micelles is completed by

the time step 3900. Then the micelles are quite stable in the solution meanwhile, they are

located away from the surfaces. The micelles tend to form two layers. For example, there are

six micelles in each layer at the time step 3900. The two-layer structure is also found at the

139

Page 159: Fiber Lubrication: A Molecular Dynamics Simulation Study

time step 20000. The difference is that there are five micelles in the first layer but seven

micelles in the second layer. Interestingly, at the time step 20000, two micelles merge into

one. The random-walk space of the hydrophobic PPO segments in solutions becomes

reduced because it is confined by the hydrophilic cellulose surfaces.

Figure 6.7 displays the order parameters as a function of time step for the 10% v/v

E19P29E19 confined by cellulose surfaces. The trend is similar to what is seen in the bulk

solution, there are three stages: (I) 0—1600 time steps; (II) 1600—3900 time steps; (III)

3900—20000 time steps. The PPO phase is confined randomly in the stage (I). The stage (II)

is the primary formation of micelles. However, the stage (II) is shorter than what is observed

for the bulk solution shown in Figure 6.3. Therefore, the confinement within cellulose

surfaces may accelerate the formation of micelles.

100 1600 2000 2500

3900 5000 10000 20000

Figure 6.6 Time evolution of morphology for the aqueous 10% v/v

E19P29E19 solution confined by cellulose surfaces. The time steps from the upper left to the lower right are: 100, 1600, 2000, 2500, 3900, 5000, 10000, 20000. Only PPO phase is displayed for clarity.

140

Page 160: Fiber Lubrication: A Molecular Dynamics Simulation Study

0 5000 10000 15000 200000.00

0.01

0.02

0.03 IIIII

Ord

er p

aram

eter

Time step

PEO PPO WaterI

Figure 6.7 Order parameter with simulation time steps for 10% E19P29E19

solution confined by cellulose surfaces. The dash lines are at x=1600 and x=3900.

100 300 500 1000

1500 2000 3000 20000

Figure 6.8 Time evolution of morphology for the aqueous 10% v/v

E19P29E19 solution confined by PP surfaces. The time steps from the upper left to the lower right are: 100, 300, 500, 1000, 1500, 2000, 3000, 20000. Water phase is removed for clarity. PPO—green and PEO—blue.

141

Page 161: Fiber Lubrication: A Molecular Dynamics Simulation Study

0 5000 10000 15000 200000.00

0.01

0.02

0.03

0.04

0.05III

Ord

er p

aram

eter

Time step

PEO PPO WaterII

Figure 6.9 Order parameter with simulation time steps for 10% E19P29E19

solution confined by PP surfaces.

The morphology of E19P29E19 solution confined by the PP surfaces is shown in

Figure 6.8. PP surfaces are well known as being hydrophobic.25 At the time step 100, the

initial condition for confined models is applied and there is not a continuous phase of

E19P29E19 on the surfaces. Then, at the time step 300 the continuous phase of the PPO

segments is observed on the PP surfaces, meanwhile the phase separation of PPO and PEO

occurs in the solution. From the time step 500 to 3000, the thickness of the E19P29E19 layer

increases until there is no longer a continuous phase of E19P29E19 in the center of the

confined space. From the time step 3000, the polymers are divided into two layers in the

lattice. The PPO segments adhere to PP surfaces and form a continuous layer. The PEO

segments form another layer covering the PPO layer so that the PEO layer is directly in

contact with water phase (The empty space is occupied by water).

142

Page 162: Fiber Lubrication: A Molecular Dynamics Simulation Study

The order parameters of the E19E29E19 solution confined by PP surfaces are

displayed in Figure 6.9. There is no stage (I) indicating that the polymer aggregation occurs

rapidly at the beginning of the simulation. The stage (II) and stage (III) appear to overlap

each other and there is not an obvious separation between the two stages.

3.3 Phase morphology of the aqueous E19P29E19 solutions confined and sheared

As discussed in the previous section, micelles of E19P29E19 are formed in a confined

lattice without any external force on the system. In order to obtain a better understanding of

fiber lubrication, an external force is employed in this section to create a shearing movement.

The top and bottom surfaces (PP or cellulose) are forced to move in opposite directions

relative to each other at a constant shear rate of 0.001 ns-1. In the simulation method, the final

frames generated in the last section are directly adopted here to be the initial frame for the

shear models.

3.3.1 Shearing on cellulose surfaces

The morphology of the aqueous 10% v/v E19P29E19 solution sheared by cellulose

surfaces as a function of the time step is shown in Figure 6.10. The empty space corresponds

to water in the lattice. In the shear model, at the time step 100 the micelles that formed in the

confined solutions start to distort due to the applied shear force. There is a rapid decrease in

the order parameter at the beginning of the simulation shown in Figure 6.11. Then at the time

step 1000, irregular worm-like micelles are formed and then begin breaking rapidly. Both

PPO and PEO segments tend to aggregate to form large blocks as the time step increases

143

Page 163: Fiber Lubrication: A Molecular Dynamics Simulation Study

from 1000 to 2000. In Figure 6.11, there is a dramatic drop in the order parameters as the

time step increases from 2300 to 2500. A new structure of E19P29E19 polymer in the

aqueous solution is formed in the lattice, which we refer to as a ‘wall-like micelle’. In Figure

6.11, the order parameter is found to increase after the time step 2500, and there are still

some micelle-like structures surrounding the wall-like micelle. Additionally, the PPO cores

become exposed to water under shearing. At the time step 3500, the E19P29E19 molecules

merge into wall-like micelles. During the time steps from 3500 to 5000, the wall-like

micelles grow thinner and wider from the top to the bottom of the lattice. Interestingly, there

is not a contact between the cellulose surfaces and the wall-like micelle. Therefore, it is

obvious that there is no permanent absorption of the E19P29E19 on the cellulose surface,

which is also reported by Song’s QCM measurements25 and Liang’s nanoindentation

experiments.26

In order to get better understanding of the properties of the wall-like micelles,

compressibility energy contribution (CEC) can be obtained using MesoDyn. In the shear

models, the CEC indexes the viscosity of the solution and is defined as: 27-29

V

TRZPCEC ⋅== (6.9)

where P refers to the system pressure; Z is the compressibility factor, usually a constant; R, T

and V respectively denote gas constant, temperature and volume. The CECs over time step

are shown in Figure 6.12. It is found that the CEC increase with time steps, resulting in a

higher viscosity for the wall-like micelles, and indicating that the E19P29E19 molecules are

packed with a high viscosity.

144

Page 164: Fiber Lubrication: A Molecular Dynamics Simulation Study

100 1000 2000 2500

3000 3500 4000 5000

Figure 6.10 The morphology for the 10% v/v E19P29E19 solution sheared by cellulose surfaces. The time steps from the upper left to the lower right are: 100, 1000, 2000, 2500, 3000, 3500, 4000, and 5000. Water phase is removed for clarity. PPO—green and PEO—blue. The yellow arrows denote the shearing directions of surfaces.

0 1000 2000 3000 4000 50000.00

0.01

0.02

0.03

0.04

0.05

Ord

er p

aram

eter

Time step

PEO PPO Water

Micelles broken

The wall-like micelle formed

Figure 6.11 Order parameter with simulation time steps for the 10%

E19P29E19 solution sheared by cellulose surfaces.

145

Page 165: Fiber Lubrication: A Molecular Dynamics Simulation Study

0 1000 2000 3000 4000 50000.000

0.001

0.002

0.003

0.004

0.005

0.006

Com

pres

sibi

lity

Ene

rgy

Con

tribu

tion

(RT/

Vol

ume)

Time step

Compressibility Energy Contribution

Figure 6.12 Compressibility energy contribution with the time steps for the

10% v/v E19P29E19 solution sheared by cellulose surfaces.

Therefore, the aqueous E19P29E19 solution might not be suitable as a boundary

lubricant used on the cellulose surfaces. There is little of adsorption onto the surfaces and

bare surfaces could easily wear under boundary lubrication conditions. The high viscosity

leads to more heat generated by frictional forces which could damage the surface.

3.3.2 Shearing on PP surfaces

Figure 6.13 shows the morphology of aqueous E19P29E19 solution sheared by the PP

surfaces. In the time range of MD simulations, the E19P29E19 molecules attach on two PP

surfaces (upper and lower surfaces) to form layers which fully cover the surfaces. The

thickness of the layers is measured at 3-4 nm using the image. There is not a significant

change in morphology among the time steps 100, 1000 and 5000, indicating that the binding

146

Page 166: Fiber Lubrication: A Molecular Dynamics Simulation Study

interaction is strong enough to withstand the 0.001 ns-1 shear rate. The strong interaction

between PPO and PP surfaces is in agreement with the conclusions stated in Chapter 5 and

the experimental results generated from Song25 and Liang26.

Figure 6.14 displays the order parameter as a function of the time step for the PPO,

PEO and water. Although there is a linearly increasing trend over time for the three

components under the shearing, the changes are actually quite small. This indicates that the

molecules slowly organize themselves into layers and do not undergo significant change in

molecular structure under the application of shear.

Therefore, the E19P29E19 is an appropriate lubricant for PP surfaces due to the large

amount of adsorption, as well as the continuous thin layer built up to protect the PP surfaces.

This layer is vital in boundary lubrication.

100 1000 5000

Figure 6.13 The morphology for the 10% v/v E19P29E19 solution sheared

by PP surfaces. The time steps from the upper left to the lower right are: 100, 1000, and 5000. Water phase is removed for clarity. PPO—green and PEO—blue. The yellow arrows denote the shearing directions of PP surfaces.

147

Page 167: Fiber Lubrication: A Molecular Dynamics Simulation Study

0 1000 2000 3000 4000 50000.00

0.01

0.02

0.03

0.04

0.05

Ord

er p

aram

eter

Time step

PEO PPO Water

Figure 6.14 Order parameter with simulation time steps for the 10%

E19P29E19 solution sheared by PP surfaces.

4. Conclusion

In this chapter, the morphology of a PAG triblock copolymer, E19P29E19, was

investigated in bulk solution, under the confined condition and the confined and sheared

condition. The E19P29E19 chains are found to form unique self-assembled structures that are

dependent of the solution concentration.

In the bulk aqueous solution with a concentration of 10% v/v, the micelle structure is

formed which exhibits in a fashion of PPO blocks inside and PEO blocks outside. When the

concentration increases from 25% to 50%, the worm-like micelles and irregular cylinders are

observed in the solutions, respectively. The micelle core radius of the E19P29E19 in the 10%

solution is calculated as 31.1 Å, and the aggregation number, N, is approximately 46. The

micelles are formed from the time step 1600 to 5700.

148

Page 168: Fiber Lubrication: A Molecular Dynamics Simulation Study

When the aqueous 10% v/v E19P29E19 solution is confined by cellulose surfaces, the

resulting micelles are located apart from the cellulose surfaces. Cellulose is hydrophilic and

PPO is hydrophobic. The formation of micelles is faster than what is observed in the bulk

solution, and merging between micelles is also found. When the solution is confined by

cellulose surfaces, the random-walk space of the PPO segments increases and there is a high

probability of the segments to contact each other, resulting in the rapid formation of micelles.

When confined by PP surfaces, the E19P29E19 copolymers form thin layers adsorbed onto

the top and bottom of the PP surfaces.

The E19P29E19 solution is not only studied in a confined condition without any

external force, but is also subjected to the application of a shear force. The model frames

confined by PP and cellulose are used to apply a shear rate of 0.001 ns-1. In the shear model

of cellulose surfaces, the micelles are distorted due to the shearing. The distorted micelles

continue to aggregate into the new structure, wall-like micelles, which are not in direct

contact with the cellulose surface. The order parameter also confirms the dynamics of the

micelle formation. The compressibility energy contribution (CEC) indicates that the wall-like

micelle structures lead to a higher viscosity. E19P29E19 is not considered a proper lubricant

for cellulose surfaces in boundary lubrication. On the other hand, when being sheared by PP

surfaces, the aqueous E19P20E19 solution does not have a significant change in phase

behavior. The PPO segments are assembled into two thin films adsorbed to PP surfaces.

These thin films can prevent the PP surfaces from friction and wear. Therefore, the

E19P29E19 is a good lubricant for PP surfaces in boundary lubrication.

149

Page 169: Fiber Lubrication: A Molecular Dynamics Simulation Study

References:

1. Fraaije, J. G. E. M., Dynamic density functional theory for microphase separation kinetics of block copolymer melts. J. Chem. Phys. 1993, 99, (11), 9202-12.

2. Fraaije, J. G. E. M.; Van Vlimmeren, B. A. C.; Maurits, N. M.; Postma, M.; Evers, O. A.; Hoffmann, C.; Altevogt, P.; Goldbeck-Wood, G., The dynamic mean-field density functional method and its application to the mesoscopic dynamics of quenched block copolymer melts. J. Chem. Phys. 1997, 106, (10), 4260-4269.

3. Groot, R. D., Electrostatic interactions in dissipative particle dynamics-simulation of polyelectrolytes and anionic surfactants. Journal of Chemical Physics 2003, 118, (24), 11265-11277.

4. Nobel, A. Stability of aluminate solutions. 439028, 2000.

5. Fraaije, J. G. E. M.; Zvelindovsky, A. V.; Sevink, G. J. A., Computational soft nanotechnology with mesodyn. Molecular Simulation 2004, 30, (4), 225-238.

6. van Vlimmeren, B. A. C.; Maurits, N. M.; Zvelindovsky, A. V.; Sevink, G. J. A.; Fraaije, J. G. E. M., Simulation of 3D mesoscale structure formation in concentrated aqueous solution of the triblock polymer surfactants (ethylene oxide)13(propylene oxide)30(ethylene oxide)13 and (propylene oxide)19(ethylene oxide)33(propylene oxide)19. Application of dynamic mean-field density functional theory. Macromolecules 1999, 32, (3), 646-656.

7. Lam, Y.-M.; Goldbeck-Wood, G., Mesoscale simulation of block copolymers in aqueous solution: Parameterisation, micelle growth kinetics and the effect of temperature and concentration morphology. Polymer 2003, 44, (12), 3593-3605.

8. Guo, S. L.; Hou, T. J.; Xu, X. J., Simulation of the phase behavior of the (EO)13(PO)30(EO)13(Pluronic L64)/water/p-xylene system using MesoDyn. Journal of Physical Chemistry B 2002, 106, (43), 11397-11403.

9. Li, Y.; Hou, T.; Guo, S.; Wang, K.; Xu, X., The Mesodyn simulation of pluronic water mixtures using the 'equivalent chain' method. Phys. Chem. Chem. Phys. 2000, 2, (12), 2749-2753.

10. Zhao, Y.; Chen, X.; Yang, C.; Zhang, G., Mesoscopic simulation on phase behavior of pluronic P123 aqueous solution. Journal of Physical Chemistry B 2007, 111, (50), 13937-13942.

11. Yang, S.; Yuan, S.; Zhang, X.; Yan, Y., Phase behavior of tri-block copolymers in solution: Mesoscopic simulation study. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2008, 322, (1-3), 87-96.

12. Spyriouni, T.; Vergelati, C., A molecular modeling study of binary blend compatibility of polyamide 6 and poly(vinyl acetate) with different degrees of hydrolysis: An atomistic and mesoscopic approach. Macromolecules 2001, 34, (15), 5306-5316.

13. Jawalkar, S. S.; Adoor, S. G.; Sairam, M.; Nadagouda, M. N.; Aminabhavi, T. M., Molecular modeling on the binary blend compatibility of poly (vinyl alcohol) and poly (methyl methacrylate): An atomistic simulation and thermodynamic approach. Journal of Physical Chemistry B 2005, 109, (32), 15611-15620.

14. Jawalkar, S. S.; Nataraj, S. K.; Raghu, A. V.; Aminabhavi, T. M., Molecular dynamics simulations on the

150

Page 170: Fiber Lubrication: A Molecular Dynamics Simulation Study

blends of poly(vinyl pyrrolidone) and poly(bisphenol-A-ether sulrone). Journal of Applied Polymer Science 2008, 108, (6), 3572-6.

15. Altevogt, P.; Evers, O. A.; Fraaije, J. G. E. M.; Maurits, N. M.; van Vlimmeren, B. A. C. In The MesoDyn project: software for mesoscale chemical engineering, Lisbon, Portugal, 1999; Elsevier: Lisbon, Portugal, 1999; pp 139-43.

16. Yuan, S.; Xu, G.; Luan, Y.; Liu, C., The interaction between polymer and AOT or NaDEHP in aqueous solution: Mesoscopic simulation study and surface tension measurement. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2005, 256, (1 SPEC ISS), 43-50.

17. Zhang, X.; Liu, Q., PAn-PEG-PAn self-assembly in aqueous solution by DPD simulation. Huagong Xuebao/Journal of Chemical Industry and Engineering (China) 2007, 58, (11), 2948-2952.

18. Yang, C.; Chen, X.; Qiu, H.; Zhuang, W.; Chai, Y.; Hao, J., Dissipative particle dynamics simulation of phase behavior of aerosol OT/water system. Journal of Physical Chemistry B 2006, 110, (43), 21735-21740.

19. Pedersen, J. S.; Sommer, C., Temperature dependence of the virial coefficients and the chi parameter in semi-dilute solutions of PEG. Prog. Colloid Polym. Sci. 2005, 130, 70-78.

20. Knoll, A.; Magerle, R.; Krausch, G., Phase behavior in thin films of cylinder-forming ABA block copolymers: Experiments. Journal of Chemical Physics 2004, 120, (2), 1105-1116.

21. Knoll, A.; Horvat, A.; Lyakhova, K. S.; Krausch, G.; Sevink, G. J. A.; Zvelindovsky, A. V.; Magerle, R., Phase Behavior in Thin Films of Cylinder-Forming Block Copolymers. Physical Review Letters 2002, 89, (3), 035501.

22. Horvat, A.; Lyakhova, K. S.; Sevink, G. J. A.; Zvelindovsky, A. V.; Magerle, R., Phase behavior in thin films of cylinder-forming ABA block copolymers: Mesoscale modeling. Journal of Chemical Physics 2004, 120, (2), 1117-1126.

23. Lyakhova, K. S.; Horvat, A.; Zvelindovsky, A. V.; Sevink, G. J. A., Dynamics of terrace formation in a nanostructured thin block copolymer film. Langmuir 2006, 22, (13), 5848-5855.

24. Jain, N. J.; George, A.; Bahadur, P., Effect of salt on the micellization of pluronic P65 in aqueous solution. Colloids and Surfaces A: Physicochemical and Engineering Aspects 1999, 157, (1-3), 275-283.

25. Song, J., Adsorption of amphoteric and nonionic polymers on model thin films. 2008; p 232.

26. Liang, J., Investigation of synthetic and natural lubricants. 2008; p 149.

27. Gill, D. S.; Vermani, B. K.; Sharma, R. P., Molar conductance, viscosity and isentropic compressibility of some cobalt (III) complexes in organic solvents. Journal of Molecular Liquids 2006, 124, (1-3), 58-62.

28. Gromakovskij, D. G.; Marinin, V. B., On volume viscosity and lubricant compressibility during forming contact hydrodynamic effects. Trenie i Iznos 1992, 33, (5), 801-810.

29. Wahab, A.; Mahiuddin, S., Isentropic compressibility and viscosity of aqueous and methanolic calcium chloride solutions. Journal of Chemical and Engineering Data 2001, 46, (6), 1457-1463.

151

Page 171: Fiber Lubrication: A Molecular Dynamics Simulation Study

Chapter 7

Future Work

Future work can include but not limited to:

(1) Confined amorphous polymer layers

The amorphous polymers are of great interest because they exist in many applications.

We have successfully built a confined amorphous layer of cellulose with use of a xenon

crystal. Except for the pure cellulose discussed previously, there is a large number of

cellulose derivatives products consumed in industrial and civilian application, such as methyl

cellulose (MC), hydroxypropyl cellulose (HPC), and carboxymethyl cellulose (CMC), etc.

The xenon method could be applied to build confined amorphous layers of these cellulose

derivatives. Furthermore, metal nanoparticles have been introduced in nanofibers made from

cellulose in the applications of filtration, catalyst, and antibacterial materials, Metal

nanoparticles can be included in the amorphous polymer layers so that we can investigate the

effects of nanoparticles on the polymer matrix.

(2) Contact angle simulations

The contact angle models we proposed are on a basis of the comprehensive force field.

It is proven that a high accuracy can be obtained to predict the contact angles on a surface.

The shape of water nano-droplet can be obtained from the weight distribution, which

provides an easy visual tool to investigate the contact angle. Therefore, we can create nano-

droplets made from polymer melts on various surfaces. The simulation results can provide a

video to discuss the affinity of polymer melts on surfaces.

152

Page 172: Fiber Lubrication: A Molecular Dynamics Simulation Study

(3) Temperature influence on the single molecule PWS system

Temperature is the vital state variable to specify the thermodynamic state of the PWS

system. To study the trend of intermolecular interaction energy in the PWS system could

further validate the PWS interaction calculation method and guide the proper temperature in

experiments.

(4) PWS system with multiple polymer molecules

Multiple polymer molecules can be set in the aqueous solution confined with surface

to study the configuration of multiple polymer molecules. The self-assembly properties of

multiple polymer molecules in solution would be helpful to understand the ultra-thin

polymeric film impacted by surfaces.

(5) Shear models of MD simulation

The shear models in MD simulations are not included in this dissertation. Only

mesoscale shear models are applied in Chapter 6. MesoDyn shear models could provide the

morphology information of the lubricant solutions and surfaces. But the atomic information

can not be obtained. In MD simulations, the PWS interaction can be a dynamic value and the

polymer configurations can be acquired at the atomic level. Also the velocity profile, the

temperature profile and the shear stress could be provided to further understand the

lubrication at nanoscale.

(6) Flory-Huggins interaction parameters

MesoDyn dynamics is based on the Flory-Huggins interaction parameters. The

interaction parameter with non-aqueous system is obtained. However in the aqueous system,

we still don’t have a full understanding on the interaction parameters due to the influence of

153

Page 173: Fiber Lubrication: A Molecular Dynamics Simulation Study

hydrogen bonding. Even the developers of MesoDyn method roughly estimate the interaction

parameter between PPO and PEO. The interaction energy is discussed in Chapters 4 and 5,

which is based on the accurate forcefiled. It is interested to build the relationship between the

interaction energy with the Flory-Huggins parameters.

(7) Other parameters discussed in MesoDyn models

In the MesoDyn models, a few other parameters could be further discussed, such as

the distance of the two confined surfaces and the solution concentrations. Their relationship

with the morphology of polymer solutions could be investigated. In addition, the confined

surfaces could be modified with charged groups. The results can help us to design the

polymer morphology as the applications require.

154

Page 174: Fiber Lubrication: A Molecular Dynamics Simulation Study

APPENDICES

Appendix 1. The perl script to calculate the interaction energy of two layers #!perl # Original Author: Stephen Todd from Acclerys Software Inc., San Diego, CA USA # Revised by Hongyi Liu, NCSU, Raleigh, NC USA #1. the forcefiel option is added. So it can use COMPASS now. #2. Previously this program can not calculate PPO and PEO. The charge Tolerance added. #3. The unit of interaction energy is clarified. And we change it as cal/m^2. ########################################################################### # This script calculates the interaction energy between two layers in a structure. # The layers have to be defined as sets called Layer 1 and Layer 2 - these are the # default names assigned by the Layer builder. # This also uses the default settings for Forcite - ie Universal with current charges. # You should change the settings if you want to use another forcefield or summation method. # This script also assumes that the surface area is in the AB-plane. # A study table is produced containing the layers, energies, interaction energy, and then # the interaction energy per angstrom^2. # Note. The input trajectory should have a large vaccuum, greater than the # cut-off you are using in the non-bond calculation. use strict; use MaterialsScript qw(:all); ########################################################################### # Begin user editable settings # my $filename = "Layer"; # Name of the trajectory without extension # # End user editable settings ###########################################################################

155

Page 175: Fiber Lubrication: A Molecular Dynamics Simulation Study

# Defines the trajectory document my $doc = $Documents{"$filename.xtd"}; # Create a new study table to hold the structures and energies and set the headings my $newStudyTable = Documents->New("$filename"."_IntEnergy.std"); my $calcSheet = $newStudyTable->Sheets->Item(0); $calcSheet->ColumnHeading(0) = "$filename Cell"; $calcSheet->ColumnHeading(1) = "Energy of $filename Cell"; $calcSheet->ColumnHeading(2) = "Layer 1"; $calcSheet->ColumnHeading(3) = "Energy of Layer 1"; $calcSheet->ColumnHeading(4) = "Layer 2"; $calcSheet->ColumnHeading(5) = "Energy of layer 2"; $calcSheet->ColumnHeading(6) = "Interaction Energy"; $calcSheet->ColumnHeading(7) = "Interaction Energy (cal/m^2)"; # Get the total number of frames in the trajectory my $numFrames = $doc->Trajectory->NumFrames; print "Calculating interaction energy for $numFrames frames on $filename trajectory\n"; # Calculates the area of the surface based on the surface being in the A-B plane my $length1 = $doc->Lattice3D->LengthA; my $length2 = $doc->Lattice3D->LengthB; my $surfaceArea = $length1 * $length2; print "The surface area is $surfaceArea angstroms^2\n"; # Initialise Forcite. If you want to change the settings, you can load them or # use a change settings command here. my $forcite = Modules->Forcite; $forcite->ChangeSettings([ CalculationQuality => "Medium", '3DPeriodicElectrostaticSummationMethod' => "Group based", '3DPeriodicvdWSummationMethod' => "Group based", CurrentForcefield => "COMPASS", TotalChargeTolerance=> 0.17, #for EO segment ]);

156

Page 176: Fiber Lubrication: A Molecular Dynamics Simulation Study

# Starts to loop over the frames in the trajectory for ( my $count=1; $count<=$numFrames; $count++) { print "Calculating energy for frame $count\n"; # Define the frame of the trajectory $doc->Trajectorys->Item(0)->CurrentFrame = $count; # Create three documents to hold the temporary layers my $allDoc = Documents->New("all.xsd"); my $layer1Doc = Documents->New("Layer1.xsd"); my $layer2Doc = Documents->New("Layer2.xsd"); # Create a copy of the structure in temporary documents and delete # the layers that you don't need. $allDoc->CopyFrom($doc); $layer1Doc->CopyFrom($doc); $layer1Doc->DisplayRange->Sets("Layer 2")->Atoms->Delete; $layer2Doc->CopyFrom($doc); $layer2Doc->DisplayRange->Sets("Layer 1")->Atoms->Delete; # Put the structures in a study table for safekeeping $calcSheet->Cell($count-1, 0) = $allDoc; $calcSheet->Cell($count-1, 2) = $layer1Doc; $calcSheet->Cell($count-1, 4) = $layer2Doc; #Calculate the energies for all the layers and put in the study table $forcite->Calculation->Run($allDoc, [Task => 'Energy']); $calcSheet->Cell($count-1, 1) = $allDoc->PotentialEnergy; $forcite->Calculation->Run($layer1Doc, [Task => 'Energy']); $calcSheet->Cell($count-1, 3) = $layer1Doc->PotentialEnergy; $forcite->Calculation->Run($layer2Doc, [Task => 'Energy']); $calcSheet->Cell($count-1, 5) = $layer2Doc->PotentialEnergy; # Calculate the Interaction Energy from the cells in the Study Table

157

Page 177: Fiber Lubrication: A Molecular Dynamics Simulation Study

my $totalEnergy = $calcSheet->Cell($count-1,1); my $layer1Energy = $calcSheet->Cell($count-1,3); my $layer2Energy = $calcSheet->Cell($count-1,5); my $interactionEnergy = $totalEnergy - ($layer1Energy + $layer2Energy); $calcSheet->Cell($count-1, 6) = $interactionEnergy; my $interactionEnergyArea = $interactionEnergy / (6.022 * $surfaceArea); $calcSheet->Cell($count-1, 7) = $interactionEnergyArea; # Discard the temporary documents $allDoc->Discard; $layer1Doc->Discard; $layer2Doc->Discard; } print "Calculation complete.\n";

158

Page 178: Fiber Lubrication: A Molecular Dynamics Simulation Study

Appendix 2. The perl script to calculate the interaction energies of PWS system #!perl # This script calculates the interaction energy in PWS system. # three sets in the system: polymer, water, and surface. # Reference: http://forums.accelrys.org/eve/forums/a/tpc/f/3091072081/m/3031042881 use strict; use MaterialsScript qw(:all); ########################################################################### # my $filename = "filename"; # Name of the trajectory without extension # To initialize Forcite. # To use a change settings command here. my $forcite = Modules->Forcite; $forcite->ChangeSettings([ CalculationQuality => "Medium", '3DPeriodicElectrostaticSummationMethod' => "Group based", '3DPeriodicvdWSummationMethod' => "Group based", CurrentForcefield => "COMPASS", TotalChargeTolerance => 0.17, ]); # ########################################################################### # To define the current trajectory document my $doc = $Documents{"$filename.xtd"}; # To create a new study table to hold the structures and energies and set the headings my $newStudyTable = Documents->New("$filename"."_IntEnergy.std"); my $calcSheet = $newStudyTable->Sheets->Item(0); $calcSheet->ColumnHeading(0) = "$filename Cell"; $calcSheet->ColumnHeading(1) = "Energy of $filename Cell";

159

Page 179: Fiber Lubrication: A Molecular Dynamics Simulation Study

$calcSheet->ColumnHeading(2) = "Surface or Polymer"; $calcSheet->ColumnHeading(3) = "Energy of SP"; $calcSheet->ColumnHeading(4) = "Polymer or Water"; $calcSheet->ColumnHeading(5) = "Energy of PW"; $calcSheet->ColumnHeading(6) = "Water or Surface"; $calcSheet->ColumnHeading(7) = "Energy of WS"; $calcSheet->ColumnHeading(8) = "Surface"; $calcSheet->ColumnHeading(9) = "Energy of Surface"; $calcSheet->ColumnHeading(10) = "Water"; $calcSheet->ColumnHeading(11) = "Energy of Water"; $calcSheet->ColumnHeading(12) = "Polymer"; $calcSheet->ColumnHeading(13) = "Energy of Polymer"; $calcSheet->ColumnHeading(14) = "Energy of SP cross"; $calcSheet->ColumnHeading(15) = "Energy of PW cross"; $calcSheet->ColumnHeading(16) = "Energy of WS cross"; $calcSheet->ColumnHeading(17) = "Energy of Triangle"; $calcSheet->ColumnHeading(18) = "Energy of SPinteraction (kcal/mol)"; $calcSheet->ColumnHeading(19) = "Energy of PWinteraction (kcal/mol)"; $calcSheet->ColumnHeading(20) = "Energy of WSinteraction (kcal/mol)"; #$calcSheet->ColumnHeading(21) = "SP interacton per angstrom^2"; #for solution. # To get the total number of frames in the trajectory my $numFrames = $doc->Trajectory->NumFrames; print "Calculating interaction energy for $numFrames frames on $filename trajectory\n"; # Starts to loop over the frames in the trajectory for ( my $count=1; $count<=$numFrames; $count++) { print "Calculating energy for frame $count\n"; # Define the frame of the trajectory $doc->Trajectorys->Item(0)->CurrentFrame = $count; # Create three documents to hold the temporary layers

160

Page 180: Fiber Lubrication: A Molecular Dynamics Simulation Study

my $total = Documents->New("total.xsd"); my $SP = Documents->New("SP.xsd"); my $PW = Documents->New("PW.xsd"); my $WS = Documents->New("WS.xsd"); my $Surface = Documents->New("Surface.xsd"); my $Water = Documents->New("Water.xsd"); my $Polymer =Documents->New("Polymer.xsd"); # Create a copy of the structure in temporary documents and delete # the layers that you don't need. $total->CopyFrom($doc); $SP->CopyFrom($doc); $SP->DisplayRange->Sets("water")->Atoms->Delete; $PW->CopyFrom($doc); $PW->DisplayRange->Sets("surface")->Atoms->Delete; $WS->CopyFrom($doc); $WS->DisplayRange->Sets("polymer")->Atoms->Delete; $Surface->CopyFrom($SP); $Surface->DisplayRange->Sets("polymer")->Atoms->Delete; $Water->CopyFrom($PW); $Water->DisplayRange->Sets("polymer")->Atoms->Delete; $Polymer->CopyFrom($PW); $Polymer->DisplayRange->Sets("water")->Atoms->Delete; # Put the structures in a study table $calcSheet->Cell($count-1, 0) = $total; $calcSheet->Cell($count-1, 2) = $SP; $calcSheet->Cell($count-1, 4) = $PW; $calcSheet->Cell($count-1, 6) = $WS; $calcSheet->Cell($count-1, 8) = $Surface; $calcSheet->Cell($count-1, 10) = $Water; $calcSheet->Cell($count-1, 12) = $Polymer; #Calculate the energies

161

Page 181: Fiber Lubrication: A Molecular Dynamics Simulation Study

$forcite->Calculation->Run($total, [Task => 'Energy']); $calcSheet->Cell($count-1, 1) = $total->PotentialEnergy; $forcite->Calculation->Run($SP, [Task => 'Energy']); $calcSheet->Cell($count-1, 3) = $SP->PotentialEnergy; $forcite->Calculation->Run($PW, [Task => 'Energy']); $calcSheet->Cell($count-1, 5) = $PW->PotentialEnergy; $forcite->Calculation->Run($WS, [Task => 'Energy']); $calcSheet->Cell($count-1, 7) = $WS->PotentialEnergy; $forcite->Calculation->Run($Surface, [Task => 'Energy']); $calcSheet->Cell($count-1, 9) = $Surface->PotentialEnergy; $forcite->Calculation->Run($Water, [Task => 'Energy']); $calcSheet->Cell($count-1, 11) = $Water->PotentialEnergy; $forcite->Calculation->Run($Polymer, [Task => 'Energy']); $calcSheet->Cell($count-1, 13) = $Polymer->PotentialEnergy; # Calculate the Interaction Energy from the cells in the Study Table my $totale = $calcSheet->Cell($count-1,1); my $spe = $calcSheet->Cell($count-1,3); my $pwe = $calcSheet->Cell($count-1,5); my $wse = $calcSheet->Cell($count-1,7); my $se = $calcSheet->Cell($count-1,9); my $we = $calcSheet->Cell($count-1,11); my $pe = $calcSheet->Cell($count-1,13); my $spintera = $totale - $pwe - $wse + $we; $calcSheet->Cell($count-1,18) = $spintera; # Discard the temporary documents $total->Discard; $SP->Discard; $PW->Discard;

162

Page 182: Fiber Lubrication: A Molecular Dynamics Simulation Study

$WS->Discard; $Surface->Discard; $Water->Discard; $Polymer->Discard; } print "Calculation complete.\n";

163