15.08.2013 Views

Abilash Nair: Dissertation - acumen - The University of Alabama

Abilash Nair: Dissertation - acumen - The University of Alabama

Abilash Nair: Dissertation - acumen - The University of Alabama

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

MULTISCALE SIMULATION OF POLYMER NANO-COMPOSITES (PNC) USING<br />

MOLECULAR DYNAMICS (MD) AND GENERALIZED INTERPOLATION<br />

MATERIAL POINT METHOD (GIMP)<br />

by<br />

ABILASH R. NAIR<br />

A DISSERTATION<br />

Submitted in partial fulfillment <strong>of</strong> the requirements<br />

for the degree <strong>of</strong> Doctor in Philosophy in the<br />

Department <strong>of</strong> Aerospace Engineering and Mechanics<br />

in the Graduate School <strong>of</strong><br />

<strong>The</strong> <strong>University</strong> <strong>of</strong> <strong>Alabama</strong><br />

TUSCALOOSA, ALABAMA<br />

2010


Copyright <strong>Abilash</strong> R. <strong>Nair</strong> 2010<br />

ALL RIGHTS RESERVED


ABSTRACT<br />

Recent mechanical characterization experiments with pultruded E-Glass / polypropy-<br />

lene (PP) and compression molded E-Glass/Nylon-6 composite samples with 3-4 weight%<br />

nanoclay and baseline polymer (polymer without nanoclay) confirmed significant im-<br />

provements in compressive strength (∼122%) and shear strength (∼60%) in the nan-<br />

oclay modified nanocomposites, in comparison with baseline properties. Uniaxial tensile<br />

tests showed a small increase in tensile strength (∼3.4%) with 3 wt % nanoclay load-<br />

ing. While the synergistic reinforcing influence <strong>of</strong> nanoparticle reinforcement is obvious,<br />

a simple rule-<strong>of</strong>-mixtures approach fails to quantify the dramatic increase in mechanical<br />

properties. Consequently, there is an immediate need to investigate and understand the<br />

mechanisms at the nanoscale that are responsible for such unprecedented strength en-<br />

hancements. In this work, an innovative and effective method to model nano-structured<br />

components in a thermoplastic polymer matrix is proposed. Effort will be directed to-<br />

wards finding fundamental answers to the reasons for significant changes in mechanical<br />

properties <strong>of</strong> nanoparticle-reinforced thermoplastic composites. This research ensues a<br />

multiscale modeling approach in which (a) a concurrent simulations scheme is developed<br />

to visualize atomistic behavior <strong>of</strong> polymer molecules as a function <strong>of</strong> continuum scale<br />

loading conditions and (b) a novel nanoscale damage mechanics model is proposed to<br />

capture the constitutive behavior <strong>of</strong> polymer nano composites (PNC). <strong>The</strong> proposed re-<br />

search will contribute towards the understanding <strong>of</strong> advanced nanostructured composite<br />

materials, which should subsequently benefit the composites manufacturing industry.<br />

ii


DEDICATION<br />

Dedicated to my teachers<br />

and<br />

my parents Ambika and Rajendran <strong>Nair</strong>.<br />

iii


LIST OF ABBREVIATIONS AND SYMBOLS<br />

ρ current density <strong>of</strong> material<br />

ρ 0 undeformed (initial) density <strong>of</strong> material<br />

J Jacobian (or determinant) <strong>of</strong> deformation gradient<br />

F deformation gradient tensor<br />

C right Cauchy-Green deformation tensor<br />

S 2 nd Piola-Kirch<strong>of</strong>f stress tensor<br />

σ Cauchy stress tensor<br />

ɛ small strain tensor<br />

E Green-Lagrange strain tensor<br />

D damage tensor<br />

T ambient temperature<br />

A Helmholtz free energy<br />

kB<br />

Boltzmann constant<br />

∆t time-step for continuum update<br />

∆τ time-step for discrete (atomistic) update<br />

1 fs 1 femto(10 −15 ) seconds<br />

1 ps 1 pico(10 −12 ) seconds<br />

1 ˚A 1 Angstrom (10 −10 ) meters<br />

1 nm 1 nano-(10 −9 ) meter<br />

RVE Representative Volume Element<br />

MD Molecular Dynamics<br />

GIMP Generalized Interpolation Material Point Method<br />

FEA Finite Element Analysis<br />

PNC Polymer Nano-composite<br />

iv


ACKNOWLEDGEMENTS<br />

I would like to use this space to acknowledge the help and encouragement without<br />

which this dissertation would not have come to realization. While the work undertaken in<br />

this research is challenging, the words <strong>of</strong> wisdom provided by my advisor Dr. Samit Roy<br />

encouraged me to learn new methodologies and equip myself with the necessary tools<br />

to pursue this research. Being a student <strong>of</strong> mechanics, I have had the opportunity to<br />

learn many elements <strong>of</strong> physics and chemistry that is an integral part <strong>of</strong> this dissertation.<br />

This work was supported by the NSF-EPSCoR funded GRSP program. Applying to<br />

the GRSP, has opened my eyes to an important part <strong>of</strong> a career in research: writing<br />

proposals. I thank the AEM department at UA, for providing me with the teaching<br />

assistantship that allowed me to complete this research. I like to thank Mr. Rezwanur<br />

Rehman for his help with the nano-clay modeling and Mr. Kamesh Narasimhan for his<br />

discussions on nano-mechanics <strong>of</strong> polymer nano-composites. I would like to thank my<br />

committee members for taking the time to read my dissertation and providing comments<br />

to improve on the manuscript. I thank Mr. Cameron Purvis at O.I.T. for his help with<br />

compiling the parallel version <strong>of</strong> LAMMPS.<br />

I thank my family whose constant support and encouragement helped me through<br />

the hardest times in my academic career. I thank them for their trust and confidence in<br />

my decisions. I would like to close by recognizing the invaluable help and support <strong>of</strong> my<br />

close friends who have been family to me for the past four years. While it may not seem<br />

long, most lifetimes are lived within a breath and its only in the briefest <strong>of</strong> moments like<br />

these that we appreciate the ones who were there when it really mattered. It seems only<br />

fitting that this dissertation marks the end <strong>of</strong> an era when I had my friends graduate and<br />

move on, and now my time has come with the proverbial “passing <strong>of</strong> the buck”.<br />

<strong>The</strong> air, pregnant with excitement beckons a future still unknown, but full <strong>of</strong> hope. To<br />

the pages that have been written and the many pages that are waiting to be done.<br />

v


CONTENTS<br />

ABSTRACT .................................................................................. ii<br />

LIST OF ABBREVIATIONS AND SYMBOLS ............................................ iv<br />

ACKNOWLEDGEMENTS .................................................................. v<br />

LIST OF TABLES ........................................................................... ix<br />

LIST OF FIGURES .......................................................................... x<br />

1. Introduction ............................................................................... 1<br />

1.1. Motivation............................................................................ 2<br />

1.2. Literature Review.................................................................... 7<br />

2. Generalized Interpolation Material Point Method (GIMP) ............................ 16<br />

2.1. Equations <strong>of</strong> Motion <strong>of</strong> a Deformable body........................................ 18<br />

2.1.1. Discretizing governing equation ............................................ 19<br />

2.2. Explicit Time Integration in GIMP ................................................ 26<br />

2.2.1. A Finite Deformation Hyper-Elastic Model for Polymeric Materials .... 29<br />

2.2.2. <strong>The</strong> Verlet Integration in GIMP............................................ 31<br />

2.3. Dynamic Fracture Analysis using GIMP ........................................... 33<br />

3. Molecular Dynamics (MD) ............................................................... 37<br />

3.1. Introduction .......................................................................... 38<br />

3.1.1. Equations <strong>of</strong> Motion......................................................... 41<br />

3.1.2. <strong>The</strong> Verlet Algorithm ....................................................... 42<br />

3.2. Molecular Force Fields .............................................................. 44<br />

3.2.1. Bond stretching .............................................................. 46<br />

3.2.2. Angle bendings .............................................................. 47<br />

3.2.3. Bond torsion ................................................................. 47<br />

3.2.4. Non-bonded Interactions .................................................... 48<br />

vi


3.3. Modeling <strong>The</strong>rmoplastic Polymer Nanocomposites in MD ....................... 50<br />

3.3.1. <strong>The</strong>rmoplastic Polymers .................................................... 51<br />

3.3.2. Nanoclay ..................................................................... 55<br />

3.4. Applications <strong>of</strong> the Molecular Dynamics Algorithm............................... 57<br />

3.4.1. Determination <strong>of</strong> Free-Energy due to deformation ........................ 58<br />

3.4.2. Multi-scale modeling <strong>of</strong> <strong>The</strong>rmoplastic composites ....................... 62<br />

4. Concurrent Multiscale Modeling <strong>of</strong> <strong>The</strong>rmoplastic PNC ............................... 65<br />

4.1. Introduction .......................................................................... 65<br />

4.2. Embedded Statistical Coupling in GIMP .......................................... 69<br />

4.3. Setup <strong>of</strong> Coupled Simulation model ................................................ 73<br />

5. Hierarchical Modeling <strong>of</strong> Damage in PNCs ............................................. 77<br />

5.1. Damage Mechanics using the Internal State Variable Approach ................. 79<br />

5.1.1. Definition <strong>of</strong> Damage Tensor ............................................... 79<br />

5.1.2. <strong>The</strong>rmodynamics <strong>of</strong> Damage and Constitutive equations ................ 81<br />

5.2. A Nano-scale Damage Model ....................................................... 84<br />

5.2.1. A Damage Model for Interphase Debond and Void Nucleation in<br />

<strong>The</strong>rmoplastic Polymer Nanocomposites .................................. 92<br />

6. Results ..................................................................................... 98<br />

6.1. Concurrent Coupling <strong>of</strong> <strong>The</strong>rmoplastic Polymers ................................. 98<br />

6.1.1. Case I: Benchmark <strong>of</strong> Coupling algorithm using a Simple Tension Test. 103<br />

6.1.2. Case II: Mode I Crack Growth in <strong>The</strong>rmoplastic Polymer ............... 105<br />

6.1.3. Case III: Mode II crack growth in <strong>The</strong>rmoplastic Polymer ............... 107<br />

6.2. Free Energy Calculations <strong>of</strong> <strong>The</strong>rmoplastic Polymers ............................. 111<br />

6.2.1. Calculation <strong>of</strong> Persistence Length <strong>of</strong> a Polymer ........................... 112<br />

6.2.2. Calculation <strong>of</strong> Free energy <strong>of</strong> Polymers undergoing Mechanical Deformation<br />

..................................................................... 115<br />

6.3. Multi-scale Damage model for <strong>The</strong>rmoplastic Nanocomposites .................. 122<br />

6.3.1. Simulation results for <strong>The</strong>rmoplastic Polymer Nano-composite .......... 124<br />

vii


6.3.2. Simulation results for pure Polypropylene without Nanoscale Reinforcement.....................................................................<br />

138<br />

6.3.3. Comparison <strong>of</strong> Stiffness and Strength between Polymer Nanocomposite<br />

and Pure Polymer .................................................... 143<br />

6.3.4. Summary ..................................................................... 145<br />

7. Discussion ................................................................................. 147<br />

7.1. Significance <strong>of</strong> Research ............................................................. 147<br />

7.2. Future Work ......................................................................... 149<br />

7.2.1. Coupled simulation <strong>of</strong> Fiber reinforced PNC .............................. 149<br />

7.2.2. Multi-scale modeling <strong>of</strong> Polymer Nanocomposite ......................... 151<br />

7.2.3. A Generalized Three-Dimensional Damage Model ........................ 153<br />

References ..................................................................................... 155<br />

A. Derivation <strong>of</strong> Stress-Strain law based on Dilatational and Deviatoric response<br />

<strong>of</strong> a material ............................................................................... 164<br />

viii


LIST OF TABLES<br />

3.1. Position <strong>of</strong> Nanoclay atoms in space group ........................................ 56<br />

6.1. Values <strong>of</strong> scaling parameter λ to determine ∆A ................................... 119<br />

6.2. Bond parameters for Nanoclay...................................................... 126<br />

6.3. Angle parameters for Nanoclay ..................................................... 126<br />

6.4. Torsion parameters for Nanoclay ................................................... 126<br />

6.5. Non-bond parameters for Nanoclay-Polymer hybrid .............................. 126<br />

6.6. Linear Regression and Analysis for PNC deformation Case 1 .................... 130<br />

6.7. Linear Regression and Analysis for PNC deformation Case 2 .................... 133<br />

6.8. Linear Regression and Analysis for PNC deformation Case 3 .................... 135<br />

6.9. Linear Regression and Analysis for PNC deformation Case 4 .................... 137<br />

6.10. Linear Regression and Analysis for PP deformation Case 1 ...................... 140<br />

6.11. Linear Regression and Analysis for PP deformation Case 2 ...................... 140<br />

6.12. Comparison <strong>of</strong> modulus <strong>of</strong> PP and PNC predicted by NDM ..................... 145<br />

6.13. Comparison <strong>of</strong> strength <strong>of</strong> PP and PNC predicted by NDM ..................... 145<br />

ix


LIST OF FIGURES<br />

1.1. Change in compressive strength <strong>of</strong> E-glass/polypropylene composite with<br />

nanoclay loading ..................................................................... 3<br />

1.2. Change in compressive modulus <strong>of</strong> E-glass/polypropylene composite with<br />

nanoclay loading ..................................................................... 4<br />

1.3. Change in compressive modulus <strong>of</strong> pure polypropylene with nanoclay loading . 4<br />

1.4. Length scales involved in multi-scale modeling .................................... 5<br />

1.5. <strong>The</strong> multi-scale modeling paradigm ................................................ 6<br />

2.1. <strong>The</strong> GIMP description .............................................................. 17<br />

2.2. Schematic <strong>of</strong> continuum deformation............................................... 18<br />

2.3. A two-dimensional GIMP particle .................................................. 20<br />

2.4. <strong>The</strong> one-dimensional characteristic particle shape function ...................... 20<br />

2.5. <strong>The</strong> one-dimensional grid interpolation function .................................. 23<br />

2.6. <strong>The</strong> one-dimensional GIMP interpolation function................................ 25<br />

2.7. <strong>The</strong> two-dimensional GIMP interpolation function ............................... 26<br />

2.8. <strong>The</strong> GIMP crack ..................................................................... 34<br />

2.9. Schematic for determining strain energy release rate in GIMP................... 35<br />

3.1. Common molecular dynamics force fields for polymers ........................... 45<br />

3.2. Schematic <strong>of</strong> bond stretching ....................................................... 46<br />

3.3. Schematic <strong>of</strong> angle bending ......................................................... 47<br />

3.4. Schematic <strong>of</strong> bond torsion........................................................... 48<br />

3.5. Schematic <strong>of</strong> non-bonded interactions .............................................. 49<br />

3.6. Schematic <strong>of</strong> van der Waals interaction energy .................................... 50<br />

3.7. Chemical formula <strong>of</strong> polypropylene................................................. 53<br />

3.8. Coarse grained polypropylene monomer ........................................... 53<br />

3.9. Pr<strong>of</strong>ile views <strong>of</strong> nanoclay in XY and XZ planes ................................... 55<br />

3.10. Calculation <strong>of</strong> free energy change using thermodynamic integration............. 61<br />

x


3.11. Schematic <strong>of</strong> the MD simulation model for polymer nanocomposite ............. 62<br />

3.12. Volumetric and deviatoric deformation <strong>of</strong> simulation box ........................ 63<br />

4.1. Schematic <strong>of</strong> Concurrent coupling .................................................. 66<br />

4.2. Schematic <strong>of</strong> the Direct Coupling methodology ................................... 67<br />

4.3. Schematic <strong>of</strong> the Statistical Coupling .............................................. 68<br />

4.4. Layout <strong>of</strong> Coupled Model ........................................................... 69<br />

4.5. Setup <strong>of</strong> atomistic model for Concurrent coupling ................................ 74<br />

4.6. Setup <strong>of</strong> ESCM/GIMP model for Concurrent coupling ........................... 75<br />

5.1. Development <strong>of</strong> nano-scale damage model (NDM) ................................ 78<br />

5.2. Characterization <strong>of</strong> damage entities ................................................ 80<br />

5.3. Representative volume element (RVE) and nano-clay orientation in damage<br />

model ............................................................................ 85<br />

5.4. Schematic <strong>of</strong> interface debond and void nucleation in damage model for<br />

deformations along principle axis ................................................... 86<br />

5.5. Schematic <strong>of</strong> interface debond in RVE under simple-shear deformation ......... 87<br />

5.6. Schematic <strong>of</strong> nano-clay with dimensions ........................................... 93<br />

5.7. Schematic <strong>of</strong> voids in pure polymers and polymer nano-composites ............. 94<br />

5.8. Determination <strong>of</strong> nano-scale damage parameters .................................. 96<br />

6.1. Layout <strong>of</strong> zones in coupled simulation ............................................. 99<br />

6.2. Layout <strong>of</strong> a non-periodic MD simulation box ...................................... 101<br />

6.3. Computation <strong>of</strong> displacement gradient ............................................. 101<br />

6.4. Schematic <strong>of</strong> coupled model in a simple tension setup ............................ 103<br />

6.5. Time progression <strong>of</strong> vertical displacements (v) for concurrent simulation<br />

(Case 1) .............................................................................. 104<br />

6.6. Time progression <strong>of</strong> displacement gradient (∂v/∂Y ) for concurrent simulation<br />

(Case 1) ....................................................................... 105<br />

6.7. Setup <strong>of</strong> coupled model for mode I crack growth .................................. 106<br />

6.8. Time progression <strong>of</strong> displacement gradient (∂v/∂Y ) for concurrent simulation<br />

(Case 2) ....................................................................... 107<br />

xi


6.9. Time progression <strong>of</strong> vertical displacement (v) for concurrent simulation<br />

(Case 2) .............................................................................. 108<br />

6.10. Fracture surface <strong>of</strong> mode I crack from concurrent simulation .................... 108<br />

6.11. Setup <strong>of</strong> coupled model for mode II crack growth ................................. 109<br />

6.12. Time progression <strong>of</strong> horizontal displacement (u) for concurrent simulation<br />

(Case 3) .............................................................................. 110<br />

6.13. Time progression <strong>of</strong> displacement gradient ∂u/∂Y for concurrent simulation<br />

(Case 3) ......................................................................... 111<br />

6.14. Time progression <strong>of</strong> displacement gradient ∂v/∂X for concurrent simulation<br />

(Case 3) ......................................................................... 112<br />

6.15. Schematic <strong>of</strong> polymer model for determination <strong>of</strong> bending stiffness EI ......... 113<br />

6.16. Dependence <strong>of</strong> loading rate on bending stiffness .................................. 114<br />

6.17. Schematic <strong>of</strong> the dilatational and deviatoric deformation used to determine<br />

free energy change associated with mechanical deformation ...................... 116<br />

6.18. Free-energy versus volumetric strain ............................................... 120<br />

6.19. Free-energy versus deviatoric strain ................................................ 121<br />

6.20. Schematic <strong>of</strong> primary RVE deformations in MD .................................. 123<br />

6.21. Structure <strong>of</strong> nano-clay with dimensions ............................................ 125<br />

6.22. σ11 versus ɛ11 for PNC deformation (Case 1) ...................................... 130<br />

6.23. σ22 versus ɛ11 for PNC deformation (Case 1) ...................................... 131<br />

6.24. σ33 versus ɛ11 for PNC deformation (Case 1) ...................................... 131<br />

6.25. Deformed and undeformed configurations <strong>of</strong> PNC deformation (Case 1) ........ 132<br />

6.26. σ11 versus ɛ22 for PNC deformation (Case 2) ...................................... 133<br />

6.27. σ22 versus ɛ22 for PNC deformation (Case 2) ...................................... 133<br />

6.28. σ33 versus ɛ22 for PNC deformation (Case 2) ...................................... 134<br />

6.29. Deformed and undeformed configurations <strong>of</strong> PNC deformation (Case 2) ........ 134<br />

6.30. σ11 versus ɛ33 for PNC deformation (Case 3) ...................................... 135<br />

6.31. σ22 versus ɛ33 for PNC deformation (Case 3) ...................................... 136<br />

6.32. σ33 versus ɛ33 for PNC deformation (Case 3) ...................................... 136<br />

xii


6.33. σ23 versus γ23 for PNC deformation (Case 4) ...................................... 137<br />

6.34. σ13 versus γ13 for PNC deformation (Case 5) ...................................... 138<br />

6.35. σ12 versus γ12 for PNC deformation (Case 6) ...................................... 139<br />

6.36. σ11 versus ɛ22 for PP deformation (Case 1) ........................................ 141<br />

6.37. σ22 versus ɛ22 for PP deformation (Case 1) ........................................ 141<br />

6.38. σ33 versus ɛ22 for PP deformation (Case 1) ........................................ 142<br />

6.39. Deformed and undeformed configurations <strong>of</strong> PP deformation (Case 1) .......... 142<br />

6.40. Schematic <strong>of</strong> void formation in PP under simple shear deformation ............. 143<br />

6.41. σ12 versus γ12 for PP deformation (Case 2) ........................................ 143<br />

6.42. Deformed and undeformed configurations <strong>of</strong> PP deformation (Case 2) .......... 144<br />

7.1. Morphology <strong>of</strong> a typical Polymer nano-composite................................. 152<br />

7.2. Schematic <strong>of</strong> randomly oriented nano-clay ......................................... 153<br />

xiii


Chapter 1<br />

Introduction<br />

<strong>The</strong> use <strong>of</strong> polymer matrix composites (PMC) in aerospace and civil engineering, as<br />

well as in sports and leisure applications is rapidly increasing. PMC has found wide range<br />

<strong>of</strong> applications in the structural components where the substitution <strong>of</strong> PMC for metals has<br />

substantially improved performance and reliability. Among the most important features<br />

in PMC, the following unique properties make it a very attractive engineering material:<br />

high specific static and fatigue strength, as high as eight times that <strong>of</strong> steels, and high<br />

corrosion resistance.<br />

Carbon fiber-reinforced polymer matrix composites are now being used widely in<br />

aerospace industry for structural components. <strong>The</strong> new Boeing 787 and the Airbus A380<br />

are cases in point. In addition to their extensive use in aircraft, composites are becoming<br />

a material <strong>of</strong> choice in the automotive, marine, civil infrastructure construction and<br />

consumer products industries. Several major manufacturers are prototyping concept cars<br />

almost totally with composite materials. Other current examples include race cars, racing<br />

boats, top-<strong>of</strong>-the-line motorcycles, bicycles and snowmobiles just to name a few. Potential<br />

new application areas include all-composite bridge I-beams, bridge deck, dry board and<br />

other building components.<br />

<strong>The</strong> high strength in PMC derives from the high strength in the fibers embedded in<br />

the matrix. Fibers have high strength in tension, however, their compressive strength<br />

is generally much lower due to the fact that under compression, the fibers tend to fail<br />

through micro-buckling well before compressive fracture occurs in conventional compos-<br />

ites. Matrix material has to contribute significantly to the overall compressive strength<br />

<strong>of</strong> the PMC, as matrix strength is much lower than that <strong>of</strong> fibers. <strong>The</strong> overall com-<br />

pressive strength <strong>of</strong> the PMC is, in general, only about 50% <strong>of</strong> the tensile strength or<br />

1


lower. Similarly, significant improvement in the interlaminar shear strength <strong>of</strong> a compos-<br />

ite can be achieved using nanoparticle reinforcement, since it is also a matrix dominated<br />

property. As most composite structures are used under multiaxial stress states, typically<br />

involving compression and/or shear in certain orientations, the lower compressive and/or<br />

shear strength is a major obstacle for the full realization <strong>of</strong> light-weight yet high-strength<br />

potential <strong>of</strong> composite applications.<br />

1.1 Motivation<br />

Preliminary results on nanoclay reinforced pultruded E-glass/polypropylene (PP) test<br />

specimens published by Roy et al. (2007) indicates significant improvements in the com-<br />

pressive strength as well as compressive modulus <strong>of</strong> nanoclay reinforced E-glass/PP com-<br />

posite, as shown in fig.(1.1) and fig.(1.2). Significant enhancements in compressive mod-<br />

ulus <strong>of</strong> neat polypropylene resin was also observed (fig.1.3). Interestingly, more than<br />

100% improvement in mechanical properties (<strong>of</strong> both neat resin and composite) was ob-<br />

served with only 10 wt% addition <strong>of</strong> nanoclay. Further, more than 60% improvement in<br />

interlaminar shear strength was recorded for 3 wt% nanoclay loading (Roy et al., 2007).<br />

While the synergistic reinforcing influence <strong>of</strong> nanoparticle reinforcement is obvious, a<br />

simple rule-<strong>of</strong>-mixtures approach fails to quantify the dramatic increase in mechanical<br />

properties. Consequently, there is an immediate need to investigate and understand<br />

the mechanisms at the nanoscale that are responsible for such unprecedented strength<br />

enhancements. It is envisioned that a better understanding <strong>of</strong> the mechanisms at the<br />

nanoscale will lead to optimization <strong>of</strong> processing variables at the macroscale, which in<br />

turn will lead to the manufacture <strong>of</strong> nanocomposites more efficiently and at lower cost.<br />

Figure 1.4 shows the length scales involved in a typical nanoparticle reinforced fiber com-<br />

posite. <strong>The</strong> nanoclay reinforcements are nanometer (10 −9 meter) thick particles that<br />

interact intimately with the polymer matrix molecules to provide strengthening <strong>of</strong> the<br />

2


Figure 1.1: Change in compressive strength <strong>of</strong> E-glass/polypropylene composite with<br />

nanoclay loading<br />

polymer matrix at the nano-scale. <strong>The</strong> size <strong>of</strong> the intercalating polymer molecules are<br />

roughly on the order <strong>of</strong> the nanoclay thickness, depending on the degree <strong>of</strong> polymeriza-<br />

tion. Moving up the scale we find that the carbon fiber diameter is usually on the order<br />

<strong>of</strong> 5 to 10 microns (1 micron, or 1µm = 10 −6 meter), although its length may extend<br />

up to several meters. <strong>The</strong> composite (matrix reinforced with fiber) is in the macro or<br />

engineering-scale <strong>of</strong> analysis. <strong>The</strong> combination <strong>of</strong> fiber reinforcements in the polymer ma-<br />

trix gives the macro-scale composite orthotropic properties (Jones, 1998). It is evident<br />

from the discussions above that the nanoscale interaction between polymer molecules<br />

and nanoclay-species is a crucial factor in determining the macro-scale strength <strong>of</strong> the<br />

composite. Hence, a satisfactory theoretical model must appeal to the physics <strong>of</strong> the<br />

problem at the nano-scale and at the same time be robust enough to predict perfomance<br />

at a macroscale. During material fracture, the reinforcing forces ahead <strong>of</strong> the crack tip<br />

3


Figure 1.2: Change in compressive modulus <strong>of</strong> E-glass/polypropylene composite with<br />

nanoclay loading<br />

Figure 1.3: Change in compressive modulus <strong>of</strong> pure polypropylene with nanoclay loading<br />

4


Figure 1.4: Length scales involved in multi-scale modeling<br />

is governed by the complex inter play between non-bonded interactions and the driving<br />

forces that want to extend the crack at the nanoscale. While, the influence <strong>of</strong> atomistic<br />

forces can be simulated to a satisfactory extent with robust molecular dynamics based<br />

algorithms (Allen & Tildesley, 1987), the technique is usually intractable when it comes<br />

to the study <strong>of</strong> larger systems under the influence <strong>of</strong> an external forcing environment. On<br />

the other hand, in a purely continuum analysis, the micro-structural detail <strong>of</strong> the ma-<br />

terial is overlooked. In the interest <strong>of</strong> lowering computational expense, an approximate<br />

phenomenological model is usually used to describe the microscale processes. Most con-<br />

tinuum scale models however, do not address the obvious length and time scale issues in<br />

fracture processes. This deficiency could undermine predictions in material behavior. A<br />

robust multiscale simulation technique will be cognizant <strong>of</strong> the discrete material behavior<br />

at atomistic level and will correlate microstructural phenomena to macroscale loading or<br />

environmental condition. However, the challenge is to seamlessly overcome disparities<br />

in both time and length scales and transfer information efficiently. As can be seen in<br />

fig.(1.5), in the modeling <strong>of</strong> atomistic systems there has been mainly two computational<br />

methodologies <strong>of</strong> interest to researchers: quantum mechanics (QM) based calculations<br />

5


Figure 1.5: <strong>The</strong> multi-scale modeling paradigm<br />

are used to simulate atomic (or molecular) systems, however as system sizes increase the<br />

molecular dynamics (MD) method has proven to be far more effective and efficient com-<br />

putationally although, not as accurate (Allen & Tildesley, 1987). In this research, the<br />

molecular domain is simulated under free boundary conditions acting under the influence<br />

<strong>of</strong> external forces from the macro-scale using molecular dynamics. Molecular dynamics<br />

is a subset <strong>of</strong> computational chemistry codes which creates system trajectories consistent<br />

with thermo-dynamic ensembles (Allen & Tildesley, 1987). <strong>The</strong> molecular dynamics pro-<br />

gram LAMMPS (large-scale atomic/molecular massively parallel simulator) (Plimpton,<br />

1995) is a scalable parallel computation algorithm with the added functionality to model<br />

polymer systems. In general the results to an atomistic MD computation (in particular<br />

polymer models) are system size dependent. However, with parallel MD algorithms such<br />

as LAMMPS, large (over 10,000 atoms) molecular systems can be simulated and hence<br />

6


makes it a more attractive choice over other approaches such as ab-initio calculations,<br />

which in general are limited to the calculation <strong>of</strong> a few 100 atoms (Leach, 1996). At the<br />

engineering scale <strong>of</strong> a coupled simulation most researchers have used the finite element<br />

analysis (FEA) to model continuum response (see e.g. Liu et al. (2006)). In this research,<br />

the continuum will be modeled using the generalized interpolation material point method<br />

or GIMP (Bardenhagen & Kober, 2004). GIMP is a class <strong>of</strong> arbitrary Euler-Lagrange<br />

(AEL) based meshless particle-in-cell method hence it automatically avoids mesh en-<br />

tanglement and distortion problems that usually plague FEA during simulation <strong>of</strong> large<br />

deformation. Furthermore, since GIMP discretizes the model into material points, it<br />

forms a natural connection to the discrete molecular system to which it is coupled (Liu<br />

et al., 2006). Using an implicit time-stepping strategy in GIMP one can also eliminate<br />

the errors associated with explicit GIMP dynamics (<strong>Nair</strong> & Roy, 2009).<br />

1.2 Literature Review<br />

As described above the need for low cost, low weight and multifunctional materials<br />

has propelled interest in studying advanced materials and in particular nano-composites<br />

(Alivisatos et al., 1998). <strong>The</strong> vast size differences in a “multi-scale” system allows for the<br />

improvement <strong>of</strong> material properties through reinforcement at the molecular level. In gen-<br />

eral, polymers can either be thermoplastics or thermosets. <strong>The</strong>rmoplastic polymers are<br />

tougher and has better impact resistance compared to thermosets. <strong>The</strong> harder, more rigid<br />

and brittle behavior <strong>of</strong> thermosets are a manifestation <strong>of</strong> the strong covalent cross-links<br />

between polymer chains which makes the macromolecule resistant to chain deformation<br />

and rearrangement. <strong>The</strong> molecular morphology in a thermoset also affects its post cure<br />

characteristics. In general cured thermoset polymers do not s<strong>of</strong>ten on re-heating, and<br />

will only char and break down at high temperatures. <strong>The</strong>rmoplastics on the other hand,<br />

owing to its weak inter molecular forces <strong>of</strong> attraction in between chains is amenable to re-<br />

7


molding and reshaping on the application <strong>of</strong> heat. <strong>The</strong>se properties renders thermoplastic<br />

polymers recyclable and feasible to applications in green technology. This research will<br />

focus on thermoplastic polymers and the effect <strong>of</strong> nanoclay platelets on the thermoplastic<br />

polymer matrix.<br />

Beginning in the early 90’s there has been an interest in the application <strong>of</strong> nanoclay<br />

reinforced polymers as structural materials. Researchers at Toyota (Kojima et al., 1993)<br />

were the first to use nanoclay reinforced thermoplastic (nylon-6) resin in manufacturing<br />

car bumpers. <strong>The</strong>y had observed a significant improvement in mechanical and impact<br />

properties <strong>of</strong> the composite (polymer + nanoclay) with low weight additions <strong>of</strong> nanoclay.<br />

<strong>The</strong> improvement in mechanical properties provided impetus for research on nanoclay<br />

reinforced composites for multi-functional applications. It has been shown that nanoclay<br />

reinforced polymers had improved barrier properties (Berta et al., 2006; Golebiewski<br />

& Galeski, 2007), moisture diffusion (Sun et al., 2007) in addition to enhancement <strong>of</strong><br />

mechanical properties (<strong>Nair</strong> et al., 2002; Subramaniyan & Sun, 2006a,b; Roy et al., 2007;<br />

Yalcin et al., 2008; Dong & Bhattacharyya, 2008). Application <strong>of</strong> nanoclay has also been<br />

extended to various resin systems including epoxies (Park & Jana, 2003; Timmerman<br />

et al., 2002), high density polyethylene (HDPE)/wood composite (Lei et al., 2007). In this<br />

research we are focused on understanding the reasons behind significant improvement in<br />

compressive and shear dominated properties <strong>of</strong> the matrix due to nanoclay reinforcement.<br />

<strong>The</strong> strength and stiffness enhancements provided by fiber reinforced composites with<br />

improved matrix dominated properties (such as compression and shear behavior) makes<br />

it an attractive choice for engineering structural applications.<br />

Montmorillonite (MMT) clay occurs in nature in the form <strong>of</strong> platelets with relatively<br />

large aspect ratios (∼100:1). <strong>The</strong> nanoclay allows both for an intercalated and/or ex-<br />

foliated morphology in the material system in which its dispersed (Mani et al., 2005).<br />

<strong>The</strong> enhancement <strong>of</strong> mechanical properties <strong>of</strong> the nanocomposite has been found to be<br />

8


strongly correlated to the quality <strong>of</strong> dispersion <strong>of</strong> nanoclay in the resin system (Chen<br />

et al., 2004). In general most naturally occurring nanoclays are organophobic and hy-<br />

drophyllic hence, nanoclays are surfactant modified to render them organophyllic. <strong>The</strong><br />

extent <strong>of</strong> nanoclay reinforcement is a function <strong>of</strong> its spatial distribution within the host<br />

polymer and the compatability <strong>of</strong> surfactant species with the polymer molecules (Yang &<br />

Tsai, 2006). In addition to the decreased processability, a high weight percent <strong>of</strong> nanoclay<br />

also degrades the mechanical performance <strong>of</strong> the nanocomposite due to the formation <strong>of</strong><br />

agglomerates in the polymer matrix. On the other hand, in a well dispersed system the<br />

polymer molecules tend to ingress into gallery spacings in between individual nanoclay<br />

platelets. Due to topological constraints imposed by the nanoclay, the local dynamics<br />

(i.e. chain relaxation, segmental mobility) <strong>of</strong> the polymer molecules is subdued (Krish-<br />

namoorti et al., 1996). It has been realised that the intercalation <strong>of</strong> the polymer molecule<br />

into the nanoclay galleries is entirely energetic (Krishnamoorti et al., 1996; Yang & Tsai,<br />

2006) and a favorable interaction between polymer and nanoclay is required to overcome<br />

the entropic loss due to confinement <strong>of</strong> polymer molecules.<strong>The</strong> small thickness <strong>of</strong> the<br />

nanoclay implies a larger surface area that allows for enhanced load transfer through<br />

non-bonded interactions at the polymer-particle interface.<br />

Yang & Tsai (2006) obeserved that the decreased mobility <strong>of</strong> the polymer molecules<br />

suggests that the viscoelastic response <strong>of</strong> the polymer could be used as an indicator <strong>of</strong><br />

dispersion <strong>of</strong> nanoclay in the polymer system. <strong>The</strong> hypothesis is verified by the enhanced<br />

creep resistance <strong>of</strong> polymer nanocomposites (Yang et al., 2006). <strong>The</strong> introduction <strong>of</strong> nan-<br />

oclay into the polymer also affects phase transformations and crystallinity <strong>of</strong> the polymer<br />

system. Yuan & Misra (2006); Yuan et al. (2008) has observed that the introduction <strong>of</strong><br />

nanoclay act as nucleation sites for the potential growth <strong>of</strong> spherulites (spherical crystal-<br />

lites) in the bulk material. Yuan & Misra (2006) observed that the growth <strong>of</strong> spherulites<br />

is correlated to the change in the primary mechanism <strong>of</strong> failure (in dynamic, IZOD<br />

9


tests) which is plastic deformation from crazing and vein-type in neat polypropylene to a<br />

microvoid-coalescence-fibrillation process in the nanocomposite. In addition to the effect<br />

<strong>of</strong> surfactant modification on nanoclay reinforcement the effects <strong>of</strong> nanoclay stacking and<br />

orientation on the mechanical properties <strong>of</strong> thermoplastic polymers is well documented<br />

(Galgalia et al., 2004; Looa & Gleason, 2004; Srinath & Gnanamoorthy, 2005).<br />

In recent years, numerous modeling methodologies has been proposed to provide better<br />

insight into the physics <strong>of</strong> nanoclay reinforcement. Zeng et al. (2008) provides an extensive<br />

review <strong>of</strong> the work that has been done in this field. In studying the system behavior<br />

<strong>of</strong> nanocomposites molecular dynamics (MD) has been the primary method <strong>of</strong> choice.<br />

Kuppa et al. (2003) used MD simulation to study poly-ethylene oxide (PEO) /nanoclay<br />

hybrid structures. <strong>The</strong>y were able to correlate the results from the analysis with wide-<br />

angle neutron diffraction (WAND) and differential scanning calorimetry (DSC) studies<br />

<strong>of</strong> purely intercalated PEO in MMT. Sinsawat et al. (2003) used coarse-grained MD to<br />

study the influence <strong>of</strong> polymer matrix architecture on nanoclay intercalation. Atomistic<br />

simulations have also been able to show that the effective dispersion and intercalation is<br />

strongly dependent on the functionalization <strong>of</strong> the constituent groups in the nanoclay,<br />

polymer matrix (Smith et al., 2003; Minisini & Tsobnang, 2005; Katti et al., 2006).<br />

Simpler theoretical models have also been suggested to explain the dynamics <strong>of</strong> melt<br />

intercalation in polymer nanocomposites (Vaia & Giannelis, 1997). Tsai & Sun (2004)<br />

used continuum shear lag to predict load transfer efficiency <strong>of</strong> nanoclay platelets. <strong>The</strong><br />

stress transfer between the (polymer) matrix and nanoclay inclusion was found to be<br />

dependent on the degree <strong>of</strong> exfoliation <strong>of</strong> the nanoclay in the polymer matrix.<br />

For the purpose <strong>of</strong> engineering applications, constitutive modeling <strong>of</strong> nanoclay rein-<br />

forced polymers is <strong>of</strong> much interest to researchers. Drozdov et al. (2010) developed a<br />

visco-elastic continuum model <strong>of</strong> polypropylene matrix with nanoclay fillers. <strong>The</strong> model<br />

was able to capture the retardation <strong>of</strong> stress relaxation with increase in nanoclay content.<br />

10


Sheng et al. (2004) determined the elastic modulii <strong>of</strong> MXD6-clay and nylon-6 compos-<br />

ites using a micromechanical model introduced in Danielsson et al. (2002). <strong>The</strong> analysis<br />

was able to incorporate geometry parameters <strong>of</strong> the hybrid structure such as the par-<br />

ticle aspect ratio (L/t) and silicate layer gallery spacing (d001) in determining material<br />

properties <strong>of</strong> the nanocomposite. <strong>The</strong> model was able to closely predict experimental<br />

observations for moduli better than Halpin-Tsai (HT) or Mori-Tanaka (MT) based cal-<br />

culation. Hbaieb et al. (2007) used MT to predict stiffness <strong>of</strong> randomly oriented nanoclay<br />

polymer composites. Zhu & Narh (2004) calculated effective tensile modulus <strong>of</strong> nanoclay<br />

polymers using a micromechanics representative volume element (RVE) model <strong>of</strong> aligned<br />

nanoclay platelets in a polymer matrix. <strong>The</strong>ir analysis included an “interface” region<br />

between the nanoclay and the bulk polymer to simulate actual load transfer conditions<br />

at the atomistic level. While the material models suggested above were successful in<br />

predicting (at least qualitatively) the enhancements <strong>of</strong> modulus for nanocomposites, the<br />

biggest drawback is the assumption <strong>of</strong> a well bonded continous interface at the nanoscale<br />

between nanoclay and polymer material. In reality, the length scale issues are unavoid-<br />

able, a realistic model <strong>of</strong> the material must explicitly account for the discrete material<br />

behavior especially at interfaces. <strong>The</strong> objective <strong>of</strong> this research is to incorporate the<br />

interactions at the interfaces and its associated properties by multiscale modeling.<br />

In studying molecular systems, a multiscale and multi-physics simulation approach is<br />

considered computationally more viable since relying on a single time or length scale can<br />

take a huge amount <strong>of</strong> computational resources and can potentially lead to inaccurate<br />

results. In recent years numerous efforts have been directed toward modeling nanocom-<br />

posites in order better understand the reasons behind the enhancement in mechanical<br />

properties, even with slight addition (a few weight percent) <strong>of</strong> nano-materials (Valavala<br />

et al., 2007; Buryachenko et al., 2005; Riddick et al., 2006b,a; Abraham et al., 1999; Ogata<br />

et al., 2001; Wagner & Liu, 2003; Xiao & Belytschko, 2004; Ma et al., 2006).<strong>The</strong>re are<br />

11


two unique choices in multiscale modeling, namely, the hierarchical and concurrent sim-<br />

ulation schemes. In the hierarchical approach the physical system is studied in isolation<br />

to far-field stimuli and the results are translated to a continuum response using curve fits<br />

and/or statistical averaging. Valavala et al. (2007) used energy equivalence <strong>of</strong> continuum<br />

and atomistic models <strong>of</strong> polymer systems to characterize the hyperelastic stress-strain<br />

response <strong>of</strong> polymers (in particular, polycarbonate and polyimide). Buryachenko et al.<br />

(2005) used the Eshelby and Mori-Tanaka methods to determine the effective proper-<br />

ties <strong>of</strong> nanocomposite materials. Riddick et al. (2006a), employed molecular dynamics<br />

(MD) to determine the force-displacement curves between nano inclusions (CNT) and the<br />

polymer system. Riddick et al. (2006b) used equivalent modeling <strong>of</strong> carbon nanotubes<br />

in polymers to study the fracture toughness <strong>of</strong> polymer matrix composites with carbon<br />

nanotubes (CNTs) embedded in them.<br />

In general, it was observed that the strengthening behavior observed in nanocomposite<br />

materials due to nano sized inclusions can only be explained using theoretical formalisms<br />

that appeal to fundamental physics <strong>of</strong> interactions at nanometer length-scales. True ma-<br />

terial behavior cannot be accurately simulated through effective or “smeared” continuum<br />

analysis since these techniques assume a continuous and “perfect” bonding behavior <strong>of</strong><br />

the nanocomposite system or in some cases “smears away” the interface altogether. By<br />

definition the “smeared” analyses fail to capture the local details which are responsible<br />

for the stiffness and strength enhancements at the nanoscale. It is evident that the stress-<br />

strain response <strong>of</strong> the system is a function <strong>of</strong> localized behavior and that the strength<br />

enhancement is primarily due to attenuation <strong>of</strong> internal damage due the complex interac-<br />

tion <strong>of</strong> voids/stress concentrators with the nanosized inclusions in the material. It is for<br />

this reason that regular continuum material models that are commonly used in engineer-<br />

ing cannot accurately predict the material response <strong>of</strong> nanocomposite systems. From a<br />

design standpoint, it is necessary that we are able to describe the material response with<br />

12


no prior assumptions in order to reliably and accurately model responses <strong>of</strong> the material<br />

under all loads and ambient conditions.<br />

Advances in computational tools has allowed us to examine and understand small<br />

scale phenomena using molecular dynamics and macro-scale responses using continuum<br />

level simulations. However, due to limited computer resources and the inefficiency and<br />

inaccuracy involved in using a single method to analyze large (or small) scale events has<br />

led to the coupling <strong>of</strong> two (or more) methods <strong>of</strong> analysis. For example, the continuum<br />

method (such as finite element analysis (FEA)), can be used to solve the governing<br />

continuum field equations subject to given traction and displacement boundary conditions<br />

and then apply the forces and tractions from its analysis as boundary conditions to the<br />

region described by atomic interactions (e.g. molecular dynamics). By taking advantage<br />

<strong>of</strong> this synergy between these vastly different methods we get the appropriate resolution<br />

to examine extremely localized events such as crack initiation, dislocations etc.<br />

However, the challenge is to simultaneously process all information available from<br />

both length scales and efficiently pass this information continually across both time and<br />

length scales. Concurrent simulations have the unique advantage that it can tie local<br />

events at nanoscale to stimuli from a larger time and length scale during real time simu-<br />

lation. One <strong>of</strong> the earliest forays into concurrent coupling was by Abraham et al. (1999)<br />

in their scheme, MAAD. <strong>The</strong> scheme involved simultaneous coupling <strong>of</strong> tight-binding<br />

(TB), molecular dynamics (MD) and finite element analysis (FEA). <strong>The</strong> scheme was<br />

successfully applied to study fracture behavior <strong>of</strong> silicon systems. However, the method<br />

had some shortcomings in it in that it was limited by the timestep <strong>of</strong> the tight-binding<br />

step (on the order <strong>of</strong> 10 −15 seconds) which made the procedure computationally demand-<br />

ing. Ogata et al. (2001) used FEA to concurrently couple with density function theory<br />

(DFT) to study the surface oxidation <strong>of</strong> silicon system. Some <strong>of</strong> the more recent efforts in<br />

multiscale concurrent simulations are the bridging-scale technique (Wagner & Liu, 2003;<br />

13


Xiao & Belytschko, 2004) and the MD-MPM coupling technique by (Ma et al., 2006). In<br />

Ma et al. (2006), the atoms within the handshake region <strong>of</strong> the MD domain are coupled<br />

point-wise to MPM particles. <strong>The</strong> continuum is hence refined down to atomic sizes and<br />

the boundary material points are assumed to overlap with boundary atoms <strong>of</strong> the MD<br />

zone at all times. <strong>The</strong> outcome <strong>of</strong> this refinement is that the molecular dynamics domain<br />

induces spurious high frequency oscillations into the continuum domain as the simula-<br />

tion proceeds due to thermal vibrations <strong>of</strong> the atoms. Recently, Saether et al. (2009),<br />

proposed a statistical averaging technique to overcome the high frequency vibrations, in<br />

their concurrent coupling scheme to study grain boundary structures in aluminum.<br />

Although there have been many attempts to model standard solid lattice structures<br />

and couple them to various continuum methods, to our knowledge, large scale concur-<br />

rent coupling <strong>of</strong> nanoparticle reinforced polymer systems has not been attempted before<br />

partly due to the complexity involved in modeling polymers. Under ambient conditions<br />

the polymer model is amorphous (depending on the degree <strong>of</strong> crystallinity) and lacks<br />

specific ordered structure. Further, a polymer MD model takes longer time to equilibri-<br />

ate as compared with an ordered lattice. In this dissertation we propose two approaches<br />

to modeling nanocomposites. First, we propose to couple the generalized interpolation<br />

material point method (continuum scale) and molecular dynamics (discrete scale) in sim-<br />

ulating atomic interactions in polymer systems using the statistical averaging technique<br />

proposed by Saether et al. (2009) in order to visualize the nanoscale failure mechanism<br />

in correlation with macroscale mechanical loads. Next we propose a continuum dam-<br />

age mechanics (CDM) based equivalent 3D RVE model <strong>of</strong> nanoclay-polymer that can be<br />

incorporated in to continuum scale analysis to capture the strength <strong>of</strong> nanoclay compos-<br />

ites as a function <strong>of</strong> interface degradation or micro-void nucleation. We envision that<br />

the work in this paper will contribute towards the understanding <strong>of</strong> nanoscale interac-<br />

tions in nanostructured composite materials. <strong>The</strong> dissertation is organized as follows: In<br />

14


Chapter 2, the GIMP algorithm is discussed along with its application to modeling crack<br />

propagation. In Chapter 3 the MD algorithm is introduced along with a technique to de-<br />

termine free energy from molecular simulations <strong>of</strong> polymer (or polymer nanocomposite)<br />

systems. <strong>The</strong> coupling algorithm for coupling MD and GIMP using the embedded statis-<br />

tical scheme (ESCM) is discussed in Chapter 4. In Chapter 5, a hierarchical simulation<br />

method is proposed to model stiffness and damage behavior <strong>of</strong> nanoclay composites. In<br />

Chapter 6, results from the proposed analyses methods are presented, followed by a list<br />

<strong>of</strong> conclusions and recommendations for future research, in Chapter 7.<br />

15


Chapter 2<br />

Generalized Interpolation Material Point Method (GIMP)<br />

<strong>The</strong> material point method (MPM) was first suggested by Sulsky et al. (1995). It<br />

belongs to a class <strong>of</strong> particle-in-cell (PIC) algorithms first developed at Los Alamos Na-<br />

tional laboratories. MPM was initially proposed to model and solve complex problems<br />

in fluid flows however its mixed Euler-Lagrange formulation has made the method vi-<br />

able to applications in finite-deformation solid mechanics (Sulsky et al., 1994; Sulsky &<br />

Schreyer, 1996). In particular, it provides as an alternative to FEA which is susceptible to<br />

numerical instabilities like mesh entanglement in large deformation simulations. In recent<br />

years, MPM has increasingly found application in a wide variety <strong>of</strong> engineering problems<br />

such as fluid-structure interactions (York et al., 2000; Guilkey et al., 2003), simulating<br />

granular materials (Zhang et al., 2009; Wieckowski, 2004), studying crack growth (Tan<br />

& <strong>Nair</strong>n, 2002; Daphalapurkar et al., 2007), analysis <strong>of</strong> multi-phase flows (Gilmanov &<br />

Acharya, 2008; Zhang et al., 2008) and multi-scale modeling (Ma et al., 2005; Lu et al.,<br />

2006; Ayton et al., 2001).<br />

<strong>The</strong> material point method (MPM) is a class <strong>of</strong> arbitrary Euler-Lagrange (AEL)<br />

scheme. <strong>The</strong> solution method uses Lagrangian material points embedded in an Eulerian<br />

mesh. <strong>The</strong> use <strong>of</strong> Lagrangian material points simultaneously avoids the need to advect<br />

state variables (such as stress and strain) and the numerical diffusion problems associated<br />

with a purely Eulerian mesh (see, fig.2.1). <strong>The</strong> Eulerian mesh on the other hand, ensures<br />

that mesh entanglement is avoided. However, certain drawbacks such as cell-crossing<br />

anomalies, that are primarily attributed to shortcomings in the original MPM interpo-<br />

lating functions are a major concern in applying MPM to real world large deformation<br />

problems. In the classic MPM algorithm the equations have to be re-ordered and written<br />

in terms <strong>of</strong> momentum in order to account for problems with particles crossing cells <strong>of</strong><br />

16


Figure 2.1: <strong>The</strong> GIMP description<br />

the grid in case <strong>of</strong> large deformations. This leads to unsymmetric equations that have to<br />

be solved on the background grid. <strong>The</strong> generalized interpolation material point method<br />

(GIMP) as proposed by Bardenhagen & Kober (2004), explicitly accounts for the shape<br />

<strong>of</strong> the particle and the background grid structure in formulating the interpolation basis<br />

functions. In this way GIMP naturally overcomes the cell crossing anomalies <strong>of</strong> MPM,<br />

and provides a better time description <strong>of</strong> the solution.<br />

<strong>The</strong> advantages <strong>of</strong> using a meshless particle method in couple atomistic-continuum<br />

simulation are many as listed in Ma et al. (2005) and Lu et al. (2006). In particular,<br />

the particle description <strong>of</strong> the material form a natural transition from a continuum to<br />

the discrete atoms in the handshake region. In GIMP, the exchange <strong>of</strong> information<br />

between the discrete atomistic system and the continuous domain is enhanced through<br />

the use <strong>of</strong> the finite sized particles. In this Chapter the theoretical foundation for GIMP<br />

17


is introduced and an numerical implementation <strong>of</strong> the explicit time-stepping algorithm<br />

is discussed. <strong>The</strong> equations <strong>of</strong> motion and GIMP discretization is described in section<br />

2.1. In section 2.2 an algorithmic implementation <strong>of</strong> the GIMP equations and the Verlet<br />

explicit time-stepping strategy in GIMP is developed. Finally, section 2.3 discusses the<br />

application <strong>of</strong> GIMP in studying crack propagation problems.<br />

2.1 Equations <strong>of</strong> Motion <strong>of</strong> a Deformable body<br />

Figure 2.2: A solid undergoing continuous deformation φ(X, t) which maps every point<br />

X in the reference body to x on the deformed body<br />

Consider a solid deformable body, undergoing continuous deformation φ(X, t) which<br />

maps the equilibrium (initially undeformed; zero stress state) to a deformed volume Ω<br />

at time ‘t’. Let x be the current coordinates <strong>of</strong> a point X on the undeformed body such<br />

that x ≡ x(X, t). <strong>The</strong> deformation causes the body to develop a stress state σ(x, t)<br />

18


where σ is the Cauchy stress tensor. <strong>The</strong> following equations for conservation <strong>of</strong> mass<br />

and momentum must be satisfied by the deformed body,<br />

dρ<br />

dt<br />

+ ρ∇ · v = 0 (2.1)<br />

ρa = ∇ · σ + ρb (2.2)<br />

Where ρ is the current mass density, a is the acceleration <strong>of</strong> the body, v is the velocity<br />

and b is the body force. It should be noted that d/dt denotes the material derivative.<br />

<strong>The</strong> weak form <strong>of</strong> the momentum equation can be found by multiplying eq.(2.2) by an<br />

admissible velocity test function δv and integrating the resulting equation over the current<br />

volume Ω <strong>of</strong> the body gives,<br />

<br />

Ω<br />

<br />

ρa · δvdΩ +<br />

Ω<br />

<br />

σ · ∇δvdΩ =<br />

Ω<br />

<br />

ρb · δvdΩ +<br />

∂Ωτ<br />

τ · δvdS (2.3)<br />

∂Ωv is the part <strong>of</strong> the body where velocity is specified while ∂Ωτ is the part <strong>of</strong> the<br />

surface where a traction force is specified. <strong>The</strong> current boundary ∂Ω <strong>of</strong> the volume Ω is<br />

the union <strong>of</strong> the two disjoint set <strong>of</strong> surfaces, i.e., ∂Ω = ∂Ωv ∪ ∂Ωτ.<br />

2.1.1 Discretizing governing equation<br />

<strong>The</strong> following section is a review <strong>of</strong> the discretization procedure in GIMP. For further<br />

details on the scheme, please refer to Bardenhagen & Kober (2004). Like MPM in GIMP,<br />

the particle is characterized by the particle characteristic function χp. In MPM, this<br />

function is simply the delta function (Sulsky et al., 1995). However, in GIMP the particle<br />

function is chosen such that the particle is given a discrete shape and simultaneously<br />

satisfies the partition <strong>of</strong> unity (Bardenhagen & Kober, 2004). Shown in fig.(2.3) is a<br />

typical (rectangular) GIMP particle (in two-dimensions) with Cartesian coordinates <strong>of</strong><br />

the centroidal location given by xp = (x (1)<br />

p , x (2)<br />

p ) and (l (1)<br />

p , l (2)<br />

p ) are the half-lengths <strong>of</strong> the<br />

19


particle in each direction.<br />

X2<br />

X1<br />

xp<br />

(x (1)<br />

p , x (2)<br />

p )<br />

2l (1)<br />

p<br />

2l (2)<br />

p<br />

Figure 2.3: A two-dimensional GIMP particle<br />

In GIMP, the simplest choice <strong>of</strong> particle characteristic function is:<br />

χp(x) = H(x − (xp − lp)) + H(x − (xp + lp)) (2.4)<br />

Where, H(x) is the unit step function, xp is the particle centroid position while lp is<br />

Figure 2.4: <strong>The</strong> one-dimensional characteristic particle shape function<br />

the half length <strong>of</strong> the particle. Figure (2.4) is the plot <strong>of</strong> the one-dimensional particle<br />

shape function. We can extend the one-dimensional formula to further dimensions using<br />

20


a multiplicative decomposition, i.e., for a space coordinate x = (x1, x2, x3), we have<br />

χp(x) = χp(x1) · χp(x2) · χp(x3).<br />

We can now define particle properties using the characteristic shape function eq.(2.4),<br />

<br />

Vp =<br />

Ω<br />

χp(x)dΩ (2.5)<br />

Where, Ω is the volume <strong>of</strong> the continua at current time ’t to be discretized. Using this<br />

definition the particle mass and momenta can be defined as follows, if ρ is the current<br />

density field <strong>of</strong> the body, then,<br />

<br />

mp =<br />

<br />

pp =<br />

It follows from above that ρp = mp<br />

Vp and vp = p p<br />

mp<br />

Ω<br />

Ω<br />

ρ(x)χp(x)dΩ (2.6)<br />

ρ(x)v(x)χp(x)dΩ (2.7)<br />

is the particle mass density and<br />

velocity, respectively. <strong>The</strong> following discretization procedure ensures the conservation <strong>of</strong><br />

mass and momenta (Bardenhagen & Kober, 2004). In GIMP for any material point data<br />

fp, the continuous representation <strong>of</strong> data f(x) is possible using the characteristic material<br />

point function, i.e.,<br />

f(x) = <br />

fpχp(x) (2.8)<br />

p<br />

We can now re-write eq.(2.3) using the rate <strong>of</strong> change <strong>of</strong> particle momentum density<br />

( ˙p p<br />

Vp ), Cauchy stress σp <strong>of</strong> the particle and above mentioned material point representation<br />

eq.(2.8).<br />

21


p<br />

Ω∩Ωp<br />

˙p p<br />

χp(x) · δvdΩ + <br />

<br />

Vp<br />

= <br />

<br />

p<br />

Ω∩Ωp<br />

mp<br />

Vp<br />

p<br />

Ω∩Ωp<br />

<br />

χp(x)b · δvdΩ +<br />

σpχp(x) : ∇δvdΩ<br />

∂Ωτ<br />

τ · δvdS (2.9)<br />

Where, Ωp is the domain <strong>of</strong> the particle. Similar to the representation <strong>of</strong> particle data,<br />

eq.(2.8), continuous data on the grid data can be represented using grid interpolation<br />

functions. For example if g(x) is a continuous variable and gi is the grid nodal data,<br />

then,<br />

g(x) = <br />

giSi(x) (2.10)<br />

i<br />

Where, Si(x) is the grid interpolation function, A bi-linear (in two-dimensions) shape<br />

function is used in the formal MPM methodology Sulsky et al. (1995). Shown in Fig.(2.5)<br />

is a one-dimensional tent function. As before, the shape function in 3-dimensions is<br />

written as, Si(x) = Si(x1) · Si(x2) · Si(x3). <strong>The</strong> interpolation function a point ‘x’ to a<br />

node ‘i’ <strong>of</strong> coordinate Xi in a regular grid <strong>of</strong> cell spacing ‘L’ is given by,<br />

⎧<br />

⎪⎨<br />

Si(x) =<br />

⎪⎩<br />

0 x − Xi ≤ −L<br />

1 + x−Xi<br />

L<br />

1 − x−Xi<br />

L<br />

−L < x − Xi ≤ 0<br />

0 < x − Xi ≤ L<br />

0 x − Xi > L<br />

22<br />

(2.11)


Figure 2.5: <strong>The</strong> one-dimensional grid interpolation function<br />

Using eq.(2.10) we can now write the admissible velocity field as a summation <strong>of</strong> the<br />

nodal velocities, δv = <br />

i δvi. Substituting, this approximation into eq.(2.9) and taking<br />

into account the arbitrariness <strong>of</strong> the nodal velocities, we arrive at the discrete form <strong>of</strong><br />

equations that needs to be solved at nodes <strong>of</strong> the background grid,<br />

˙p i = f int<br />

i + f ext<br />

i<br />

23<br />

(2.12)


Where,<br />

˙p i = <br />

p<br />

f int<br />

i = − <br />

Sip ˙p p<br />

p<br />

f ext<br />

i = f b<br />

i + f τ<br />

i<br />

f b<br />

i = <br />

f τ<br />

i<br />

=<br />

<br />

p<br />

∂Ωτ<br />

σp · ∇SipVp<br />

mpbpSip<br />

τ Si(x)dS<br />

(2.13)<br />

In the above equation ˙p i, f int<br />

i and f ext<br />

i are nodal rate <strong>of</strong> momentum, internal and external<br />

forces, respectively. <strong>The</strong> external force on the node can be due to a spatially varying<br />

traction force on the particles τ (x) acting on surface ∂Ωτ or a body force b acting on<br />

the volume enclosed by Ω. In the above equations, Sip and ∇Sip represent the volume<br />

averaged GIMP interpolation function and derivative. <strong>The</strong>y are defined as follows (in<br />

one-dimension),<br />

Sip(xp) = 1<br />

=<br />

2lp<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

xp+lp<br />

xp−lp<br />

Si(x)dx (2.14)<br />

0 xp − Xi ≤ −L − lp<br />

(L+lp+(xp−Xi)) 2<br />

4Llp<br />

1 + xp−Xi<br />

L<br />

1 − (xp−Xi) 2 +l 2 p<br />

2Llp<br />

1 − xp−Xi<br />

L<br />

(L+lp−(xp−Xi)) 2<br />

4Llp<br />

−L − lp < xp − Xi ≤ −L + lp<br />

−L + lp < xp − Xi ≤ −lp<br />

−lp < xp − Xi ≤ lp<br />

lp < xp − Xi ≤ L − lp<br />

L − lp < xp − Xi ≤ L + lp<br />

0 L + lp < xp − Xi<br />

24


∇Sip(xp) = 1<br />

2lp<br />

xp+lp<br />

xp−lp<br />

∇Si(x)dx (2.15)<br />

In the above equations, ∇ ≡ ∂/∂xp. Shown in fig.(2.6) is the plot <strong>of</strong> the GIMP<br />

interpolation function in one-dimension. <strong>The</strong> extension <strong>of</strong> the interpolation function to<br />

three dimensions is simply given by, Sip = Sip(x (1)<br />

p ) · Sip(x (2)<br />

p ) · Sip(x (3)<br />

p ). For example in<br />

fig.(2.7) the two-dimensional interpolation function is shown for a particle located at the<br />

origin <strong>of</strong> a uniformly spaced grid.<br />

Figure 2.6: <strong>The</strong> one-dimensional GIMP interpolation function<br />

25


Figure 2.7: <strong>The</strong> two-dimensional GIMP interpolation function<br />

2.2 Explicit Time Integration in GIMP<br />

In this section the algorithmic implementation <strong>of</strong> GIMP is discussed. <strong>The</strong> solution strat-<br />

egy follows explicit time-stepping and hence numerical stability and accuracy <strong>of</strong> the al-<br />

gorithm is time-step (∆t) dependent. For time-steps lower than the CFL (Courant et al.,<br />

1928) condition for elastic wave propagation in the material good numerical accuracy can<br />

be achieved (Reddy, 2006; Bathe, 1982).<br />

At the beginning <strong>of</strong> timestep ‘t’ the mass mp, velocity v t p, stress σ t p external and<br />

internal forces on particle ‘p’ is extrapolated to its nearest node neighbors,<br />

26


Note that V t<br />

p = mp<br />

ρ t p<br />

m t i = <br />

p<br />

p t i = <br />

p<br />

f int,t<br />

i = − <br />

f ext,t+∆t<br />

i<br />

=<br />

<br />

Sipmp<br />

Sipmpv t p<br />

p<br />

∂Ωτ<br />

σ t p∇SipV t<br />

p<br />

(2.16)<br />

(2.17)<br />

(2.18)<br />

τ t+∆t Si(x)dS (2.19)<br />

is the volume <strong>of</strong> the particle at the beginning <strong>of</strong> the timestep with<br />

ρ t p as its current mass density. <strong>The</strong> momenta on the grid is then updated using a forward<br />

differencing scheme.<br />

p t+∆t<br />

i<br />

= p t i + ∆tf tot,t+∆t<br />

i<br />

Where, the total force on the grid is given by f tot,t+∆t<br />

i<br />

= f int,t<br />

i<br />

the grid update the particle positions and velocities are updated next.<br />

x t+∆t<br />

p = x t p + ∆t <br />

v t+∆t<br />

p = v t p + ∆t <br />

i<br />

i<br />

Sip<br />

Sip<br />

p t+∆t<br />

i<br />

m t i<br />

f tot,t+∆t<br />

i<br />

m t i<br />

(2.20)<br />

ext,t+∆t<br />

+ f . Following<br />

i<br />

(2.21)<br />

(2.22)<br />

Note that for large deformation problems, the particle lengths in each direction<br />

(l (1)<br />

p , l (2)<br />

p ) is tracked by the GIMP algorithm as the material deforms. <strong>The</strong> advantages<br />

<strong>of</strong> tracking lengths are two-fold: the particle volumes are tracked and it circumvents<br />

particle-splitting (e.g., Ma et al. (2006)) in finite deformation simulations. <strong>The</strong> imple-<br />

mentation <strong>of</strong> corner tracking as suggested by Ma et al. (2006) alleviates this issue to a<br />

large extent. In order to update particle stresses, the gradient <strong>of</strong> the particle displace-<br />

27


ments have to be determined. This gradient can be found using gradient <strong>of</strong> the GIMP<br />

particle shape function and nodal velocity field <strong>of</strong> the background grid.<br />

Where, ∂∆up<br />

∂x t<br />

∂∆up<br />

<br />

= ∆t<br />

∂xt i<br />

∇Sip<br />

p t+∆t<br />

i<br />

m t i<br />

(2.23)<br />

is a two-tensor and the gradient <strong>of</strong> the incremental displacement field<br />

with respect to the particle coordinates (x t p) at the beginning <strong>of</strong> the time-step ‘t’ (Bathe,<br />

1982). For finite deformation, the incremental deformation gradient can be determined<br />

from,<br />

dFp = I + ∂∆up<br />

∂x<br />

(2.24)<br />

Where, I is the second order identity tensor. <strong>The</strong> total deformation gradient is given by<br />

the multiplicative decomposition (which follows from the the chain rule <strong>of</strong> differentiation<br />

(Malvern, 1969)),<br />

F t+∆t<br />

p<br />

= F t<br />

p · dFp<br />

For small deformation, the displacement gradient can be approximated as,<br />

∂ut+∆t p<br />

∂X = ∂utp ∂∆up<br />

+<br />

∂X ∂xt <strong>The</strong> small strain tensor follows from above,<br />

ɛ t+∆t<br />

p<br />

(2.25)<br />

(2.26)<br />

= 1<br />

⎡<br />

⎤<br />

T<br />

t+∆t<br />

t+∆t<br />

∂u<br />

⎣ p ∂up +<br />

⎦ (2.27)<br />

2 ∂X ∂X<br />

Where, T is the transpose <strong>of</strong> the deformation tensor. <strong>The</strong> Cauchy stress (Malvern,<br />

1969) on the particle can be updated using the deformation gradient eq.(2.25) or eq.(2.26)<br />

depending on the severity <strong>of</strong> the deformation coupled with the material constitutive law.<br />

28


For a linearly elastic isotropic material with Lame constants λ and µ undergoing small<br />

deformation, the cauchy stress on the particle can be updated using,<br />

Where, e t+∆t<br />

p<br />

σ t+∆t<br />

ij,p<br />

= λet+∆t p δij + 2µɛ t+∆t<br />

ij,p<br />

(2.28)<br />

= tr(ɛt+∆t p ) the trace <strong>of</strong> the small deformation strain tensor and δij is<br />

the Kronecker delta function. After the particle state variables are updated, the grid<br />

is reset while the particle positions are updated using eq.(2.21). While the Lagrangian<br />

material points are “convected” in this fashion, the background grid is restored to its<br />

initial (undeformed) state. <strong>The</strong> arbitrary Euler-Lagrange solution strategy naturally<br />

overcomes issues involved with mesh entanglement in large deformation problems.<br />

2.2.1 A Finite Deformation Hyper-Elastic Model for Polymeric<br />

Materials<br />

In this dissertation a hyperelastic material model is assumed to simulate large deformation<br />

behavior <strong>of</strong> thermoplastic polymers. Consider the following strain energy potential “W”<br />

(from Valavala et al. (2007)) for a compressible hyperelastic material.<br />

W = c1 (I3 − 1) 2 + c2<br />

<br />

I1<br />

I 1/3 +<br />

3<br />

I3 2<br />

I2 3<br />

− 30<br />

<br />

(2.29)<br />

Where, I1, I2 and I3 are the invariants <strong>of</strong> the right Cauchy-Green deformation tensor<br />

C (C = F T F, where F is the deformation gradient) in continuum mechanics (Malvern,<br />

1969). <strong>The</strong>y are defined as follows,<br />

29


I1 = tr(C) (2.30)<br />

I2 = 1 2 2<br />

[tr(C)] − tr(C )<br />

2<br />

(2.31)<br />

I3 = det(C) (2.32)<br />

Where, tr(C) and det(C) is the trace and determinant <strong>of</strong> tensor C. <strong>The</strong> Lagrangian<br />

strain tensor E is defined as follows,<br />

E = 1<br />

(C − I) (2.33)<br />

2<br />

<strong>The</strong> 2 nd Piola-Kirch<strong>of</strong>f stress S is energy conjugate to the Lagrangian strain tensor E.<br />

Hence we have,<br />

S = ∂W<br />

∂E<br />

= ∂W<br />

∂C<br />

· ∂C<br />

∂E<br />

= 2∂W<br />

∂C<br />

(2.34)<br />

Substituting eq.(2.29) into eq.(2.34), and employing the chain rule <strong>of</strong> differenciation<br />

we get,<br />

S = 2<br />

<br />

<br />

6c1I3(I3 − 1) − c2<br />

3<br />

+2c2<br />

<br />

1<br />

I 1/3<br />

3<br />

I1<br />

I 1/3 + 6<br />

3<br />

I3 2<br />

I2 3<br />

+ 3I1I 2 2<br />

I 2 3<br />

<br />

<br />

C −1<br />

I<br />

I − 6c2<br />

2 2<br />

I2 C (2.35)<br />

3<br />

Where, the following identities were employed in the above derivation,<br />

∂I1<br />

∂C<br />

= I<br />

∂I2<br />

∂C = I1I − C (2.36)<br />

∂I3<br />

∂C<br />

=<br />

−1<br />

I3C<br />

30


<strong>The</strong> Cauchy stress in the current configuration can be found using the Piola transfor-<br />

mation:<br />

σ = 1<br />

J FSFT<br />

σ = 1<br />

<br />

<br />

J<br />

2<br />

3<br />

6c1I3(I3 − 1) − c2<br />

I1<br />

I 1/3<br />

+2c2<br />

<br />

1<br />

I 1/3<br />

3<br />

+ 3I1I 2 2<br />

I 2 3<br />

<br />

3<br />

+ 6 I3 2<br />

I 2 3<br />

<br />

I<br />

B − 6c2<br />

2 2<br />

I2 B<br />

3<br />

2<br />

I<br />

<br />

(2.37)<br />

(2.38)<br />

Where, J = det(F) is the Jacobian <strong>of</strong> the deformation gradient F, B is the left<br />

Cauchy-Green tensor (B = FF T ). As can be seen, J = √ I3 = det(C) = det(B).<br />

2.2.2 <strong>The</strong> Verlet Integration in GIMP<br />

As observed in the earlier section the internal force on the nodes eq.(2.18) will always<br />

“lag” behind the current external force hence the dynamic force equilibrium is satisfied<br />

only in the limit <strong>of</strong> small time-step in an explicit time-stepping strategy. A better esti-<br />

mate <strong>of</strong> the internal nodal force vector is possible through implicit iterations where the<br />

error in solution per time-step is reduced by iterating to a pre-decided error norm (<strong>Nair</strong><br />

& Roy, 2009). However, in the interest <strong>of</strong> expediting calculations during a multi-scale<br />

coupled simulation an alternate explicit time-stepping strategy is employed in this re-<br />

search, namely the “Verlet” integration scheme wherein a mid-step internal force vector<br />

is determined.<br />

<strong>The</strong> algorithm begins with interpolating particle state variables to nodes <strong>of</strong> back-<br />

ground grid using equations (2.16) - (2.19). A temporary velocity field is determined<br />

using the following equation,<br />

v ∗ i =<br />

p t i + ∆t<br />

2<br />

<br />

f int,t<br />

i<br />

31<br />

m t i<br />

<br />

ext,t+∆t<br />

+ f<br />

i<br />

(2.39)


<strong>The</strong> temporary nodal velocity field is used to update the positions and the deformation<br />

gradients <strong>of</strong> the material points, using equations (2.21) and (2.23).<br />

x t+∆t<br />

p = x t p + ∆t <br />

∂∆u ∗ p<br />

∂x<br />

= ∆t <br />

i<br />

i<br />

∇Sipv ∗ i<br />

Sipv ∗ i<br />

(2.40)<br />

(2.41)<br />

<strong>The</strong> “mid” timestep deformation gradient is used to create the mid-interval deforma-<br />

tion gradient,<br />

F t∗<br />

p = F t<br />

<br />

p · I + ∂∆u∗ <br />

p<br />

∂x<br />

(2.42)<br />

<strong>The</strong> updated Cauchy stress will follow from the updated deformation gradient eq.(2.42)<br />

and the material constitutive law. For an isotropic material undergoing small deformation<br />

the Cauchy stress is updated using eq.(2.28). However, in this dissertation a hyper-elastic<br />

material model is assumed to simulate large deformation behavior <strong>of</strong> thermoplastic poly-<br />

mers. <strong>The</strong> Cauchy stress is updated using equations (2.41), (2.25) and (2.38). <strong>The</strong><br />

updated Cauchy stress is used to update the internal nodal force vector.<br />

f int,t∗<br />

i<br />

= − <br />

p<br />

σ t∗<br />

p ∇SipV t<br />

p<br />

(2.43)<br />

In the above equation t ∗ denotes the mid-step update <strong>of</strong> nodal and particle variables.<br />

It should be further noted that since the equations <strong>of</strong> motion on the grid are solved using<br />

an updated Lagrangian formulation (Bathe, 1982) the gradient <strong>of</strong> the GIMP interpolation<br />

function ∇Sip is with respect to the coordinate <strong>of</strong> the particle and node at the beginning<br />

<strong>of</strong> time-step ‘t’, i.e., x t p and Xi. <strong>The</strong> total force and nodal momentum is next updated,<br />

32


f tot,t+∆t<br />

i = f int,t∗<br />

i + f ext,t+∆t<br />

i<br />

p t+∆t<br />

i = m t iv ∗ i + ∆t<br />

2<br />

f tot,t+∆t<br />

i<br />

(2.44)<br />

(2.45)<br />

<strong>The</strong> updated momentum and force is used to update the particle velocities and stresses<br />

using equations (2.22), (2.23), (2.24), (2.25) and (2.38). For a compressible hyper-elastic<br />

material with initial density ρ 0 , the particle densities can updated using,<br />

Where, J t+∆t<br />

p<br />

= det(F t+∆t<br />

p ).<br />

ρ t+∆t<br />

p = ρ0<br />

J t+∆t<br />

p<br />

2.3 Dynamic Fracture Analysis using GIMP<br />

(2.46)<br />

An important application <strong>of</strong> the GIMP algorithm is in the modeling <strong>of</strong> crack propagation<br />

and failure analysis <strong>of</strong> structures. Contact and friction properties <strong>of</strong> crack surfaces can<br />

be easily modeled by incorporating the necessary normal and tangential behavior <strong>of</strong> the<br />

crack surface into the GIMP analysis. Since the GIMP methodology is mesh independent,<br />

crack extension and propagation in GIMP is easier to simulate than conventional FEA<br />

(Duflot et al., 2010). By construction, the Lagrangian particles track evolution <strong>of</strong> state<br />

variables (such as stress, strain, temperature etc.) within the bulk <strong>of</strong> the continuum hence<br />

in a crack propagation situation the particles could be also used to track the extension<br />

<strong>of</strong> crack surface. This unique property allows easier definition <strong>of</strong> crack since the analysis<br />

grid does not have to be remeshed in order to accommodate new crack geometry, instead<br />

the material point field variables can be adjusted during simulation time to include the<br />

new singularity (see, <strong>Nair</strong>n (2003)).<br />

In this research the crack in material point (CRAMP) algorithm by <strong>Nair</strong>n (2003) is<br />

33


Figure 2.8: <strong>The</strong> GIMP crack<br />

employed to track and simulate propagation <strong>of</strong> crack surface. <strong>The</strong> CRAMP algorithm has<br />

been used to calculate dynamic strain energy release rates (SERR) using virtual crack<br />

closure technique (VCCT) (Krueger, 2004) and the dynamic (i.e., with inertial effects<br />

included) version <strong>of</strong> the J-integral (Guo & <strong>Nair</strong>n, 2004). <strong>The</strong> CRAMP algorithm allows<br />

for the definition <strong>of</strong> multiple nodal velocity fields for those material points that lie along<br />

the crack surface, this approach implicitly allows the definition <strong>of</strong> discontinuous stress<br />

and displacement field around the vicinity <strong>of</strong> the crack. In CRAMP the crack surface<br />

is defined as line segments (in 2D) or planes (in 3D). At the beginning <strong>of</strong> the GIMP<br />

update the position <strong>of</strong> the material point relative to the crack surface is determined using<br />

a line crossing algorithm. This is illustrated as follows: consider for example, the crack<br />

schematic in fig.(2.8). According to the CRAMP formulation an imaginary line is drawn<br />

from material point mp1 to node 1, since the line intersects the crack surface from above<br />

34


we accordingly assign that material point mp1 is on “top” <strong>of</strong> the crack surface similarly<br />

we find that mp2 is at the “bottom” in its position relative to the crack surface. Next<br />

in order to account for the discontinuity (crack) multiple velocity fields are introduced<br />

at nodes 1 and 2 (and other nodes in the neighborhood <strong>of</strong> the crack). At these special<br />

nodes the two equations <strong>of</strong> motion are solved; those for the GIMP particles above the<br />

crack surface and for those below the crack surface. At the other nodes (such as node 3<br />

in fig.2.8) the regular gimp update follows. Further details on the scheme can be found<br />

in <strong>Nair</strong>n (2003).<br />

Figure 2.9: Schematic for determining strain energy release rate in GIMP<br />

In this work the SERR at the crack tip is calculated using the VCCT approach (Tan<br />

& <strong>Nair</strong>n, 2002). Figure 2.9 is a schematic <strong>of</strong> mode I crack in GIMP. In the schematic<br />

points 1 and 1’ are nodes behind the crack tip located on top and bottom <strong>of</strong> the crack<br />

surface respectively. Note that the displacements are exaggerated to show the opening <strong>of</strong><br />

the crack. Further, it should be understood that according to the CRAMP scheme points<br />

35


1 and 1’ are the same node but with multiple velocity fields there exists a discontinuity in<br />

the velocity at this node (as indicated by the unequal displacements, uy(1) and uy(1 ′ ))a<br />

unique velocity field. <strong>The</strong> crack tip is located at node 2 and node 3 is located ahead <strong>of</strong><br />

the crack tip. For a traction free crack, the crack closure integral in mode I is given by,<br />

GI = lim<br />

δa→0<br />

1<br />

2δa<br />

δa<br />

0<br />

σyy(x)∆uy(x − δa)dx (2.47)<br />

Similarly the mode II SERR (GII) can be determined using the following virtual crack<br />

closure integral,<br />

GII = lim<br />

δa→0<br />

1<br />

2δa<br />

δa<br />

0<br />

σxy(x)∆ux(x − δa)dx (2.48)<br />

To determine the nodal stress values the interpolation scheme suggested in Tan &<br />

<strong>Nair</strong>n (2002) can be used. Writing the stresses as a linear interpolation between nodal<br />

values at nodes (2) and (3) and the displacements as linear interpolations between nodes<br />

(1) or (1’) and (2), and substituting into the crack closure integral leads to,<br />

GI = u(1) y − u (1′ )<br />

y<br />

(2σ (2)<br />

yy + σ (3)<br />

yy ) (2.49)<br />

12<br />

GII = u(1) x − u (1′ )<br />

x<br />

(2σ (2)<br />

xy + σ (3)<br />

xy ) (2.50)<br />

12<br />

36


Chapter 3<br />

Molecular Dynamics (MD)<br />

<strong>The</strong> classical dynamical equations <strong>of</strong> motion are valid for slow and heavy particles with<br />

typical velocities v ≪ c where, c is the speed <strong>of</strong> light and masses m ≫ me, me being the<br />

electron mass (Liu et al., 2006). <strong>The</strong>refore a typical molecular dynamics (MD) simulation<br />

can predict trajectories <strong>of</strong> a molecular system evolving under the influence <strong>of</strong> internal<br />

forces which the constituent entities <strong>of</strong> the atomistic system exert on each other. Thus,<br />

molecular dynamics algorithm finds application in many areas <strong>of</strong> atomistic simulation. It<br />

has been used to calculate stiffness <strong>of</strong> novel engineering materials such as nanoclay (Katti<br />

et al., 2005; Suter et al., 2007, 1995), carbon nanotube (CNT) (Frankland et al., 2002;<br />

Liew et al., 2004), transport properties such as thermal conductivity (Maruyama, 2002;<br />

Abramson et al., 2002; Berber et al., 2000; Volz & Chen, 1999), electrical conductivity<br />

(Yuge et al., 2005; Rey-Castro & Vega, 2006; Tang et al., 2001) and viscosity (Erpenbeck,<br />

1984; Hess, 2002). Other popular molecular simulation techniques includes ab initio based<br />

approaches such as density functional theory (DFT), tight-binding (TB) calculations (see<br />

e.g.Leach (1996)). However, first principle calculations come at a high computational<br />

cost. Molecular dynamics based methods on the other hand can leverage state-<strong>of</strong>-the-<br />

art computing to study larger molecular systems. Billion atom simulations with MD is<br />

possible under parallel computing scenarios (Vashishta et al., 1996; Abraham et al., 2002;<br />

Rountree et al., 2002).<br />

<strong>The</strong> efficiency enhancements in MD are solely due to the empirical energy functions<br />

that are used to describe the inter-atomic interaction. However, these enhancements come<br />

at the price <strong>of</strong> accuracy <strong>of</strong> trajectory prediction. In general, MD is unable to accurately<br />

predict long term responses <strong>of</strong> the atomistic system due to quantuum errors and curve<br />

fitting approximations <strong>of</strong> force field parameters. Some authors have hence, proposed cou-<br />

37


pling quantuum mechanics based methods to molecular dynamics e.g.(Abraham et al.,<br />

1999; Khare et al., 2007; Qian et al., 2008) in the process zone to achieve a better pre-<br />

diction <strong>of</strong> material behavior.<br />

In studying the system behavior <strong>of</strong> nano-composites, molecular dynamics (MD) has<br />

been the primary method <strong>of</strong> choice. Kuppa et al. (2003) used MD simulation to study<br />

poly-ethylene oxide (PEO) /nanoclay hybrid structures. <strong>The</strong>y were able to correlate the<br />

results from the analysis with wide-angle neutron diffraction (WAND) and differential<br />

scanning calorimetry (DSC) studies <strong>of</strong> purely intercalated PEO in MMT. Sinsawat et al.<br />

(2003) used coarse-grained MD to study the influence <strong>of</strong> polymer matrix architecture<br />

on nanoclay intercalation. In this chapter we focus on multi-scale modeling strategies<br />

for thermoplastic polymer nano-composites using the MD algorithm. We begin with<br />

description <strong>of</strong> statistical thermodynamics in the canonical (or NVT) ensemble and its<br />

applications to this research. Later, we derive the thermostatted equations <strong>of</strong> motion<br />

and the common force fields that are used in the modeling <strong>of</strong> polymers.<br />

3.1 Introduction<br />

Molecular dynamics simulations can be viewed as a bridge between the deterministic<br />

behavior <strong>of</strong> the macroscopic material system and the statistical nature <strong>of</strong> atomistic in-<br />

teractions. If we can imagine the position (r) and momenta (p) <strong>of</strong> a N atom MD system<br />

represented in a 6N dimensional space (r N , p N ) called phase space, during simulation<br />

time the MD system will occupy various regions <strong>of</strong> the phase space as the system evolves.<br />

We can imagine the collection <strong>of</strong> possible configurations in phase space (also called mi-<br />

crostates) <strong>of</strong> the atomistic system to lie inside a 6N dimensional hyper-sphere called<br />

an ensemble. <strong>The</strong> concept <strong>of</strong> an ensemble plays a central role in the determination <strong>of</strong><br />

macroscopic properties <strong>of</strong> the atomistic system. Consider for example the evolution <strong>of</strong> an<br />

observable F (r N , p N ) (F is any variable <strong>of</strong> interest; it may refer to the internal energy,<br />

38


Helmholtz energy, enthalpy etc.) for the N-atom system. <strong>The</strong>n according the postu-<br />

late by Gibbs and Boltzmann the macroscopic observable is an average over all possible<br />

micro-states in phase-space <strong>of</strong> the atomistic model. Mathematically this is written as,<br />

<br />

〈F 〉 =<br />

dp N dr N F (r N , p N )ρ(r N , p N ) (3.1)<br />

Where, dp N dr N is short-hand for the 6N space volume element,<br />

dp N dr N = dp 1 xdp 1 ydp 1 z . . . dp N x dp N y dp N z dr 1 xdr 1 ydr 1 z . . . dr N x dr N y dr N z ,the double integral alludes<br />

to the volume integral in 6N dimensional space and ρ(r N , p N ) is the probability associated<br />

with a particular micro-state. In a statistical sense, the ensemble average <strong>of</strong> the observable<br />

is the expectation value <strong>of</strong> the observable. In the molecular dynamics method the classical<br />

equations <strong>of</strong> motion are time integrated to simulate the behavior <strong>of</strong> the atomistic system<br />

in its phase space. Hence, to extract the macroscopic observable in MD, a time averaging<br />

has to be performed.<br />

F τ = 1<br />

τ<br />

τ<br />

F (r<br />

t=0<br />

N (t), p N (t))dt (3.2)<br />

In the limit <strong>of</strong> sufficiently long simulation time the following equality will hold,<br />

lim<br />

τ→∞ F τ = F = 〈F 〉 (3.3)<br />

This is the statement <strong>of</strong> the ergodic hypothesis. It is the fundamental assumption<br />

that bridges molecular dynamics and statistical mechanics. <strong>The</strong> probability distribution<br />

ρ(rN, p N) is <strong>of</strong> particular importance in statistical mechanics. For a NVT (constant<br />

number <strong>of</strong> particles N, constant volume V and temperature T) ensemble also called the<br />

canonical ensemble, the probability distribution is assumed to follow the Boltzmann<br />

distribution,<br />

39


ρ(r N , p N ) = exp −H(r N , p N )/kBT <br />

QNV T<br />

(3.4)<br />

Where, H(r N , p N ) is the total energy (Hamiltonian <strong>of</strong> the system) associated with a<br />

particular configuration and QNV T is the partition function for the NVT ensemble, kB is<br />

the Boltzmann constant and T is the temperature. <strong>The</strong> partition function for the NVT<br />

ensemble is written as follows,<br />

QNV T =<br />

1<br />

N!h 3N<br />

<br />

dp N dr N exp −H(r N , p N )/kBT <br />

(3.5)<br />

Where, N! is a factor used to prevent ‘over-count’ <strong>of</strong> the particle states and h is<br />

a quantuum-correction factor usually chosen to be the Plank’s constant . Once the<br />

partition function has been established, all thermodynamic properties can be defined.<br />

For example in a NVT ensemble, the internal energy (U) and Helmholtz free energy (A)<br />

can be derived from the partition function from eq.(3.5),<br />

U =<br />

=<br />

2 kBT ∂Q<br />

Q ∂T<br />

<br />

dp N dr N H(r N , p N ) exp −H(rN , pN )/kBT <br />

Q<br />

(3.6)<br />

(3.7)<br />

A = −kBT ln Q (3.8)<br />

<br />

= kBT ln dp N dr N <br />

N N H(r , p )<br />

exp<br />

ρ(r<br />

kBT<br />

N , p N <br />

) (3.9)<br />

While the statistical nature <strong>of</strong> thermodynamic quantities are well established, the<br />

calculation <strong>of</strong> these variables from simulation is dependent on the efficiency and quality<br />

<strong>of</strong> MD algorithm. In the following sections, the equations <strong>of</strong> motion in a thermostatted<br />

atomistic system are derived and the application <strong>of</strong> the MD algorithm is discussed.<br />

40


3.1.1 Equations <strong>of</strong> Motion<br />

From Newton’s law, we know that the classical equations <strong>of</strong> motion for a point mass is,<br />

mia τ i = −∇x τ i U + Fi, i=1,2,. . .,N (3.10)<br />

Where, mi, x τ i and a τ i could be considered as the mass, current position and acceler-<br />

ation vector <strong>of</strong> atom i at time τ, ∇x τ i is the spatial derivative in the three orthogonal<br />

space directions, Fi is a non-conservative or external force on the i th particle, and U is the<br />

energy <strong>of</strong> interaction <strong>of</strong> atom i with all other atoms in the molecular system. In section<br />

3.2, the various forms <strong>of</strong> energy interactions are discussed. For a non conservative system,<br />

vector force Fi could act as a dissipative force. For example, the force could represent a<br />

velocity dependent damping force,<br />

F S<br />

i = −miγv τ i (3.11)<br />

In the above equation, γ is the damping coefficient that adds a momentum dependent<br />

retardation on the atom ‘i’ while v τ i is the current velocity <strong>of</strong> atom ‘i’. Here, F S<br />

i is called<br />

the viscous, or Stoke’s friction force. In this dissertation, temperature control is achieved<br />

using the Langevin thermostat (Allen & Tildesley, 1987). <strong>The</strong> formulation is based on<br />

the addition <strong>of</strong> a stochastic external force Ri that represents thermal collisions <strong>of</strong> the<br />

atom ‘i’ with hypothetical solvent molecules. In the case <strong>of</strong> a thermally excited system<br />

the stochastic force will simulate Brownian dynamics <strong>of</strong> the atoms in the solute. <strong>The</strong><br />

thermostatted equation <strong>of</strong> motion is now,<br />

mia τ i = −∇x τ ,iU + Ri(τ), i=1,2,. . .,N (3.12)<br />

It should be noted that the stochastic force Ri is chosen such that it satisfies the<br />

41


following relationships,<br />

3.1.2 <strong>The</strong> Verlet Algorithm<br />

τ<br />

1<br />

lim Ri(t)dt = 0<br />

τ→∞ τ 0<br />

(3.13)<br />

τ<br />

1<br />

lim Ri(t) · Rj(t0 + t)dt = aδ(t0)δij<br />

τ→∞ τ 0<br />

(3.14)<br />

<strong>The</strong> typical molecular dynamics algorithm simulates evolution <strong>of</strong> the atomistic system in<br />

its phase space through time integration <strong>of</strong> the equations <strong>of</strong> motion. <strong>The</strong> time integration<br />

is performed for a given set <strong>of</strong> boundary (usually a periodic boundary condition, see Allen<br />

& Tildesley (1987)) and prescribed temperature and pressure conditions (or alternatively,<br />

prescribed temperature and volume). For a simulation algorithm to be successful, it needs<br />

to satisfy the following conditions (Allen & Tildesley, 1987),<br />

1. It should be fast, and require little memory.<br />

2. It should permit the use <strong>of</strong> a large time step ∆t.<br />

3. It should duplicate the classical trajectory as closely as possible.<br />

4. It should satisfy the known conservation laws <strong>of</strong> energy and momentum, and be<br />

time-reversible.<br />

5. It should be simple in form and easy to program.<br />

<strong>The</strong> Verlet algorithm (Verlet, 1967) is the most widely used method <strong>of</strong> integrating<br />

equations <strong>of</strong> motion. <strong>The</strong> method is based on current positions x τ , accelerations a τ , and<br />

the positions x τ−∆τ from the previous step. Using the Taylor series expansion <strong>of</strong> the<br />

vector functions around the known system configuration at time ‘τ’,<br />

x τ+∆τ = x τ + ∆τv τ + (1/2)∆τ 2 a τ + . . . (3.15)<br />

42


x τ−∆τ = x τ − ∆τv τ + (1/2)∆τ 2 a τ − . . . (3.16)<br />

Combining the above two equations we get the update for position as follows,<br />

x τ+∆τ = 2x τ − x τ−∆τ + ∆τ 2 a τ<br />

<strong>The</strong> velocities can be calculated as follows,<br />

v τ = xτ+∆τ − x τ−∆τ<br />

2∆τ<br />

(3.17)<br />

(3.18)<br />

We see that the error in eqn.(3.17) is <strong>of</strong> the order ∆τ 4 while the error in velocity<br />

(eq.3.18)is <strong>of</strong> the order ∆τ 2 . We can observe from eqn.(3.17) the position update is<br />

time-reversible (since, x τ+∆τ and x τ−∆τ play symmetrical roles). <strong>The</strong> algorithm however<br />

introduces some numerical imprecision since in update eq.(3.17), a small term O(∆τ 2 )<br />

is added to the difference <strong>of</strong> large terms O(∆τ 0 ). To improve on these shortcomings a<br />

half-step ‘leap-frog’ scheme (Hockney, 1970) was introduced. In this algorithm a mid-step<br />

velocity is determined as follows,<br />

x τ+∆τ = x τ ∆τ<br />

τ+<br />

+ ∆τv 2 (3.19)<br />

∆τ<br />

∆τ<br />

τ+ τ−<br />

v 2 = v 2 + ∆τa τ<br />

(3.20)<br />

<strong>The</strong> velocities for energy calculations are computed using a mid-time interval update,<br />

v τ+∆τ =<br />

∆τ<br />

∆τ<br />

τ−<br />

vτ+ 2 + v 2<br />

2<br />

(3.21)<br />

However, as evident from the above equation (eq.3.21) the leap-frog algorithm does<br />

not handle round-<strong>of</strong>f errors in predicting velocities . In order to overcome these errors<br />

a velocity Verlet algorithm was proposed (Swope et al., 1982). <strong>The</strong> algorithm uses the<br />

43


following update equations,<br />

x τ+∆τ = x τ + ∆τv τ + 1<br />

2 ∆t2 a τ<br />

v τ+∆τ = v τ + 1<br />

2 ∆τ a τ + a τ+∆τ<br />

(3.22)<br />

(3.23)<br />

In general, when the velocity Verlet algorithm is implemented a mid-step velocity is<br />

determined,<br />

∆τ<br />

τ+<br />

v 2 = v τ + ∆τ<br />

2 aτ<br />

<strong>The</strong> position and velocity updates then follow in a single update step,<br />

(3.24)<br />

x τ+∆τ = x τ ∆τ<br />

τ+<br />

+ ∆τv 2 (3.25)<br />

v τ+∆τ ∆τ<br />

τ+<br />

= v 2 + 1<br />

2 ∆τaτ+∆τ<br />

(3.26)<br />

At this point, the kinetic energy is calculated and the potential energy is calculated in<br />

the force loop. <strong>The</strong> velocity Verlet algorithm as implemented in the molecular dynamics<br />

algorithm LAMMPS (Plimpton, 1995) is the atomistic trajectory predictor algorithm<br />

used in this dissertation.<br />

3.2 Molecular Force Fields<br />

In general quantuum mechanics (QM) based methods provide a better time-description<br />

<strong>of</strong> atoms in phase space. However since the energy calculated in a QM calculation is<br />

a function <strong>of</strong> the motions and interactions <strong>of</strong> electrons within each atom the method is<br />

computationally restricted to solving smaller atomistic systems. <strong>The</strong> Born-Oppenheimer<br />

approximation allows the energy <strong>of</strong> the system to be written as a product <strong>of</strong> mutually<br />

independent functions <strong>of</strong> nuclear and electron coordinates (Leach, 1996). In MD, the<br />

44


position and configuration <strong>of</strong> electrons within the shell <strong>of</strong> atoms are ignored and focus<br />

is entirely on the energy due to the position <strong>of</strong> nuclei with respect to each other. This<br />

approximation allows MD to be robust and scalable to larger thermodynamic systems.<br />

<strong>The</strong> molecular force fields employed in MD algorithm are in general empirical force fits<br />

to quantuum mechanical calculations. In certain cases the molecular dynamics algorithm<br />

is able to provide results as good as the highest quantuum mechanical calculations, for a<br />

fraction <strong>of</strong> computer time (Leach, 1996). For a N-body system, the total energy in the<br />

system can be written as the sum <strong>of</strong> contributions from individual potentials,<br />

U(r) = <br />

Vij(ri, rj) + <br />

Vijk(ri, rj, rk) + . . . (3.27)<br />

i<br />

j<br />

Figure 3.1: Common molecular dynamics force fields for polymers<br />

i<br />

Where, Vij is a two-body interaction potential and Vijk is the potential due to three-<br />

body interaction. As discussed above molecular mechanics is based on simple interaction<br />

<strong>of</strong> nuclei centers, hence the energy <strong>of</strong> interactions can be written purely as a function<br />

<strong>of</strong> nuclei coordinates. In a polymer system, the energy contributions are due to bonded<br />

interactions such as bond stretching (Ebond), the opening and closing <strong>of</strong> angles between<br />

45<br />

j<br />

k


onds (Eang), bond torsions (Etors), and non-bonded interactions such as van der Waals<br />

forces (Enb). <strong>The</strong> total energy (U) for these components can be written as a sum <strong>of</strong> these<br />

individual contributions,<br />

U(r) = Ebond + Eang + Etors + Enb<br />

(3.28)<br />

In the following sections some the popular molecular force field forms are discussed.<br />

3.2.1 Bond stretching<br />

Figure 3.2: Schematic <strong>of</strong> bond stretching<br />

<strong>The</strong> energy due to bond stretch accounts for the contribution to the total configu-<br />

rational energy <strong>of</strong> the system due to stretching (or compressing) <strong>of</strong> the intra-molecular<br />

bonds during simulation. <strong>The</strong> simplest form <strong>of</strong> this energy force field is the harmonic<br />

potential given below.<br />

Ebond = <br />

bonds<br />

kl 2<br />

(l − l0)<br />

2<br />

(3.29)<br />

Where, kl is the stiffness <strong>of</strong> the bonds, l is the current length <strong>of</strong> the bond while l0<br />

is the ‘equilibrium’ length that the bond adopts when all other terms in the force field<br />

46


contribute (Leach, 1996). Note that in the summation above, the energy contribution<br />

from all bonds in the system are accounted for.<br />

3.2.2 Angle bendings<br />

Figure 3.3: Schematic <strong>of</strong> angle bending<br />

Angle bending interactions are usually included in describing molecules to include<br />

the energetic contribution <strong>of</strong> angles deviating from their equilibrium configurations. It is<br />

usually described by Hooke’s law,<br />

Eangle = <br />

angles<br />

kθ<br />

2<br />

(θ − θ0)<br />

2<br />

(3.30)<br />

Where, kθ is the stiffness <strong>of</strong> the bonds, θ is the current angle <strong>of</strong> the bond while θ0 is<br />

the ‘equilibrium’ angle.<br />

3.2.3 Bond torsion<br />

<strong>The</strong> bond stretch and angle bending accounts for the stiffer response in the molecular<br />

system to configurational change (Leach, 1996). In general, the conformational change<br />

<strong>of</strong> the polymer is attributed to the complex interplay between torsional and non-bonded<br />

47


Figure 3.4: Schematic <strong>of</strong> bond torsion<br />

interactions. <strong>The</strong> existence <strong>of</strong> energy barriers to rotation about chemical bonds is <strong>of</strong> prime<br />

importance in defining the structure property <strong>of</strong> the polymer molecule. <strong>The</strong> barrier to<br />

rotation arises from the repulsive forces in the non-bonded interactions between the end<br />

atoms <strong>of</strong> the torsional unit, hence in principle, it is possible to model structure changes<br />

through non-bonded interactions. Most torsional interactions are almost modeled using<br />

a cosine series expansion (Leach, 1996). One form <strong>of</strong> interaction is,<br />

Etors = <br />

N<br />

tors n=0<br />

Alternatively it could also be expressed as,<br />

Etors = <br />

Vn<br />

[1 + cos(nω − γ)] (3.31)<br />

2<br />

N<br />

tors n=0<br />

3.2.4 Non-bonded Interactions<br />

Cn cos(ω) n<br />

(3.32)<br />

Most physical properties <strong>of</strong> the molecular system can be explained by the non-bonded<br />

interactions between atoms <strong>of</strong> the physical system. <strong>The</strong> non-bonded interactions can<br />

be either due to the electro-static forces that occur between ionically charged atoms or<br />

48


Figure 3.5: Schematic <strong>of</strong> non-bonded interactions<br />

due to the quantuum-mechanical nuclear interactions between atoms. In this work, the<br />

electrostatic (or Columbic) interactions are ignored and hence, the contribution to non-<br />

bonded interactions are entirely attributed to the dispersive-repulsive forces called “van<br />

der Waals” (Allen & Tildesley, 1987) that quantifies the quantuum-statistical interactions<br />

between atoms <strong>of</strong> the molecular system. <strong>The</strong>se are short ranged nuclear forces with an<br />

influence zone in the vicinity <strong>of</strong> a few Angstroms. <strong>The</strong> van der Waal force is in general<br />

the sum <strong>of</strong> instantaneous attractive and repulsive forces between atoms (Leach, 1996).<br />

<strong>The</strong> dispersive or London force is an attractive force due to instantaneous induced dipoles<br />

with in the atoms. <strong>The</strong> repulsive forces on the other hand is primarily attributed to the<br />

forces arising from the Pauli exclusion principle. Figure (3.6) is the typical energy versus<br />

inter-atomic separation pr<strong>of</strong>ile for van der Waals interactions.<br />

<strong>The</strong> Lennard-Jones potential is a common empirical potential used to model non-<br />

bonded interactions and is given by,<br />

Enb = <br />

σ 4ɛ<br />

r<br />

i<br />

j>i<br />

12<br />

−<br />

<br />

σ<br />

<br />

6<br />

r<br />

(3.33)<br />

Where, r=|rij| is the distance between atoms i and j, ɛ is the depth <strong>of</strong> the potential<br />

well and σ is the distance at which inter-particle energy is zero.<br />

49


Figure 3.6: Pr<strong>of</strong>ile <strong>of</strong> van der Waals interaction energy versus interatomic separation<br />

3.3 Modeling <strong>The</strong>rmoplastic Polymer Nanocomposites in MD<br />

In the context <strong>of</strong> MD simulation, modeling <strong>of</strong> thermoplastic or thermoset polymers rep-<br />

resents significant computational challenges. Some <strong>of</strong> the more predominant challenges<br />

are listed below:<br />

1. A realistic description <strong>of</strong> material atomistic configuration at given ambient condition<br />

is inherently difficult. Most thermoplastic polymers have an amorphous (i.e., lack<br />

<strong>of</strong> specific long range order) configuration at room temperature. For example,<br />

polypropylene is semi-crystalline at room temperature i.e., it has regions <strong>of</strong> ordered<br />

lamellar crystal configuration embedded in a matrix <strong>of</strong> amorphous molecules. <strong>The</strong><br />

visco-elasticity <strong>of</strong> the polymer material is entirely attributed to the amorphous<br />

matrix while the crystal structure gives the material stiffness.<br />

50


2. <strong>The</strong> macro-scale behavior (transport properties, thermo-mechanical behavior etc.)<br />

<strong>of</strong> the polymer (or polymer nanocomposite) atomistic model is dependent on the<br />

force fields used in the simulation. Given in sec.(3.2) is a list <strong>of</strong> energy contribu-<br />

tions that are commonly considered in modeling polymers. Because there are few<br />

force fields designed specifically for in-organic polymers, there has been an issue <strong>of</strong><br />

transferability <strong>of</strong> force fields to study polymeric molecular systems.<br />

3. <strong>The</strong> equilibration (or energy relaxation) <strong>of</strong> polymer molecules to its minimum en-<br />

ergy takes longer compared with regular solid lattice structures. <strong>The</strong> difficulties in<br />

equilibrating structures arise from configurational as well as energetic sources <strong>of</strong> the<br />

polymer in the phase space that add to the free energy <strong>of</strong> the system. It is known<br />

that the parameters for minimum energy configuration is dependent on a multitude<br />

<strong>of</strong> parameters.<br />

Additional difficulties arise when nanoparticles such as nanoclay, is introduced into<br />

the polymer atomistic model. <strong>The</strong> challenge is to develop a computationally viable model<br />

that is able to address some <strong>of</strong> the issues listed above. In this dissertation, in the interest <strong>of</strong><br />

saving computational time we use a coarse grained polymer system using the parameters<br />

described in Wei et al. (2002). In the following sections methods are suggested to model<br />

thermoplastic polymers (sec.3.3.1), a model for nanoclay (sec.3.3.2) and finally a working<br />

model for thermoplastic polymer nanocomposite (sec.3.3.2).<br />

3.3.1 <strong>The</strong>rmoplastic Polymers<br />

In the MD modeling <strong>of</strong> polymer systems, coarse graining has found wide acceptance.<br />

Coarse graining significantly reduces the computational time by eliminating degrees <strong>of</strong><br />

freedom from the system that would be considered in the full atomistic model. For<br />

example, a full MD model <strong>of</strong> polypropylene will typically include the intra- and inter-<br />

51


molecular bonded and non-bonded interactions <strong>of</strong> all the constituent atoms <strong>of</strong> the atom-<br />

istic model. A coarse grained model simplifies these interactions by creating an “equiv-<br />

alent” or “pseudo” atom that has bonded and non-bonded properties equivalent to the<br />

monomer <strong>of</strong> the full atomistic system. In this manner, the computations per monomer is<br />

effectively reduced. <strong>The</strong> reduced degrees <strong>of</strong> freedom translates to computational savings<br />

for the MD model. Another advantage <strong>of</strong> coarse graining is that it overcomes the high<br />

frequency vibration <strong>of</strong> the C-H bonds in the polymer which automatically allows for larger<br />

timestep in the MD simulation. While coarse graining has obvious advantages, certain<br />

details <strong>of</strong> the chemical structure <strong>of</strong> the molecular model are inherently overlooked such<br />

as the chain rotation and its related effect <strong>of</strong> on chain configuration and relaxation. For<br />

a full realization <strong>of</strong> polymer properties these structural details may be <strong>of</strong> significance and<br />

needs to be considered for future studies. In this dissertation, the atomistic model for<br />

polypropylene uses the coarse-grained bead-spring model with parameters suggested by<br />

Wei et al. (2002). In Wei et al. (2002), the −CH2− molecule is modeled as a single atom<br />

or “repeating unit” for their coarse grained model <strong>of</strong> polyethylene. In polypropylene, the<br />

repeating unit (or monomer) is (−C3H6−)n where n is the number <strong>of</strong> repeated units in<br />

the polymer. <strong>The</strong> structure <strong>of</strong> molecule is given in fig.3.7.<br />

<strong>The</strong> coarse grained model collapses the entire structure to two backbone pseudo atoms<br />

as shown in fig.3.8. <strong>The</strong> mass <strong>of</strong> the polypropylene monomer is 42g/mol. In this coarse<br />

grained model the mass <strong>of</strong> each pseudo monomer is taken to be 21g/mol. An alternate<br />

coarse grained model could possibly be if the molecules CH2,CH3 and CH are consid-<br />

ered as separate entities, this would provide an improved description <strong>of</strong> the monomer<br />

with marginal increases in computational demand. In order to generate the amorphous<br />

polymer model we employ a random walk algorithm to place the coarse-grained beads<br />

at randomly chosen sites inside the volume box. All polymer chains were assumed to<br />

have length <strong>of</strong> 20 beads, to minimize entropic contributions to the Helmholtz free energy.<br />

52


Figure 3.7: Chemical formula <strong>of</strong> polypropylene<br />

Figure 3.8: Coarse grained polypropylene monomer<br />

After the polymer is created, the following steps are chosen to equilibrate the model at<br />

the required ambient temperature. Note that the simulations in this work are in the NVT<br />

canonical ensemble.<br />

53


1. As mentioned above (sec.3.1.1), the Langevin thermostat is used to regulate the<br />

temperature <strong>of</strong> the thermodynamic system at a desired level. At first a s<strong>of</strong>t non-<br />

bond potential is employed to remove the over-lapping centers <strong>of</strong> the polymer beads<br />

that might have occupied the same site (or in the immediate vicinity) during the<br />

random walk operation. <strong>The</strong> NVE dynamics (with no thermostatting) automati-<br />

cally acts to remove the overlap <strong>of</strong> the centers and hence reduce the energy <strong>of</strong> the<br />

system.<br />

2. At first we run the MD model <strong>of</strong> polymer at a temperature at least three (3) times<br />

higher than the ambient temperature <strong>of</strong> the model. For example, for the polymer<br />

model considered in this dissertation the system was thermostatted at 1000K using<br />

explicit rescaling <strong>of</strong> atomic velocities (see, Plimpton (1995)) in order ensure that<br />

the polymer molecules fill the simulation box and are not restricted to a single<br />

region <strong>of</strong> the simulation box. In general a ‘high’ temperature run ensures that the<br />

volume box is evenly filled by the polymer molecules. It should be noted here that<br />

if chain entanglements are created during the random walk procedure it will be<br />

hard to move chains away from each other. <strong>The</strong> self avoiding random walk (SAW)<br />

could be used to create the polymer model, in which case chain entanglements are<br />

automatically avoided.<br />

3. Next the temperature is reduced to desired level (the temperature for this work is<br />

300K) and run using NVE and velocity rescaling to get a uniform distribution <strong>of</strong><br />

temperatures (the kinetic energy <strong>of</strong> the system was monitored).<br />

4. Finally, the Langevin thermostat is employed and dynamics is used to equilibrate<br />

the model until total energy <strong>of</strong> the system is minimized. Alternatively the auto-<br />

correlation function (ACF) for the atomistic positions <strong>of</strong> the beads could be mon-<br />

itored to ensure system convergence to equilibrium. In principle, the equilibrium<br />

54


configuration <strong>of</strong> the polymer should be independent <strong>of</strong> the initial condition for the<br />

atomistic system. <strong>The</strong> decay <strong>of</strong> the auto-correlation function ensures that the fi-<br />

nal configuration <strong>of</strong> the system will be truly un-correlated to its initial state. An<br />

additional note must be made that running a purely NVT simulation can build<br />

significant but spurious hydrostatic stresses inside the MD model that may be<br />

detrimental to future simulations and could potentially lead to erroneous results.<br />

A NPT (constant number <strong>of</strong> particles N, pressure P and temperature T ) MD at<br />

zero pressure is suggested to relieve these spurious internal stresses.<br />

3.3.2 Nanoclay<br />

Figure 3.9: Pr<strong>of</strong>ile views <strong>of</strong> nanoclay in XY and XZ planes<br />

Montmorillonite (MMT) is a well-known clay mineral. Its structure essentially consists<br />

<strong>of</strong> two silica tetrahedral sheets sandwiching an edge-shared octahedral sheet <strong>of</strong> either<br />

aluminum or magnesium hydroxide, known as t-o-t sheets. Most naturally occuring<br />

MMT has intercalated sodium (Na + ) or calcium (Ca +2 ) ions in the gallery space. <strong>The</strong><br />

55


crystal system <strong>of</strong> the nanoclay is monoclinic with lattice parameters, a = 5.20˚A,b =<br />

9.20˚A,c = 10.13˚A and α = 90 ◦ ,β = 99 ◦ and γ = 90 ◦ . <strong>The</strong> nanoclay belongs to the C2/m<br />

space group. From Toth et al. (2004), we have the fractional coordinates <strong>of</strong> the nanoclay<br />

system as given in Table 3.1.<br />

Table 3.1: Position <strong>of</strong> Nanoclay atoms in space group C2/m in terms <strong>of</strong> fractional coordinates<br />

ξ, η, ζ<br />

Atoms ξ η ζ<br />

Al 0.000 0.333 0.000<br />

O1 0.481 0.500 0.320<br />

O2 0.172 0.728 0.335<br />

O3 0.348 0.691 0.110<br />

OH 0.419 0.000 0.105<br />

H 0.320 0.000 0.170<br />

Si 0.417 0.329 0.270<br />

Figure 3.9 is the structure <strong>of</strong> the nanoclay species used in this simulation. <strong>The</strong><br />

alumina-silicate structure in this work does not consider the inclusion <strong>of</strong> Na + or K +<br />

ions because a fully exfoliated nano-clay morphology is considered in this dissertation.<br />

<strong>The</strong> following steps are chosen to create and equilibrate the polymer nano-composite<br />

model at the required ambient temperature.<br />

1. <strong>The</strong> nanoclay model is embedded in a coarse grained propylene model generated<br />

using the randolm walk algorithm. As before, an energy minimization is run using<br />

the s<strong>of</strong>t potential approach. In this case however, the non-bonded potential applied<br />

to both the polymer-nanoclay interaction and the polymer-polymer interaction.<br />

<strong>The</strong> nanoclay does not participate in this NVE dynamics (i.e., the force on it is<br />

zeroed at the end <strong>of</strong> every run).<br />

2. Once the overlap energies are dissipated, NVE dynamics coupled with high temper-<br />

ature thermostatting (with explicit velocity rescaling) is used to evenly distribute<br />

the polymer beads around the simulation box. Once again the nanoclay does not<br />

56


move during polymer dynamics.<br />

3. Next the polymer molecules are held fixed while NVE is performed on the nanoclay<br />

to equilibrate its configuration. In this manner the nanoclay will equilibrate under<br />

the influence <strong>of</strong> forces from the surrounding polymer matrix in addition to its own<br />

internal forces.<br />

4. In the final step the Langevin thermostat is employed and atomistic dynamics is<br />

used to equilibrate the polymer nanocomposite (coarse grained polymer + nanoclay)<br />

model till desired energy equilibrium is achieved.<br />

3.4 Applications <strong>of</strong> the Molecular Dynamics Algorithm<br />

As shown above the molecular dynamics algorithm can be used to develop a realistic<br />

model <strong>of</strong> atomistic interactions. As a numerical scheme it provides a viable alternative<br />

to computationally expensive quantuum mechanical calculations. Some <strong>of</strong> the applica-<br />

tions <strong>of</strong> the MD algorithm has been to study free-energy (Muller-Plathe, 1991; Kollman,<br />

1993; Mavrantzas & <strong>The</strong>odorou, 1998; Zhao & Aluru, 2008), transport properties such<br />

as diffusivity (Pant & Boyd, 1993; Heffelfinger & van Swol, 1994; Charati & Stern, 1998),<br />

thermal conductivity (Maruyama, 2002; Abramson et al., 2002; Berber et al., 2000; Volz<br />

& Chen, 1999) etc., in addition to mechanical property characterization <strong>of</strong> novel mate-<br />

rials where an experimental characterization <strong>of</strong> properties is difficult (Suter et al., 2007,<br />

1995; Frankland et al., 2002; Liew et al., 2004). In this section two cases <strong>of</strong> interest are<br />

considered: (a) the MD method is used to study evolution <strong>of</strong> free energy, specifically the<br />

Helmholtz free energy <strong>of</strong> the polymer system as a function <strong>of</strong> mechanical strain in the<br />

body and (b) a MD based scheme is developed to study constitutive relations <strong>of</strong> poly-<br />

mer nano-composites using the virial stress definition coupled with continuum damage<br />

mechanics (CDM).<br />

57


3.4.1 Determination <strong>of</strong> Free-Energy due to deformation<br />

<strong>The</strong> thermodynamic work that can be extracted from a system is dependent on the free-<br />

energy in the system. In polymers, free-energy dominates the deformation process and<br />

must be quantified for a complete description <strong>of</strong> constitutive behavior. This scenario is<br />

especially true for long chain macromoecules where the configurational entropy <strong>of</strong> the<br />

system is predominant and cannot be ignored. Through the evaluation <strong>of</strong> free energy we<br />

can quantify both energetic and entropic contributions in the deformation <strong>of</strong> polymers.<br />

<strong>The</strong> MD algorithm has been used to calculate free energy <strong>of</strong> reactions, solvation, phase<br />

change and mechanical deformation.<br />

However, the determination <strong>of</strong> free-energy <strong>of</strong> a thermodynamic system through en-<br />

semble averaging methods such as MD or Monte-Carlo (MC) is difficult due to inherent<br />

short comings in these algorithms which is explained as follows: For most macromolecules<br />

there may exist multiple minimum energy configurations separated by low-energy bar-<br />

riers. In most cases <strong>of</strong> the MD (or MC) simulation the sampling space is limited to a<br />

small zone (or lower energy levels) <strong>of</strong> the phase space. On the other hand, inspection <strong>of</strong><br />

eq.(3.9) reveals that the higher energy levels makes important contributions to the Hel-<br />

moholtz energy due to the exp (H(rN, p N)/kBT ) term in the ensemble integral. Hence<br />

an adequate phase-sampling must be performed in order to accurately determine the free<br />

energy. <strong>The</strong>re are multiple methods <strong>of</strong> interest in the determination <strong>of</strong> the Helmholtz<br />

energy, such as (Leach, 1996): thermodynamic perturbation, thermodynamic integra-<br />

tion, slow-growth method etc. In this work we shall examine the use <strong>of</strong> thermodynamic<br />

integration for the purpose <strong>of</strong> free-energy <strong>of</strong> solids undergoing deformation.<br />

While the calculation <strong>of</strong> absolute free energies are difficult for materials with disor-<br />

dered phases, the change in free energy <strong>of</strong> an amorphous system can be quantified. Since,<br />

the free energy <strong>of</strong> a thermodynamic system is a state variable (i.e., it is path independent)<br />

58


the thermodynamic integration (TI) method seeks to determine the change in free energy<br />

between two energetically distinct systems that exist at locations ‘0’ and ‘1’ separated in<br />

phase space. In evaluating the change in free energy TI prescribes a phase trajectory to<br />

the system such that we can sufficiently sample all energy levels that exist between states<br />

0 and 1. <strong>The</strong> mathematical realization <strong>of</strong> this procedure is conceived by re-writing the<br />

Hamiltonian <strong>of</strong> the system as an explicit function <strong>of</strong> the variable λ. <strong>The</strong> Hamiltonian <strong>of</strong><br />

the system is simply the sum <strong>of</strong> kinetic energy K and potential energy U. <strong>The</strong> parameter<br />

λ is introduced as a perturbation in the parameters <strong>of</strong> the interaction energy functions<br />

that are in U (see section 3.2). Consider the total potential energy <strong>of</strong> the system as<br />

follows,<br />

U(r) = <br />

bonds<br />

+ <br />

i<br />

kl<br />

2 (l − l0) 2 + <br />

j>i<br />

4ɛij<br />

angles<br />

σij 12 r<br />

kθ<br />

2 (θ − θ0) 2 + <br />

−<br />

<br />

σij<br />

6<br />

r<br />

N<br />

tors n=0<br />

Vn<br />

[1 + cos(nω − γ)]<br />

2<br />

(3.34)<br />

We then prescribe the values <strong>of</strong> the constants in the energy field (e.g., kl, kθ, etc.) in<br />

the above equation (3.34), such that as λ is ramped from 0 to 1, it forces the system to<br />

evolve from a state 0 (at λ = 0) to a final state 1 (at λ = 1). That is we have,<br />

kl(λ) = k 0 l (1 − λ) + k 0 l λ (3.35)<br />

kθ(λ) = k 0 θ(1 − λ) + k 1 θλ (3.36)<br />

Vn(λ) = V 0<br />

n (1 − λ) + V 1<br />

n λ (3.37)<br />

ɛij(λ) = ɛ 0 ij(1 − λ) + ɛ 1 ijλ (3.38)<br />

σij(λ) = σ 0 ij(1 − λ) + σ 1 ijλ (3.39)<br />

Where, the superscripts 0 and 1 pertains to the constants that characterize the various<br />

interaction energies in each state. Hence, we have an explicit relation for the Hamiltonian<br />

59


(and hence the free-energy) <strong>of</strong> the system as a function <strong>of</strong> the internal variable λ. We<br />

can now determine the change in free energy <strong>of</strong> the system (∆A) at state 1 relative to a<br />

state 0 as follows,<br />

From eq.(3.8) it follows,<br />

get,<br />

∆A = A 1 − A 0<br />

λ=1<br />

∂A<br />

∂λ<br />

=<br />

λ=0<br />

1 ∂Q<br />

= −kBT<br />

Q ∂λ<br />

(3.40)<br />

∂A<br />

dλ (3.41)<br />

∂λ<br />

(3.42)<br />

Substituting for the partition function Q for the canonical ensemble from eq.(3.5) we<br />

∂A<br />

∂λ<br />

<br />

1<br />

=<br />

Q<br />

<br />

∂H<br />

=<br />

∂λ<br />

dp N N ∂H<br />

dr<br />

λ<br />

∂λ exp (−H/kBT ) (3.43)<br />

(3.44)<br />

Where, λ denotes the ensemble average for a given value <strong>of</strong> λ, and H is the<br />

Hamiltonian. We can then evaluate the change in free-energy (∆A) by eq.(3.45), which<br />

is the area under the curve in fig.(3.10).<br />

∆A =<br />

λ=1<br />

λ=0<br />

<br />

∂H<br />

dλ (3.45)<br />

∂λ λ<br />

<strong>The</strong> Helmholtz energy can hence be evaluated by finding the ensemble average in<br />

eq.(3.45) for various values <strong>of</strong> λ (between 0 and 1) and then using numerical quadrature<br />

to determine the integral. In this dissertation a representative TI method has been<br />

suggested to find the Helmholtz free energy due to deformation for the polymer model<br />

described in section 3.3.1. For the sake illustration we consider only the interaction<br />

60


Figure 3.10: Calculation <strong>of</strong> free energy change using thermodynamic integration<br />

energies due to changes in bond-length and non-bonded interactions. <strong>The</strong> free energy<br />

is determined by parameterzing the non-bonded interactions in the internal energy as a<br />

function <strong>of</strong> λ. That is we write,<br />

Enb = <br />

<br />

<br />

σ(λ)<br />

4ɛ(λ)<br />

r<br />

i<br />

j>i<br />

12<br />

<br />

6<br />

σ(λ)<br />

−<br />

r<br />

(3.46)<br />

Where, ɛ(λ) = ɛ 0 (1 − λ) + ɛ 1 λ and σ(λ) = σ 0 (1 − λ) + σ 1 λ. We can now write the<br />

Hamiltonian for the system as,<br />

H(p N , r N ; λ) =<br />

N<br />

i=1<br />

p i · p i<br />

Where, p i is the momentum <strong>of</strong> atom ‘i’ and U(λ) is given as,<br />

bonds<br />

i<br />

j>i<br />

mi<br />

U(λ) = kl<br />

2 (l − l0) 2 + <br />

<br />

<br />

σ(λ)<br />

4ɛ(λ)<br />

r<br />

+ U(λ) (3.47)<br />

12<br />

<br />

6<br />

σ(λ)<br />

−<br />

r<br />

(3.48)<br />

Thus we can derive a closed form analytical expression for the derivative <strong>of</strong> the Hamil-<br />

61


tonian with respect to λ,<br />

∂H<br />

∂λ<br />

<br />

=<br />

i<br />

j>i<br />

4(ɛ 1 − ɛ 0 )<br />

+6(σ 1 − σ 0 ) ɛ<br />

σ<br />

<br />

2<br />

σ 12 <br />

σ<br />

<br />

6<br />

−<br />

r r<br />

<br />

σ<br />

12 <br />

σ<br />

<br />

6<br />

−<br />

r r<br />

(3.49)<br />

During MD simulation, the time average <strong>of</strong> the quantity ∂H/∂λ is calculated for each<br />

value <strong>of</strong> λ. A three-point Gaussian quadrature (Reddy, 2006) is used to evaluate the<br />

integral, hence 3 separate MD simulations has to be run for each step <strong>of</strong> deformation<br />

state. <strong>The</strong> results from this analysis is given in Chapter 6.<br />

3.4.2 Multi-scale modeling <strong>of</strong> <strong>The</strong>rmoplastic composites<br />

Figure 3.11: Schematic <strong>of</strong> the MD simulation model for polymer nanocomposite<br />

One <strong>of</strong> the primary challenges in modeling novel materials is the characterization<br />

<strong>of</strong> mechanical properties such as stiffness and strength. While the widely accepted en-<br />

gineering solution is to a subscribe a phenomenological model, MD presents a case <strong>of</strong><br />

representing stress-strain response based on actual atomistic behavior. In this work we<br />

62


consider the stress-strain (constitutive) response <strong>of</strong> polymer nanocomposite employing<br />

the atomistic stress definition (Zhou, 2003). It is envisoned that a hierarchical atomistic<br />

simulation based model will be computationally effective in characterizing the stiffness<br />

and damage induced degradation <strong>of</strong> material stiffness. To our knowledge this approach<br />

has not been attempted.<br />

<strong>The</strong> procedure involves deforming the simulation box as per a prescribed deformation<br />

gradient F. Note that although a homogeneous deformation field is assumed, local in-<br />

homogeneities such as micro-void formation, void coalescence, interface debond (in the<br />

nanocomposite) etc. are allowed to evolve in the MD system. In this dissertation we are<br />

particularly interested in modeling a nanoclay platelet oriented along the local x1-axis <strong>of</strong><br />

the simulation box as shown in fig.(3.11).<br />

Figure 3.12: Schematic <strong>of</strong> deformation <strong>of</strong> simulation box under a volumetric and deviatoric<br />

(e.g. simple shear) deformation gradients<br />

<strong>The</strong> specified orientation <strong>of</strong> nanocomposite gives the nanoscale RVE orthotropic prop-<br />

erties. This assumption enables us to decompose deformation into its deviatoric and vol-<br />

umetric parts and study each mode <strong>of</strong> deformation in isolation (as shown in fig.3.12). In<br />

this research the stress-strain response characteristics <strong>of</strong> the system involves deforming<br />

63


the volume box (in volumetric or deviatoric deformation) and determining the stress that<br />

develops inside the specimen. Historically, researchers employing atomistic simulation<br />

algorithms have used the virial stress to define stress in the discrete system,<br />

¯Παβ = 1<br />

Ω<br />

<br />

<br />

i<br />

−miv α i v β<br />

i<br />

+ 1<br />

2<br />

<br />

j>i<br />

r α ijf β<br />

ij<br />

<br />

(3.50)<br />

Where, α, β are indicial variables that could be either =1, 2 or 3, mi is mass <strong>of</strong> atom<br />

‘i’, Ω is the volume <strong>of</strong> the atomistic system, r α ij is α component <strong>of</strong> the vector joining<br />

atoms i and j, while f α ij is the interaction force between atoms i and j as derived from the<br />

total potential energy <strong>of</strong> system (eq.3.27). In Zhou (2003), the definition <strong>of</strong> true-stress<br />

(or Cauchy stress) as a measure <strong>of</strong> internal force between the particles (or atoms) <strong>of</strong> the<br />

system is invoked to show that the commonly accepted format <strong>of</strong> virial stress (eq.3.50) is<br />

incorrect. Further, it is shown that the virial stress in its definition in eq.(3.50) is only a<br />

measure <strong>of</strong> momentum flux through an imaginary plane fixed in space. <strong>The</strong> Cauchy stress<br />

is now defined as the ensemble average (or time-averaged through MD) <strong>of</strong> the following<br />

quantity,<br />

¯σαβ = 1<br />

2Ω<br />

<br />

i<br />

j>i<br />

r α ijf β<br />

ij<br />

(3.51)<br />

<strong>The</strong> stress measure converges to the Cauchy tensor in the continuum sense, when a<br />

sufficiently large time and volume average is performed. Later in Chapter 5 we develop<br />

a phenomenological model for the failure <strong>of</strong> nano-composites using the applied strain<br />

and measured stress (eq.3.51), and compare the results with the baseline case where the<br />

polymer does not contain nano-particle reinforcement.<br />

64


Chapter 4<br />

Concurrent Multiscale Modeling <strong>of</strong> <strong>The</strong>rmoplastic PNC<br />

Structural and configurational characteristics <strong>of</strong> a polymer are both time and tem-<br />

perature dependent. In particular, polymer characteristics such as viscoelastic relaxation<br />

is a manifestation <strong>of</strong> polymer chain rearrangements and cooperative motion <strong>of</strong> polymer<br />

molecules as a function <strong>of</strong> time. <strong>The</strong> recent interest in nanoclay technology to enhance<br />

polymer properties has demanded further scientific enquiry to better explain and un-<br />

derstand the physics <strong>of</strong> reinforcement. In this work the focus is on modeling material<br />

behavior using multi-physics based algorithms to capture both the nano-scale (discrete)<br />

behavior and the continuum behavior at the macroscopic length scales.<br />

<strong>The</strong> theoretical model <strong>of</strong> thermoplastic polymer nano-composites must capture the<br />

molecular physics at the atomistic scale in order to provide a better description <strong>of</strong> ma-<br />

terial behavior. It is envisioned that molecular physics based modeling approaches will<br />

eventually help in better design and risk assessments <strong>of</strong> polymer based materials for<br />

structural applications. <strong>The</strong> purpose <strong>of</strong> this chapter is to provide a physical rather than<br />

phenomenological model <strong>of</strong> the molecular rearrangements and configurational changes in<br />

polymer in relation to macroscale loading through a concurrent atomistic and continuum<br />

coupled simulation.<br />

4.1 Introduction<br />

Over the past 10 years numerous analysis techniques has been proposed to tie nanoscale<br />

events to macroscale stimuli during simulation time. <strong>The</strong>se simulations fall into the<br />

category <strong>of</strong> concurrent analysis. Typically in concurrent simulations, the continuum<br />

response is simulated using the finite element method and the region <strong>of</strong> interest or process<br />

65


zone (see, fig.4.1) is simulated using the MD algorithm. <strong>The</strong> information exchange occurs<br />

through the handshake zone where a region <strong>of</strong> pad atoms occupy the same space as the<br />

continuum finite elements.<br />

Figure 4.1: Schematic <strong>of</strong> Concurrent coupling<br />

Depending on the description <strong>of</strong> the overlap region, the concurrent coupling algo-<br />

rithm can be broadly divided in to two kinds: (a) direct coupling (e.g. Wagner & Liu<br />

(2003); Xiao & Belytschko (2004) etc.) and the recently proposed (b) statistical coupling<br />

technique (Saether et al., 2009).<br />

In direct coupling (DC), the finite element mesh is gradually refined from a coarse<br />

mesh starting from the external boundaries <strong>of</strong> the model down to atomistic sizes at the<br />

handshake region. At the interface zone the atoms from the atomistic domain is coupled<br />

one-on-one to each interface finite element node, as the arrows in figure 4.2 indicate.<br />

While DC based methods is a widely employed for coupling atomistic simulations to<br />

continuum, there are some inherent disadvantages:<br />

1. Tying individual atoms to finite element nodes at the interface means cumbersome<br />

mesh algorithms to realize the model setup.<br />

66


Figure 4.2: Schematic <strong>of</strong> the Direct Coupling methodology<br />

2. Time-step for dynamic problems are severely effected due to refined mesh at the<br />

interface.<br />

3. For finite temperature simulations <strong>of</strong> the MD domain, the atoms are under constant<br />

thermal motion, for example in solid lattices the atoms undergo thermal vibration<br />

about its equilibrium position. <strong>The</strong>se vibrations emanate out <strong>of</strong> the MD domain<br />

into the continuum region in the form <strong>of</strong> high frequency noise, and can potentially<br />

destroy the numerical solution.<br />

An alternative to direct coupling, is the recently developed statistical coupling tech-<br />

nique (Saether et al., 2009). In this work, the statistical coupling method also called the<br />

embedded statistical coupling method (ESCM) is employed to study polymer systems.<br />

Some <strong>of</strong> the salient features <strong>of</strong> the coupling technique are listed here as follows:<br />

1. <strong>The</strong> handshake zone is divided into volume cells through Voronoi tessellations (as<br />

shown in fig.4.3) such that the interface node ‘communicates’ information with<br />

the atomistic domain with a group <strong>of</strong> atoms rather than individual atoms. This<br />

construction has a two-fold advantage (1) the finite element mesh does not have to<br />

67


e refined down to atomistic size, and (2) the volume cell construction allows for<br />

proper statistical sampling <strong>of</strong> atomistic trajectories and can hence provide a better<br />

description <strong>of</strong> the material behavior<br />

2. Through the statistical coupling methodology the acoustic vibrations (phonons)<br />

emitted from the thermal environment <strong>of</strong> the atomistic zone can be captured and<br />

diffused at the handshake region hence allowing for larger time-steps in the contin-<br />

uum update (see, Saether et al. (2009)).<br />

Figure 4.3: Schematic <strong>of</strong> the Statistical Coupling methodology (from Saether et al. (2009))<br />

<strong>The</strong> primary advantage <strong>of</strong> using concurrent coupling is the ability to avoid finite<br />

boundary (or free edge) effect. Concurrent simulation strategies also imply computa-<br />

tional advantages by cutting down on the computational overhead involved in atomistic<br />

simulations by restricting the calculations to a narrow process zone. In the following<br />

section ESCM is extended to couple GIMP (see, Chapter 2) to MD simulations.<br />

68


4.2 Embedded Statistical Coupling in GIMP<br />

Figure 4.4: Layout <strong>of</strong> Coupled Model<br />

In this section the statistical scheme developed in (Saether et al., 2009) is extended to<br />

couple the GIMP algorithm to molecular dynamics (MD). GIMP is a meshless particle<br />

scheme, originally developed by (Bardenhagen & Kober, 2004) as an extension to the<br />

material point method (see e.g., Sulsky et al. (1995)) to improve the numerical stability<br />

<strong>of</strong> the method. As reviewed in Chapter 2, GIMP particles are constructed as finite<br />

domain entities that deform in a manner consistent with the deformation <strong>of</strong> the material<br />

body. In ESCM,volume cells (VC) are constructed at the overlap domain <strong>of</strong> MD/GIMP<br />

(also called handshake zone, see fig.4.4) to communicate information between the internal<br />

69


MD environment and the external continuum (GIMP) region. <strong>The</strong> volume cells serves<br />

the three-fold purpose <strong>of</strong> (a) determining averaged displacements <strong>of</strong> the atomic entities<br />

within it (b) transferring forces from the continuum region down to the atoms on the MD<br />

boundary (c) preventing the MD surface from degradation when forces are applied on it<br />

(d) prevent the artificial “heating” <strong>of</strong> the MD atoms due to the loading wave ingressing<br />

from the continuum environment or reflection <strong>of</strong> phonons emanating inside the atomistic<br />

region at the MD/GIMP interface. In GIMP, it is easy to setup the VCs since the<br />

deformation <strong>of</strong> the particle domain is tracked during real-time simulation. Hence, in a<br />

continuous deformation field the contiguous particle domains remain contiguous even after<br />

deformation. <strong>The</strong> construction <strong>of</strong> volume cells also enables easier spatial discretization<br />

<strong>of</strong> the atomistic/continuum overlap domain.<br />

In ESCM, the MD domain is divided into three (3) sub domains as shown in the<br />

fig.(4.4). <strong>The</strong> surface volume cell (SVC), internal volume cell (IVC) and the interior MD<br />

region. Since the material point (mp) domain serve as the volume cells, the IVC is also<br />

referred to as IVC mp, similarly SVC is also called SVC mp. <strong>The</strong> IVC calculates time-<br />

averaged displacements <strong>of</strong> the atoms (referred to here on as ∆u IVC<br />

p<br />

) within each IVC and<br />

applies this as incremental displacement <strong>of</strong> the material point <strong>of</strong> each IVC at time ‘t’,<br />

Where, x t−∆t<br />

p<br />

x t p = x t−∆t<br />

p<br />

+ ∆u IVC<br />

p<br />

(4.1)<br />

is the position <strong>of</strong> the material point from the previous timestep. <strong>The</strong><br />

material point IVC uses the time-averaged displacements and calculates the internal force<br />

on the background nodes <strong>of</strong> GIMP due to this change in position. This force is applied<br />

as an equally distributed external force on the atoms within the IVC. <strong>The</strong> purpose <strong>of</strong> the<br />

SVC is to prevent the degradation <strong>of</strong> the MD surface atoms due to the internal forces<br />

from the continuum.<br />

<strong>The</strong> continuum is updated using the explicit Verlet integration algorithm introduced<br />

70


in Chapter 2. In the interest <strong>of</strong> brevity, the equations are not repeated here. Given<br />

below is the explicit time-integration <strong>of</strong> the atomistic equations <strong>of</strong> motion that follows<br />

the continuum update.<br />

In the atomistic update, the IVC atoms that are a part <strong>of</strong> the IVC material points is<br />

under the influence <strong>of</strong> an external force referred to as, f IVC<br />

a , where the subscript a refers<br />

to atom within the IVC ‘p’. This external force is calculated as,<br />

f IVC<br />

a<br />

=<br />

tot,t+∆t<br />

i<br />

Sipfi N IVC<br />

a<br />

(4.2)<br />

Where, Sip is the GIMP interpolation function between IVC material point ‘p’ and<br />

node ‘i’ and N IVC<br />

a<br />

is the number <strong>of</strong> atoms in this IVC. As can be seen the interpolation<br />

function for each atom in the IVC is not calculated, instead all atoms <strong>of</strong> an IVC are<br />

assumed to have the same interpolation function. Further, we determine the velocities<br />

and accelerations <strong>of</strong> the SVC atoms as follows,<br />

v ∗,SVC<br />

a = <br />

a SVC<br />

a = <br />

i<br />

i<br />

Sipv ∗ i<br />

Sip<br />

<br />

f int,t<br />

i<br />

<br />

ext,t+∆t<br />

+ f<br />

i<br />

m t i<br />

(4.3)<br />

(4.4)<br />

<strong>The</strong> following algorithm details the update <strong>of</strong> the atoms within the interior (abbrevi-<br />

ated as INT), IVC and SVC parts <strong>of</strong> the MD domain using a Verlet integration scheme.<br />

In the description to follow the interaction energy between atoms in the MD zone is<br />

designated as the scalar function ‘φ’. For example, in a simple Leonard-Jones system,<br />

<br />

σ 12 <br />

σ 6<br />

φ = 4ɛ − where, r=|rij| is the distance between atoms i and j. Also defined<br />

r r<br />

is the discrete time step <strong>of</strong> molecular dynamics algorithm as, ∆τ = ∆t<br />

N<br />

where, ∆t is<br />

the coarse time-step and N is the number <strong>of</strong> MD steps. Begin time integration <strong>of</strong> MD<br />

71


equations <strong>of</strong> motion.<br />

1. Update velocities <strong>of</strong> atoms in the interior (INT) and IVC atoms <strong>of</strong> the MD zone<br />

at the beginning <strong>of</strong> time-step τ <strong>of</strong> MD. As mentioned in Chapter 3, the Langevin<br />

thermostat (see e.g., Allen & Tildesley (1987)) is used to maintain the temperature<br />

<strong>of</strong> the atomic zones. As per its formulation a stochastic damping force f T a is ap-<br />

plied on atom ‘a’ such that the overall temperature <strong>of</strong> the system is maintained at<br />

‘T’ (temperature units). In general, the Langevin equations are used to simulate<br />

brownian dynamics in a thermally excited system. For an atom ‘a’ the mid-time<br />

step velocities (denoted as v ∗ ) in the INT and IVC zones are,<br />

Where, ∇x τ p<br />

v ∗,INT<br />

a = v τ,INT<br />

a + ∆τ<br />

<br />

−∇x<br />

2<br />

τ p<br />

v ∗,IVC<br />

a = v τ,IVC<br />

a + ∆τ<br />

2<br />

−∇x τ p<br />

<br />

T=Γ φ + fa mINT a<br />

T=0 φ + fa mIVC a<br />

+ f IVC<br />

a<br />

(4.5)<br />

(4.6)<br />

is the spatial derivatives along the three orthogonal spatial directions.<br />

This indicates that the spatial derivatives are with respect to the current coordinates<br />

<strong>of</strong> the atom. Please note that ‘Γ’ is the temperature <strong>of</strong> the atomic zone.<br />

2. Update the position <strong>of</strong> all atoms in all zones <strong>of</strong> the MD region,<br />

x τ+∆τ<br />

a<br />

= x τ a + ∆τv ∗,I<br />

I ∈ (IVC,SVC,INT) (4.7)<br />

3. Update velocities atoms in the interior (INT), IVC and SVC atoms <strong>of</strong> the MD zone<br />

at the beginning <strong>of</strong> time-step ‘∆τ’ <strong>of</strong> the MD. For an atom ‘a’ the updated velocities<br />

72


are,<br />

Where, ∇ x τ+∆τ<br />

p<br />

v τ+∆τ,INT<br />

a = v ∗,INT<br />

a + ∆τ<br />

2<br />

v τ+∆τ,IVC<br />

a = v ∗,IVC<br />

a + ∆τ<br />

2<br />

v τ+∆τ,SVC<br />

a = v ∗,SVC<br />

a + ∆τ<br />

2 aSVC a<br />

<br />

<br />

−∇ x τ+∆τ<br />

p<br />

−∇ x τ+∆τ<br />

p<br />

m INT<br />

a<br />

φ + f T=Γ<br />

<br />

a<br />

φ + f T=0<br />

a<br />

m IVC<br />

a<br />

+ f IVC<br />

<br />

a<br />

(4.8)<br />

(4.9)<br />

(4.10)<br />

is the spatial derivative along the three orthogonal space direc-<br />

tions. This indicates that the spatial derivatives are with respect to the updated<br />

coordinates <strong>of</strong> the atom. Note that at the end <strong>of</strong> every MD update the following<br />

displacement average is recorded,<br />

∆u IVC<br />

p<br />

= 1<br />

t+N∆τ <br />

N<br />

τ=t<br />

N IVC<br />

a<br />

a=1<br />

<br />

τ,IVC xa − xt,IVC <br />

a<br />

N IVC<br />

a<br />

(4.11)<br />

At the end <strong>of</strong> the MD update the control switches back to the GIMP algorithm and<br />

the continuum update follows using the IVC displacements calculated during simulation<br />

time (eq.4.1). <strong>The</strong> proposed algorithm was implemented to study fracture initiation and<br />

evolution in the thermoplastic polymer system (introduced in Chapter 3) in relation to<br />

macro-scale loading conditions. <strong>The</strong> results from the simulations are presented in Chapter<br />

6.<br />

4.3 Setup <strong>of</strong> Coupled Simulation model<br />

<strong>The</strong> algorithmic implementation <strong>of</strong> the scheme described above is realized using Visual<br />

C++ programming. However, the need to scale to larger systems combined with the fact<br />

that both the continuum and atomistic methods are particle based methods necessitates<br />

the use <strong>of</strong> parallel implementation with distributed computing. This dissertation focusses<br />

73


on the effectiveness <strong>of</strong> the new method to model polymers and leaves room for further<br />

improvements including computational efficiency. In this section, the details <strong>of</strong> setting<br />

up a coupled simulation in a continuum/atomistic solution paradigm is explained. Figure<br />

4.5 is a schematic <strong>of</strong> the atomistic model setup.<br />

Figure 4.5: Setup <strong>of</strong> atomistic model for Concurrent coupling<br />

74


Figure 4.6: Setup <strong>of</strong> ESCM/GIMP model for Concurrent coupling<br />

Given below are the steps followed in the typical setup <strong>of</strong> the coupled thermoplastic<br />

polymer model.<br />

1. <strong>The</strong> thermoplastic polymer amorphous structure is generated using the random<br />

walk operation.<br />

2. A s<strong>of</strong>t non-bond potential is employed to remove the over-lapping centers <strong>of</strong> the<br />

polymer beads that might have occupied the same site (or in the immediate vicinity)<br />

during the random walk operation. <strong>The</strong> NVE dynamics (with no thermostatting)<br />

automatically acts to remove the overlap <strong>of</strong> the centers and hence reduce the energy<br />

<strong>of</strong> the system.<br />

3. <strong>The</strong> energy minimized MD model is then embedded in a volume box surrounded<br />

by a regular ordered lattice. <strong>The</strong> atoms <strong>of</strong> the ordered lattice will eventually couple<br />

to the continuum material points through the IVC and SVC.<br />

75


4. <strong>The</strong> system is equilibrated at high temperature using eplicit velocity rescaling in<br />

order to evenly distribute the polymer molecules inside the volume box<br />

5. Finally, the system temperature is reduced and MD is performed until the desired<br />

energy equilibration is achieved.<br />

After the atomistic coupled setup domain is complete, the particles cells are embedded<br />

around the atomistic domain and the IVC and SVC atom lists are constructed using a<br />

linear search algorithm depending on the positions <strong>of</strong> the MD atoms relative to the<br />

particle domain (see, fig.4.6).<br />

76


Chapter 5<br />

Hierarchical Modeling <strong>of</strong> Damage in PNCs<br />

While the weight savings and property enhancements that can be realized using nano<br />

composites are obvious, the availability <strong>of</strong> physical models to capture these enhancements<br />

in relation to macro-structural load history is limited. This is due to the complexities<br />

involved in the modeling <strong>of</strong> polymeric materials and the obvious length and time scales<br />

involved in the structural material. On the other hand the significant enhancements in<br />

the nanoclay nanocomposites are attributed to the richness <strong>of</strong> polymer-nanoclay interac-<br />

tions at the interface. <strong>The</strong>se interactions occur on a length scale which is comparable to<br />

the interatomic separation. Further, the heterogeneity <strong>of</strong> the interacting species at the<br />

nanoscale adds to the level <strong>of</strong> complexity. In this chapter, a novel RVE based model <strong>of</strong><br />

the nanoclay/polymer interface is proposed. <strong>The</strong> model attempts to capture the stiffness<br />

enhancement due to nanoscale reinforcement through tracking the change in energy over<br />

baseline as measured through the framework <strong>of</strong> volume-averaged atomistic stresses. Fur-<br />

ther, the model attempts to capture the stiffness degradation using an internal state vari-<br />

able (ISV) approach. <strong>The</strong> proposed analysis correlates the dissipation <strong>of</strong> internal energy<br />

<strong>of</strong> the system to deterioration <strong>of</strong> interface and bulk properties <strong>of</strong> the nanoclay polymer<br />

composite. <strong>The</strong> model employs a smeared analysis approach by assuming orthotropic<br />

symmetry to material properties on the local representative volume element (RVE) coor-<br />

dinate frame (we assume that the nanoclay is oriented along the local x1 axis). Figure 5.1<br />

is a schematic to explain the development <strong>of</strong> the damage model. As shown in the figure,<br />

the model constructs an equivalent model with a smeared representation <strong>of</strong> the nanoclay,<br />

polymer system. As the “smeared” model deforms a critical value <strong>of</strong> strain occurs at<br />

which damage appears in the nano-scale RVE (see, fig.5.1(c))(the damage initiation and<br />

evolution model is discussed in detail later). In this research, the damage in the equivalent<br />

77


model actually represents the debond at the polymer/nanoclay interface and micro-void<br />

coalescence in the bulk polymer (regions <strong>of</strong> polymer away from the nanoclay/polymer<br />

interface). Later, through atomistic simulation we will show that the loss in stiffness and<br />

material decohesion in the form <strong>of</strong> interface debond and micro-void coalescence are well<br />

correlated, and hence can be considered to be the primary modes <strong>of</strong> damage initiation<br />

and evolution. It is envisioned that the proposed method will help in the constitutive<br />

modeling <strong>of</strong> nano-composites by incorporating several elements <strong>of</strong> nano-scale physics into<br />

a simple phenomenological model.<br />

Figure 5.1: Development <strong>of</strong> the equivalent damage model (a) the nanoscale polymer<br />

nanocomposite (b) the nanoscale RVE with equivalent elastic orthotropic properties (c)<br />

interface debond and void coalescence in the actual polymer nanocomposite (d) damage<br />

development in the smeared model<br />

As mentioned above, we have the continuum treatment <strong>of</strong> damage evolution. In this<br />

effort, we review the thermodynamics <strong>of</strong> a deformable solid, followed by specializing the<br />

Clausius-Duhem equation to develop the constitutive relations with embedded dissipation<br />

78


due to damage evolution. In the following sections a nanoscale damage model (NDM) is<br />

introduced and internal state variable to define damage initiation and damage evolution<br />

is defined.<br />

5.1 Damage Mechanics using the Internal State Variable Ap-<br />

proach<br />

In this work the internal state variable (ISV) approach proposed by Talreja (1990) to<br />

model damage in composites will be modified and extended to model damage at the<br />

nano-scale. <strong>The</strong> proposed RVE is assumed to have a linear elastic response until failure<br />

initiation after which the damage evolution is described by the model in section 5.2.1. We<br />

begin with the review <strong>of</strong> the preliminary definitions and assumptions <strong>of</strong> the ISV model.<br />

5.1.1 Definition <strong>of</strong> Damage Tensor<br />

In Talreja (1990), the damage behavior is entirely described by a tensorial damage entity.<br />

In general, the damage entity is any micro-structural change in the material brought<br />

about by an internal dissipative mechanism. For example, in the macro-scale composite<br />

the damage entity may refer to the cracks in the matrix, fiber-matrix debond, inter-ply<br />

delamination, etc. Realistically, it is possible that multiple damage entities may coexist in<br />

side the material specimen and evolve at mutually independent rates. Hence, the damage<br />

entity is a concept to unify the multiple damage multiple modes as a single quantity, the<br />

combined effect <strong>of</strong> which initiates and describes the progression <strong>of</strong> material degradation<br />

in relation to macro-scale loading history, within the framework <strong>of</strong> thermodynamics.<br />

Let V be a finite volume surrounding a point P within the bulk material with dis-<br />

tributed damage at multiple damage modes (fig.5.2). Consider a damage entity inside<br />

the volume V. Let S be its surface area and n be the unit outward normal to the surface.<br />

79


Figure 5.2: Characterization <strong>of</strong> damage entities<br />

An influence vector function a is then defined on the surface. <strong>The</strong> vector a essentially<br />

represents a localized disturbance due to the damage entity in an otherwise uniform field.<br />

A definition for a with respect to nano-scale damage is provided later in this chapter. A<br />

second order tensor dij is defined as follows,<br />

<br />

dij =<br />

S<br />

ainjdS (5.1)<br />

Where, ai and nj are the components <strong>of</strong> the vectors a and n in three dimensional<br />

Cartesian coordinates. In the general formulation we assume that there are ‘N’ distinct<br />

damage modes in the material. We can define a damage tensor for each damage mode α<br />

(where, α = 1, 2, . . . , N) as follows,<br />

D (α)<br />

ij<br />

= 1<br />

V<br />

80<br />

<br />

kα<br />

(dij)kα<br />

(5.2)


Where, kα is the number <strong>of</strong> damage entities in each mode. If m is the tangential<br />

unit vector defined on surface S such that nimi = 0, nini = 1 and mimi = 1 (indicial<br />

summation on i implied) we can then write<br />

and<br />

ai = a(ni + mi) (5.3)<br />

Substituting eq.(5.3) into eq.(5.2) and dropping α for convenience we get,<br />

Where,<br />

Dij = D 1 ij + D 2 ij (5.4)<br />

= 1<br />

V<br />

d 1 ij =<br />

d 2 ij =<br />

<br />

k<br />

<br />

<br />

S<br />

S<br />

1<br />

(dij)k + (d 2 <br />

ij)k<br />

aninjdS (5.5)<br />

aminjdS (5.6)<br />

In this research we retain only the normal part <strong>of</strong> a in order to retain symmetry <strong>of</strong><br />

the damage tensor. Further, we assume only two distinct modes <strong>of</strong> damage hence,<br />

Dij = 1<br />

V<br />

2<br />

<br />

k=1<br />

S<br />

<br />

aninjdS<br />

k<br />

5.1.2 <strong>The</strong>rmodynamics <strong>of</strong> Damage and Constitutive equations<br />

(5.7)<br />

To derive the thermo-mechanical constitutive equations <strong>of</strong> damage we assume Truesdell’s<br />

principle <strong>of</strong> equipresence (Truesdell & Noll, 1992). <strong>The</strong> principle simply states that an<br />

independent variable assumed to present in one constitutive equation <strong>of</strong> the material<br />

must exist in all other constitutive relations as well, unless its presence contradicts the<br />

81


fundamental principles (see e.g. Malvern (1969), sec.6.7, page 379) for any constitutive<br />

relation. In Talreja (1990) the independent variables chosen are, the small strain tensor<br />

ɛ, temperature T , temperature gradient g = ∇T and damage tensor D (we have dropped<br />

the superscript α for convenience) . Hence we can define,<br />

σ = σ(ɛ, T, g, D) (5.8)<br />

ψ = ψ(ɛ, T, g, D) (5.9)<br />

η = η(ɛ, T, g, D) (5.10)<br />

q = q(ɛ, T, g, D) (5.11)<br />

˙D = ˙ D(ɛ, T, g, D) (5.12)<br />

Where, σ is the Cauchy stress tensor, ψ is the specific Helmholtz free energy, η is<br />

the specific entropy, q is the heat flux vector and ˙ D rate <strong>of</strong> damage accumulation. <strong>The</strong><br />

choice <strong>of</strong> stress, energy is assumed to satisfy the governing equations for linear momentum,<br />

angular momentum and energy balance (Talreja, 1990). <strong>The</strong> Clausius-Duhem in-equality<br />

(Malvern, 1969) is an alternate expression <strong>of</strong> the second law <strong>of</strong> thermo-dynamics. Given<br />

by,<br />

σ : ˙ɛ − ρ ˙ ψ − ρη ˙<br />

q · g<br />

T −<br />

T<br />

≥ 0 (5.13)<br />

For a finite-deformation damage model formulation we can re-write the above equation<br />

in the material frame as follows,<br />

S : ˙ E − ρ0 ˙ Ψ − ρ0N ˙ Γ −<br />

Q · G<br />

Γ<br />

≥ 0 (5.14)<br />

Where, S is the second Piola-Kirch<strong>of</strong>f stress, E is the Green-Lagrange strain tensor, ρ0<br />

is density <strong>of</strong> the reference configuration. If F is the deformation gradient with X and x as<br />

the material and spatial coordinates respectively, then Q(X, t) = JF −1 q(x, t), N(X, t) =<br />

82


η(x, t), Γ(X, t) = T (x, t), G = ∇Γ. We also have the thermodynamic relation:ψ = e−T η<br />

(or, Ψ = U − ΓN). However for the material frame formulation eq.(5.8) to eq.(5.12) have<br />

to be re-written in terms <strong>of</strong> independent variables in the material frame.<br />

Assuming small strains and differentiating eq.(5.9) with respect to time gives,<br />

˙ψ = ∂ψ<br />

∂ɛ<br />

∂ψ<br />

: ˙ɛ +<br />

∂T ˙ T + ∂ψ ∂ψ<br />

· ˙g +<br />

∂g ∂D : D ˙ (5.15)<br />

Substituting eq.(5.15) in to eq.(5.13) we get the internal dissipation inequality,<br />

<br />

σ − ρ ∂ψ<br />

<br />

: ˙ɛ − ρ η +<br />

∂ɛ<br />

∂ψ<br />

<br />

T˙ − ρ<br />

∂T<br />

∂ψ ∂ψ<br />

· ˙g − ρ<br />

∂g ∂D : D˙ q · g<br />

−<br />

T<br />

≥ 0 (5.16)<br />

Since eq.(5.16) must hold for any arbitrary rates ˙ɛ, ˙<br />

T and ˙g, we get the relations,<br />

σ = ρ ∂ψ<br />

∂ɛ<br />

η = − ∂ψ<br />

∂ψ<br />

∂g<br />

∂T<br />

(5.17)<br />

(5.18)<br />

= 0 (5.19)<br />

Combining equations (5.17) through (5.19) and (5.8) through (5.11) gives,<br />

<strong>The</strong> dissipative inequality 5.16 reduces to,<br />

σ = σ(ɛ, T, D) (5.20)<br />

ψ = ψ(ɛ, T, D) (5.21)<br />

η = η(ɛ, T, D) (5.22)<br />

− ρ ∂ψ<br />

∂D : D˙ q · g<br />

−<br />

T<br />

83<br />

≥ 0 (5.23)


Further assuming that the formulation is isothermal (i.e. g = ∇T = 0) and for purely<br />

mechanical response we arrive at,<br />

With the internal dissipation inequality,<br />

σ = σ(ɛ, D) (5.24)<br />

ψ = ψ(ɛ, D) (5.25)<br />

˙D = ˙ D(ɛ, D) (5.26)<br />

∂ψ<br />

∂D : ˙ D ≤ 0 (5.27)<br />

Hence, for a purely mechanical response, formulation <strong>of</strong> the damage tensor D and free<br />

energy ψ defines the constitutive behavior <strong>of</strong> the deformable body. <strong>The</strong> rate equation for<br />

stress, σ is now from eq.(5.25) and eq.(5.17),<br />

σij ˙ = ∂2ψ ∂ɛij∂ɛkl<br />

ɛkl ˙ + ∂2ψ Dkl<br />

∂ɛij∂Dkl<br />

˙<br />

(5.28)<br />

<strong>The</strong> implications <strong>of</strong> the above relation in the context <strong>of</strong> hyperelastic materials is that<br />

the constitutive relation remains polyconvex (Schroder & Neff, 2003) until initiation <strong>of</strong><br />

damage. <strong>The</strong> onset <strong>of</strong> damage and damage accumulation (as quantified by the rate<br />

equation above), the material deviates from polyconvex constitutive behavior due to<br />

damage induced s<strong>of</strong>tening.<br />

5.2 A Nano-scale Damage Model<br />

<strong>The</strong> internal state variable introduced above is now specialized for the case <strong>of</strong> damage in<br />

a nano-scale representative volume element (RVE). In this formulation the nanoclay is<br />

assumed to be oriented along the direction <strong>of</strong> the local 1 axis <strong>of</strong> the RVE (as shown in fig.<br />

84


Figure 5.3: Representative volume element (RVE) and nano-clay orientation in damage<br />

model<br />

5.1), hence even though the orientation <strong>of</strong> the nanoclay is accounted for through material<br />

properties, exact description <strong>of</strong> interfaces and features <strong>of</strong> the inclusion are avoided. Some<br />

<strong>of</strong> the key features <strong>of</strong> the damage model and primary assumptions are as follows,<br />

1. <strong>The</strong> material exhibits orthotropic behavior and preserves orthotropic symmetry<br />

even after damage initiation (hence, Dij = 0 if i = j).<br />

2. <strong>The</strong> damage is small hence Dij ≪ 1 and higher-order terms <strong>of</strong> damage tensor<br />

components can be ignored.<br />

3. Primary assumption We define nanoclay damage as that which occurs due to inter-<br />

face debond growth (mode <strong>of</strong> damage, α = 1) along the local 2-axis <strong>of</strong> the material<br />

(see, fig.5.4) and the micro-void coalescence in the x1 and x3 directions (α = 2,<br />

respectively) in the bulk polymer.<br />

<strong>The</strong> primary assumption can be explained further as follows: We assume that the<br />

interface debond occurs normal to the local 1-3 plane <strong>of</strong> the nanoclay/polymer interface<br />

and the D (1)<br />

22 (or D2) term <strong>of</strong> the damage tensor dominates the damage evolution for<br />

the case <strong>of</strong> interface debond. For void nucleation (and coalescence) we assume that the<br />

85


Figure 5.4: Schematic <strong>of</strong> interface debond and void nucleation in damage model for<br />

deformations along principle axis<br />

terms D (2)<br />

11 (D1) and D (2)<br />

33 (D3) terms dominate for voids that oriented along the local x1<br />

and x3 axis, respectively. However, for the case <strong>of</strong> small deformation simple shear the<br />

formulation needs to be refined further. Consider the case <strong>of</strong> simple shear in the x1 − x2<br />

plane, as shown in fig.(5.5). We assume that principal stretch in the body (which is<br />

normal to the plane oriented 45 ◦ to the x1 direction) accounts for the interface debond in<br />

the material. In Chapter 6, we shall revisit these assumptions and examine them more<br />

closely.<br />

For the remaining part <strong>of</strong> this chapter, for notational convenience we switch to the<br />

Voigt notation in order to condense the final form <strong>of</strong> the equations. We use σ1 = σ11,σ2 =<br />

σ22,σ3 = σ33,σ4 = σ23,σ5 = σ13 and σ6 = σ12 a similar number convention follows for the<br />

tensorial components <strong>of</strong> ɛ and D. In order to derive the necessary constitutive relations<br />

we now consider the free-energy <strong>of</strong> the system ψ. <strong>The</strong> free energy can be partitioned into<br />

86


Figure 5.5: Schematic <strong>of</strong> interface debond in RVE undergoing simple shear deformation<br />

in x1 − x2 plane with the principal stretch axis x ′ 2 oriented 45 ◦ to the local x2 axis<br />

its undamaged (superscript ‘0’ ) and damaged (superscript ‘1’ ) parts.<br />

ψ = ψ 0 + ψ 1<br />

(5.29)<br />

We first consider the undamaged part <strong>of</strong> the energy response. Since we do impose or-<br />

thotropic symmetry on the material the energy can be further partitioned to a pure<br />

volumetric and deviatoric (isochoric) deformation state, that is we have,<br />

ψ 0 = ψ 0 vol + ψ 0 dev<br />

(5.30)<br />

A quadratic energy function is chosen such that the stress-strain relation prior to damage<br />

initiation remain linear elastic; Hooke’s law behavior. Hence, we have<br />

ψ 0 vol = c1ɛ 2 1 + c2ɛ 2 2 + c3ɛ 2 3 + c4ɛ1ɛ2 + c5ɛ1ɛ3 + c6ɛ2ɛ3 (5.31)<br />

ψ 0 dev = c7ɛ 2 4 + c8ɛ 2 5 + c9ɛ 2 6 (5.32)<br />

87


In order to construct the damaged response <strong>of</strong> the system and we into account the in-<br />

variance <strong>of</strong> the tensorial quantities with respect to tensor transformations and we employ<br />

the irreducible integrity basis (Spencer, 1971). Using the integrity basis Adkins (1959)<br />

we have the following set <strong>of</strong> invariants for damage in modes, α = 1 and 2.<br />

D1,D2,D3,D 2 4,D 2 5,D 2 6, D4D5D6, ɛ4D4,ɛ5D5, ɛ6D6, ɛ4ɛ5D6, ɛ5ɛ6D4, ɛ4ɛ6D5, ɛ4D5D6,<br />

ɛ5D4D6, ɛ4D5D6.<br />

Assuming small damage the second order terms vanish that is the coefficients for<br />

terms containing D 2 4,D 2 5,D 2 6, D4D5D6 is zero. Since orthotropic symmetry is preserved at<br />

all times in the true principal coordinates, the shear-extension coupling coefficients must<br />

be absent hence, ɛ6D6, ɛ4ɛ5D6, ɛ5ɛ6D4, ɛ4ɛ6D5, ɛ4D5D6, ɛ5D4D6, ɛ4D5D6 are neglected.<br />

As a consequence the energy partitioning is still feasible,<br />

ψ 1 = ψ 1 vol + ψ 1 dev<br />

Admissible free energy function that satisfies the prescribed criterion is,<br />

ψ 1 vol = A1ɛ 2 1D1 + A2ɛ 2 2D1 + A3ɛ 2 3D1 + A4ɛ 2 1D2 + A5ɛ 2 2D2<br />

+ A6ɛ 2 3D2 + A7ɛ 2 1D3 + A8ɛ 2 2D3 + A9ɛ 2 3D3<br />

+ A10ɛ1ɛ2D1 + A11ɛ1ɛ2D2 + A12ɛ1ɛ2D3 + A13ɛ1ɛ3D1<br />

+ A14ɛ1ɛ3D2 + A15ɛ1ɛ3D3 + A16ɛ2ɛ3D1 + A17ɛ2ɛ3D2<br />

(5.33)<br />

+ A18ɛ2ɛ3D3 (5.34)<br />

ψ 1 dev = A19ɛ 2 4D1 + A20ɛ 2 5D1 + A21ɛ 2 6D1 + A22ɛ 2 4D2 + A23ɛ 2 5D2<br />

+ A24ɛ 2 6D2 + A25ɛ 2 4D3 + A26ɛ 2 5D3 + A27ɛ 2 6D3 (5.35)<br />

<strong>The</strong> Cauchy stress in the system can now be derived using equation (5.17). Hence, the<br />

constitutive law for polymer nanocomposite can be written in short-hand (Voigt notation)<br />

88


as,<br />

Where, {σ} is the 6 × 1 vector {σ1, σ2, σ3, σ4, σ5, σ6}, while<br />

{σ} = [C 0 ] + [C 1 ] {ɛ} (5.36)<br />

{ɛ} is the vector <strong>of</strong> tensorial strain components {ɛ1, ɛ2, ɛ3, ɛ4, ɛ5, ɛ6}<br />

[C 0 ] is the 6 × 6 undamaged orthotropic stiffness matrix,<br />

⎡<br />

[C 0 ⎢<br />

] = ⎢<br />

⎣<br />

2c1 c4 c5 0 0 0<br />

c4 2c2 c6 0 0 0<br />

c5 c6 2c3 0 0 0<br />

0 0 0 2c7 0 0<br />

0 0 0 0 2c8 0<br />

0 0 0 0 0 2c9<br />

⎤<br />

⎥<br />

⎦<br />

(5.37)<br />

Where, the constants ci(i = 1, 2, 3 . . . 9) are related to the undamaged nanocomposite<br />

material stiffnesses as follows,<br />

c1 = C0 11<br />

2 , c2 = C0 22<br />

2 , c3 = C0 33<br />

2 , c4 = C0 12<br />

2 , c5 = C0 13<br />

2 , c6 = C0 23<br />

2<br />

c7 = C 0 44, c8 = C 0 55, c9 = C 0 66 (5.38)<br />

Where, the undamaged material stiffness components C 0 ij can be written in terms <strong>of</strong><br />

undamaged material constants E 0 1, E 0 2, E 0 3 (Young’s modulus along x1,x2 and x3 axis<br />

directions, respectively), ν 0 23, ν 0 13, ν 0 12 (Poisson’s ratio effect in the 2-3, 1-3 and 1-2 planes),<br />

89


µ 0 23, µ 0 13 and µ 0 12 (shear modulus in the 2-3, 1-3 and 1-2 planes repectively).<br />

Where,<br />

Υ =<br />

ν 0 32 = ν 0 E<br />

23<br />

0 2<br />

E0 3<br />

ν 0 31 = ν 0 E<br />

13<br />

0 3<br />

E0 1<br />

ν 0 21 = ν 0 E<br />

12<br />

0 2<br />

E0 1<br />

C 0 11 = E 0 1(1 − ν 0 23ν 0 32)Υ (5.39)<br />

C 0 22 = E 0 2(1 − ν 0 13ν 0 31)Υ (5.40)<br />

C 0 33 = E 0 3(1 − ν 0 12ν 0 21)Υ (5.41)<br />

C 0 23 = E 0 2(ν 0 32 + ν 0 12ν 0 31)Υ (5.42)<br />

C 0 13 = E 0 3(ν 0 13 + ν 0 12ν 0 23)Υ (5.43)<br />

C 0 12 = E 0 1(ν 0 31 + ν 0 21ν 0 32)Υ (5.44)<br />

C 0 44 = µ 0 23 (5.45)<br />

C 0 55 = µ 0 13 (5.46)<br />

C 0 66 = µ 0 12 (5.47)<br />

1<br />

1 − ν 0 12ν 0 21 − ν 0 23ν 0 32 − ν 0 13ν 0 31 − 2ν 0 21ν 0 32ν 0 13<br />

90


<strong>The</strong> damaged stiffness [C 1 ] is given as,<br />

Where,<br />

⎡<br />

[C 1 ⎢<br />

] = ⎢<br />

⎣<br />

C 1 11 C 1 12 C 1 13 0 0 0<br />

C 1 12 C 1 22 C 1 23 0 0 0<br />

C 1 13 C 1 23 C 1 33 0 0 0<br />

0 0 0 C 1 44 0 0<br />

0 0 0 0 C 1 55 0<br />

0 0 0 0 0 C 1 66<br />

⎤<br />

⎥<br />

⎦<br />

(5.48)<br />

C 1 11 = 2 (A1D1 + A4D2 + A7D3) (5.49)<br />

C 1 22 = 2 (A2D1 + A5D2 + A8D3) (5.50)<br />

C 1 33 = 2 (A3D1 + A6D2 + A9D3) (5.51)<br />

C 1 12 = A10D1 + A11D2 + A12D3 (5.52)<br />

C 1 13 = A13D1 + A14D2 + A15D3 (5.53)<br />

C 1 23 = A16D1 + A17D2 + A18D3 (5.54)<br />

C 1 44 = 2 (A19D1 + A22D2 + A25D3) (5.55)<br />

C 1 55 = 2 (A20D1 + A23D2 + A26D3) (5.56)<br />

C 1 66 = 2 (A21D1 + A24D2 + A27D3) (5.57)<br />

<strong>The</strong> number constants to be determined can be reduced by using the primary assump-<br />

tions that were mentioned earlier in the section. Since we assume that the primary<br />

damage modes in each principal direction <strong>of</strong> the nanocomposite does not interact, we<br />

have A4=A7=A2=A8=A3=A6=0. <strong>The</strong> argument is extended to characterize the effect<br />

<strong>of</strong> damage on Poisson’s ratio while setting, A12=A14=A16=0. From, our discussions on<br />

91


simple shear in the x1 − x2 or the x2 − x3 plane, we assume that D2 term dominates the<br />

damage evolution in these planes, such that D1, D3 ≪ D2. Hence, A19=A25=A21=A27=0.<br />

Finally we assume that the damage entity for simple shear in the x1 − x3 (i.e., void co-<br />

alescence) is mainly dominated by the D3 term with the damage opening given by the<br />

maximum principle stretch. <strong>The</strong> damaged stiffness response reduces to,<br />

C 1 11 = 2A1D1 (5.58)<br />

C 1 22 = 2A5D2 (5.59)<br />

C 1 33 = 2A9D3 (5.60)<br />

C 1 12 = A10D1 + A11D2 (5.61)<br />

C 1 13 = A13D1 + A15D3 (5.62)<br />

C 1 23 = A17D2 + A18D3 (5.63)<br />

C 1 44 = 2A22D2 (5.64)<br />

C 1 55 = 2A26D3 (5.65)<br />

C 1 66 = 2A24D2 (5.66)<br />

<strong>The</strong> constants Ai are characterized through atomistic MD simulations. In Chapter 6, the<br />

results for the damage model are presented. <strong>The</strong> above model can describe the evolution<br />

<strong>of</strong> damage using the damage entity. In the next section we provide an analytical model<br />

for the damage entity based on the inter-atomic interaction potentials.<br />

5.2.1 A Damage Model for Interphase Debond and Void Nu-<br />

cleation in <strong>The</strong>rmoplastic Polymer Nanocomposites<br />

In the proposed damaged model the two primary and competing modes <strong>of</strong> damage in<br />

a polymer nanocomposite is the interphase debond behavior and void nucleation in the<br />

92


Figure 5.6: Schematic <strong>of</strong> nano-clay with dimensions<br />

process zone due to the atoms moving apart in the direction <strong>of</strong> maximum stretch. We<br />

arbitrarily assign interphase debond as damage mode ‘1’ and void formation in the bulk<br />

as mode ‘2’. For a general state <strong>of</strong> stress, all damage modes are assumed to coexist in<br />

the material with no interaction between the constituent damage entities. Consider a<br />

nano-particle oriented along the local 1-direction with the normal to the damage surface<br />

given by n = {0, 1, 0} T as shown in fig.(5.6), then we assume the damage entity due<br />

interphase debond is defined as,<br />

D2 = 1<br />

b 3 2a2Lnwn if λ > λC where λ = r/σ (5.67)<br />

Where, Ln is the length <strong>of</strong> the nanoclay, wn is the width <strong>of</strong> nanoclay and b is the length<br />

<strong>of</strong> the cubic volume box in each dimension and σ is the zero-point <strong>of</strong> the Leonard-<br />

Jones potential. <strong>The</strong> variable ‘a2’ is the damage influence function for interphase debond<br />

(magnitude <strong>of</strong> influence function) as introduced in eq.(5.3). For damage mode 2, consider<br />

ellipsoidal voids in the nanoscale RVE oriented along principal stretch directions x1 (n =<br />

{1, 0, 0} T ) and x3 (n = {0, 0, 1} T ) (see, fig.5.7). <strong>The</strong>n we assume the damage entity due<br />

93


Figure 5.7: Schematic <strong>of</strong> undeformed spherical void in the undeformed polymer (or polymer<br />

nanocomposite) and deformed ellipsoidal voids oriented along stretch directions x1<br />

and x3 axis, with b1, b2 and b3 as the major axes <strong>of</strong> the ellipsoid.<br />

to voids in the respective principal stretch directions to be given as,<br />

D1 = 1<br />

b 3 πa1b 2 2 if λ > λC<br />

D3 = 1<br />

b 3 πa3b 2 1 if λ > λC<br />

(5.68)<br />

(5.69)<br />

Where, b1, b2 and b3 are lengths <strong>of</strong> the major axes <strong>of</strong> the ellipsoidal void as shown in<br />

fig.(5.7), and a1 and a3 are the damage influence functions for void growth in the local 1<br />

and 3 directions. In order to construct the influence function for the degenerate damage<br />

modes, we consider the non-bonded interaction in the bulk polymer and at the local<br />

nano-particle and polymer molecule interface. <strong>The</strong> force due to non-bonded interaction<br />

94


has the following functional formula,<br />

F (r) = − 24<br />

r ɛ<br />

<br />

2<br />

<br />

σ<br />

12 −<br />

r<br />

<br />

σ<br />

<br />

6<br />

r<br />

(5.70)<br />

Where, ɛ is the interaction energy, σ is the zero-point <strong>of</strong> the energy, while r is the distance<br />

between the atoms at the interface <strong>of</strong> the polymer nano-clay. In MD simulations, the<br />

extent <strong>of</strong> interphase debond (or void coalescence) and the related damage behavior is a<br />

function <strong>of</strong> the inter-atomic potentials. Through multi-scale simulation we can embed<br />

the characteristic atomistic debond behavior into the nano-scale continuum RVE model<br />

to capture the evolution <strong>of</strong> damage in the system. In this manner, the multi-scale model<br />

avoids the need for a full atomistic description <strong>of</strong> the nano-scale damage and enables<br />

material scientists and engineers to easily incorporate the model to any design analysis<br />

s<strong>of</strong>tware to study nano-composites. In this effort, we define the influence function for the<br />

three cases <strong>of</strong> damage to be,<br />

a1(λ) = b3<br />

πb2 <br />

1 −<br />

2<br />

24ɛ<br />

<br />

1<br />

2<br />

rFmax λ<br />

a2(λ) =<br />

a3(λ) = b3<br />

πb 2 1<br />

12 <br />

6<br />

1<br />

−<br />

λ<br />

b3 <br />

1 −<br />

2Lnwn<br />

24ɛ<br />

12 <br />

6<br />

1 1<br />

2 −<br />

rFmax λ λ<br />

<br />

1 − 24ɛ<br />

12 <br />

6<br />

1 1<br />

2 −<br />

rFmax λ λ<br />

(5.71)<br />

(5.72)<br />

(5.73)<br />

Where, λ is the maximum (principle) stretch <strong>of</strong> the continuum RVE (to be discussed in<br />

detail in Chapter 6), while Fmax is the maximum value <strong>of</strong> the function F (r). Physically,<br />

the damage influence function represents the evolution <strong>of</strong> opening displacement <strong>of</strong> the<br />

damage entity following damage initiation (i.e., when λ > λC). To determine Fmax we<br />

take the first derivative <strong>of</strong> eq.(5.70) w.r.t. ‘r’ and equate it to zero to obtain the maximum<br />

95


Figure 5.8: Determination <strong>of</strong> Fmax with the nano-scale damage parameters σ and rmax<br />

indicated.<br />

as shown in fig.(5.8). This gives,<br />

rmax = 6<br />

<br />

26<br />

σ (5.74)<br />

7<br />

Where, rmax is the value <strong>of</strong> ‘r’ for which F (r) is at a maximum. Substituting, eq.(5.74)<br />

in to eq.(5.70), gives<br />

Fmax = −2.396 ɛ<br />

σ<br />

Substituting for Fmax in to equations 5.71 to 5.73 we get,<br />

a1(λ) = b3<br />

πb2 <br />

1<br />

1 + 10.015 2<br />

2<br />

λ<br />

b<br />

a2(λ) =<br />

3<br />

<br />

1<br />

1 + 10.015 2<br />

2Lnwn<br />

<br />

a3(λ) = b3<br />

πb 2 1<br />

1 + 10.015<br />

<br />

2<br />

13 <br />

7<br />

1<br />

−<br />

λ<br />

13 <br />

7<br />

1<br />

−<br />

λ λ<br />

13 <br />

7<br />

1 1<br />

−<br />

λ λ<br />

(5.75)<br />

(5.76)<br />

(5.77)<br />

(5.78)<br />

It is evident from the above equations that the minimum value for Di (i=1, 2 or 3) is 0<br />

96


(no damage, no separation) while 1 indicates complete damage and complete separation.<br />

<strong>The</strong> critical value <strong>of</strong> stretch for damage initiation is when r = rmax as shown in fig.(5.8),<br />

giving,<br />

λc = rmax<br />

σ<br />

<br />

6 26<br />

=<br />

7<br />

(5.79)<br />

<strong>The</strong> components <strong>of</strong> the damage tensor can now be found by substituting equations (5.76)<br />

through (5.78) in to eq.(5.67), eq.(5.68) and eq.(5.69), respectively.<br />

Di(λ) = 1 + 10.015<br />

<br />

2<br />

13 1<br />

−<br />

λ<br />

<br />

7<br />

1<br />

λ<br />

where, i = 1, 2 or 3 (5.80)<br />

We observe that all components <strong>of</strong> the damage tensor reduce to the same damage function,<br />

due to the fact that although the geometry <strong>of</strong> the damages are unique, the non-bond<br />

interactions that govern damage evolution is the same for all damage entities. In general,<br />

due to the heterogeneities <strong>of</strong> the interacting species at the nano-scale, the critical value <strong>of</strong><br />

stretch λC at which damage (debond) initiates needs to be scaled to a value ascertained<br />

from atomistic simulations. In particular, if ‘α’ is a correction factor determined from<br />

simulations then the corrected value <strong>of</strong> critical stretch λ ′ C<br />

λ ′ C = αλC<br />

is given by,<br />

(5.81)<br />

In Chapter 7, a discussion is presented on generalizing the current damage formulation<br />

for a generalized 3-dimensional RVE model.<br />

97


Chapter 6<br />

Results<br />

In this chapter the results to the multi-scale modeling techniques proposed in this<br />

dissertation are tabulated. Beginning in section 6.1 we look at an innovative scheme to<br />

concurrently couple atomistic polymer simulations to macro-scale continuum during sim-<br />

ulation time. In section 6.2, a technique to model the Helmholtz free-energy is discussed<br />

followed by section 6.3 where the results to the damage model are discussed.<br />

6.1 Concurrent Coupling <strong>of</strong> <strong>The</strong>rmoplastic Polymers<br />

In this section the results to 2-D coupling <strong>of</strong> MD simulation <strong>of</strong> atomistic thermoplastic<br />

polymer to continuum simulation <strong>of</strong> polymer using GIMP (see, Chapter 2) with the<br />

coupling algorithm proposed in Chapter 4 is discussed. <strong>The</strong> coupling method is used<br />

to solve three specific cases <strong>of</strong> polymer deformation: I. A coupled polymer model under<br />

simple tension (with no embedded crack), II. coupled polymer with an evolving mode I<br />

(crack face opening mode) crack and III. polymer model with an evolving mode II (crack<br />

tip under shear deformation) crack. In each <strong>of</strong> these cases, the displacement fields (note:<br />

u is the displacement along X axis and v is the displacement along Y ) and displacement<br />

gradients (i.e., ∂u ∂u ∂v , , ∂X ∂Y ∂X<br />

∂v and ) for the coupled simulation particles (material points<br />

∂Y<br />

and atomistic entities) are determined as a function <strong>of</strong> time.<br />

For the larger continuum, the response <strong>of</strong> the system is assumed to be in the form<br />

<strong>of</strong> the hyperelastic material model introduced in section 2.2.1 with material properties<br />

c1 = 34.3MPa and c2 = 9.59MPa that corresponds to a Young’s modulus <strong>of</strong> E=1.5GPa<br />

and Poissons ratio <strong>of</strong> 0.33 for small deformation. In all simulations the far-field (or macro-<br />

scale) loading is simulated using an applied stress σapp = 2.5kPa. In order to mitigate<br />

98


the effect <strong>of</strong> dynamic stress waves from unloading the crack tip during simulation time,<br />

the load in each case is linearly ramped after t = 25 ps (1 pico = 10 −12 ). Figure 6.1<br />

Figure 6.1: Layout <strong>of</strong> the interior (INT), internal volume cell (IVC) and surface volume<br />

cell (SVC) zones in MD simulation, with dimensions indicated in ˚A<br />

shows the layout <strong>of</strong> the atomistic zone that was used in all simulations. <strong>The</strong> interior MD<br />

zone (i.e., the process zone) is a cube <strong>of</strong> dimension 40˚A × 40˚A × 40˚A. <strong>The</strong> IVC and<br />

SVC regions are a 5 ˚A wide zones as indicated in fig.(6.1). <strong>The</strong> force field parameters for<br />

the thermoplastic polymer atomistic model are chosen from Wei et al. (2002). <strong>The</strong> total<br />

energy <strong>of</strong> the system is assumed to be a sum <strong>of</strong> the following contributions,<br />

U(r) = Ebond + Eangle + Etorsion + Enon-bond<br />

99<br />

(6.1)


Where,<br />

Ebond = <br />

bonds<br />

Eangle = <br />

angles<br />

Etorsion = <br />

torsions n=0<br />

kl 2<br />

(l − l0)<br />

2<br />

kθ<br />

2<br />

(cos θ − cos θ0)<br />

2<br />

3<br />

Enon-bond = <br />

σ 4ɛ<br />

r<br />

i<br />

j>i<br />

Cn cos n (ω)<br />

12<br />

−<br />

<br />

σ<br />

<br />

6<br />

r<br />

(6.2)<br />

<strong>The</strong> constants are as follows: kl = 346kJ/mol/˚A 2 , l0 = 1.53˚A, kθ = 520J/mol,θ0 =<br />

112.813 ◦ , C0 = 8, 832J/mol, C1 = 18, 087J/mol, C2 = 4, 880J/mol, C3 = −31, 800J/mol<br />

and ɛ = 0.498J/mol,σ = 2.95˚A. It should be noted that the cut-<strong>of</strong>f radius σ was adjusted<br />

to 2.95˚A from its recommended value <strong>of</strong> 3.95˚A in Wei et al. (2002). This adjustment<br />

in value was required to prevent void formation in the MD model during equilibration.<br />

<strong>The</strong> typical MD simulation algorithm requires the MD domain to adhere to one <strong>of</strong> the<br />

classical ensembles, e.g. NVT, NPT, NPH etc.(see Allen & Tildesley (1987)), these<br />

ensembles necessitates the use <strong>of</strong> periodic boundary conditions. However, the simula-<br />

tion box is typically non-periodic in a coupled simulation. In order to circumvent this<br />

problem, the MD box is mapped to a bigger box (with much greater volume) without<br />

rescaling the internal coordinates <strong>of</strong> the MD atoms. In this manner we can ‘trick’ the<br />

atomistic simulation into believing that the simulation domain is indeed periodic (see<br />

fig.6.2). <strong>The</strong> primary objective <strong>of</strong> the coupled simulation is to ensure that the continuity<br />

<strong>of</strong> field variables are preserved during simulation time. Hence, in this section the contour<br />

plots <strong>of</strong> displacements and displacement gradients <strong>of</strong> particles are included in order to<br />

illustrate the effectiveness <strong>of</strong> proposed coupled simulation in maintaining continuity <strong>of</strong><br />

field variables. <strong>The</strong> timestep for the continuum simulations was 0.5fs (1 femto = 10 −15 ).<br />

For the purpose <strong>of</strong> temporal scaling (see Chapter 4) we use a factor <strong>of</strong> 50, such that<br />

100


Figure 6.2: Layout <strong>of</strong> a non-periodic MD simulation box<br />

the scaled atomistic time-step (∆τ) is 0.01fs. Zhou (2003) argues that in order to accu-<br />

rately determine the value <strong>of</strong> continuum stress for a discrete system long time-averages<br />

and a sufficiently large atomistic (discrete) system size is required. Since the temporal<br />

space over which the system is averaged (N = 50 timesteps) is quite small, we choose<br />

to plot the gradients <strong>of</strong> the displacement field rather than the instantaneous values <strong>of</strong><br />

Cauchy stress. Given below is a description <strong>of</strong> the method employed in this work to<br />

determine the displacement gradients by post-processing the coupled simulation data. In<br />

Figure 6.3: Layout <strong>of</strong> grid and particle in post-processing <strong>of</strong> concurrent coupled simulation<br />

to compute displacement gradient<br />

101


this description we designate all particles (i.e., continuum material points and atomistic<br />

entities) as p. Let x k p and x k+1<br />

p<br />

be the current and updated <strong>of</strong> the particle at step ‘k’ (k<br />

assumes integral values). Further, let Xp be the undeformed coordinate <strong>of</strong> the particles<br />

such that x k=1<br />

p<br />

= Xp. We superpose the displacement solution at step ‘k’ on a regular<br />

and structured finite element grid (see fig.6.3). From the displacement solution we have<br />

the incremental nodal displacement vector ∆u k computed as, ∆u k = x k+1 − x k . <strong>The</strong>n,<br />

assuming that the incremental displacements are small such that the particle does not<br />

cross its current element boundaries, we can compute the incremental displacements on<br />

node I <strong>of</strong> the background grid using the finite element shape functions NpI (the shape<br />

function between node ‘I’ and particle ‘p’). That is we have,<br />

∆U k<br />

I =<br />

M I p<br />

<br />

p=1<br />

NpI∆u k p<br />

(6.3)<br />

Where, M I p is the number <strong>of</strong> particles in the neighborhood <strong>of</strong> node I. This number is<br />

simply the number <strong>of</strong> atoms that have a non-zero contribution to node I, as dictated by<br />

the atom to node shape function NpI. An average nodal incremental displacement vector<br />

is computed as follows,<br />

∆U k<br />

I = ∆UkI<br />

M I p<br />

(6.4)<br />

For a particle ‘p’ that lies within the boundaries <strong>of</strong> element ‘e’ (see, fig.6.3), the gradient<br />

<strong>of</strong> the incremental displacement field ‘∂∆u k p/∂x k ’ can be computed at each particle using<br />

the following equation,<br />

∂∆uk N<br />

p<br />

=<br />

∂xk (e)<br />

∂NpI<br />

∆Uk<br />

∂xk I<br />

I=1<br />

(6.5)<br />

Where, N (e) is the number <strong>of</strong> nodes in element ‘e’. It should be noted that the derivative<br />

<strong>of</strong> the shape function ∂NpI/∂x k is with respect to current coordinates x k p <strong>of</strong> the particle.<br />

102


<strong>The</strong> deformation gradient for particle p, F k<br />

p can then be updated using,<br />

F k+1<br />

p<br />

= F k<br />

p ·<br />

<br />

I + ∂∆uk p<br />

∂x k<br />

<strong>The</strong> displacement gradient (Malvern, 1969) can then be found using,<br />

∂u k+1<br />

p<br />

∂X<br />

<br />

(6.6)<br />

= Fk+1<br />

p − I (6.7)<br />

It should be noted that the a coarse mesh will act to “smear” the atomistic details <strong>of</strong><br />

deformation in computing the displacement gradients. <strong>The</strong> solution for displacement<br />

gradients are hence inherently mesh size dependent.<br />

6.1.1 Case I: Benchmark <strong>of</strong> Coupling algorithm using a Simple<br />

Tension Test<br />

Figure 6.4: Schematic <strong>of</strong> coupled model in a simple tension setup<br />

Given in fig.(6.4) is the schematic <strong>of</strong> the thermoplastic polymer tension specimen with<br />

103


the appropriate boundary conditions. <strong>The</strong> system evolves under the influence <strong>of</strong> the stress<br />

applied on the edge located at Y=400˚A. Figure (6.5) and (6.6) is the time progression<br />

Figure 6.5: Time progression <strong>of</strong> vertical displacements (v) <strong>of</strong> the coupled simulation<br />

particles at (a) t = 1fs (b) t= 20ps (c) t = 40ps and (d) t=60ps. In this view the<br />

particle displacements are plotted along the vertical axis and undeformed coordinates <strong>of</strong><br />

the particles are on the X-Y plane. <strong>The</strong> specimen is loaded at the edge on the right which<br />

corresponds to the line Y=400˚A<br />

<strong>of</strong> the displacements and gradients, respectively. <strong>The</strong> figures are viewed in the Y-Z with<br />

the undeformed coordinates <strong>of</strong> the 2-D simulation plotted in the X-Y plane and the field<br />

variable (displacement/displacement gradient) plotted along Z. As can be seen there is a<br />

large variation in displacements and gradients <strong>of</strong> the particles inside the atomistic zone.<br />

This is attributed to the thermal vibrations due to non-zero, finite temperature simulation<br />

inside the MD zone. However, we see that the average behavior <strong>of</strong> the atoms satisfies the<br />

necessary continuity requirements.<br />

104


Figure 6.6: Time progression <strong>of</strong> displacement gradient (∂v/∂Y ) <strong>of</strong> the coupled simulation<br />

particles at (a) t = 0.5 fs (b) t= 10 ps (c) t = 20 ps and (d) t=30 ps. In this view the<br />

displacement gradients are plotted along the vertical axis and undeformed coordinates<br />

<strong>of</strong> the particles are on the X-Y plane. <strong>The</strong> specimen is loaded at the edge on the right<br />

which corresponds to the line Y=400˚A<br />

6.1.2 Case II: Mode I Crack Growth in <strong>The</strong>rmoplastic Polymer<br />

Next we consider a more complex application <strong>of</strong> the coupling algorithm. Shown in fig.(6.7)<br />

is the layout <strong>of</strong> the coupled model to study mode I crack growth in a thermoplastic<br />

polymer system. <strong>The</strong> pre-existing continuum crack shown in the figure is extended by<br />

applying an external applied stress as indicated. <strong>The</strong> strain energy release rate (SERR)<br />

at the crack tip in the continuum is calculated using the virtual crack closure integral<br />

discussed in sec.(2.3). <strong>The</strong> critical value <strong>of</strong> SERR for mode I crack growth is chosen to<br />

be GIc = 10 −7 J/m 2 , with an initial flaw size (a) <strong>of</strong> 20 ˚A (fig.(6.7)). Since, the primary<br />

focus <strong>of</strong> the algorithm is crack growth within the process zone (i.e. atomistic polymer<br />

zone), the chosen value for critical SERR was assumed small (and in general physically<br />

105


Figure 6.7: Schematic <strong>of</strong> coupled model for mode I crack growth in thermoplastic polymer<br />

unrealistic) in order to extend the crack surface into the MD simulation zone within<br />

a reasonable number <strong>of</strong> coupled computational steps. Figure 6.8 shows the evolution<br />

<strong>of</strong> gradient <strong>of</strong> the vertical (v) displacement field in the Y-direction (∂v/∂Y ) at various<br />

instants <strong>of</strong> simulation. As can be seen the simulation shows that as the crack extends into<br />

the MD domain, significant displacement gradients occur. <strong>The</strong> spread <strong>of</strong> high gradients<br />

inside the amorphous polymer MD domain (contour level red in fig.6.8(d)) indicate that a<br />

diffused damage occurs inside the MD domain and is not a singular crack tip as one would<br />

expect in a simulation <strong>of</strong> crack propagation in an ordered solid lattice. Figure 6.9 shows<br />

the vertical displacement (v) <strong>of</strong> the particles as the mode I crack enters the MD domain.<br />

<strong>The</strong> contours indicate good continuity <strong>of</strong> displacements in the coupled simulation <strong>of</strong> a<br />

mode I propagating crack. <strong>The</strong> simulations methodology also allows the user to examine<br />

fracture surfaces in detail with embedded atomistic information (see, fig.6.10).<br />

106


Figure 6.8: ∂v/∂Y gradient pr<strong>of</strong>iles for a propagating 2D mode I crack in coupled simulation<br />

<strong>of</strong> a thermoplastic polymer system at times t = (a) 0.5 fs (b) 10 ps (c) 25 ps, and<br />

(d) 80 ps<br />

6.1.3 Case III: Mode II crack growth in <strong>The</strong>rmoplastic Polymer<br />

In the final coupled simulation benchmark example we consider mode II shear crack prop-<br />

agation in a thermoplastic polymer. In short beam shear tests <strong>of</strong> E-glass/polypropylene<br />

nanoclay and E-glass/nylon-6 nanoclay composites significant improvements in mechani-<br />

cal properties were observed. It is envisioned that the coupled simulation scheme proposed<br />

here can be utilized to study nano-composite specimens in shear. <strong>The</strong> coupling simulation<br />

metholodology becomes all the more relevant in the context <strong>of</strong> nano-composite simula-<br />

tion since nanoclay platelets have unusually high aspect ratios making them difficult to<br />

model in regular sized MD domains while keeping the computational costs involved to<br />

a minimum. However, in this example the MD model is <strong>of</strong> a purely amorphous poly-<br />

mer system with no embedded nanoclay (fig.6.11). As in mode I simulation, the initial<br />

107


Figure 6.9: Vertical displacement (v) contours for a propagating 2D mode I crack in<br />

coupled simulation <strong>of</strong> thermoplastic polymer system at times t = (a) 0.5 fs (b) 10 ps<br />

(c) 25 ps, and (d) 80 ps, with displacement (v) plotted along Z-axis and undeformed<br />

coordinates <strong>of</strong> the body in the X-Y plane<br />

Figure 6.10: Fracture surface <strong>of</strong> mode I crack with the atoms within the SVC (red), IVC<br />

(yellow) and interior (green) zones indicated<br />

108


Figure 6.11: Schematic <strong>of</strong> coupled model for mode II crack growth in thermoplastic<br />

polymer<br />

crack length is chosen to be 20 ˚A. <strong>The</strong> critical SERR for extension <strong>of</strong> crack surface<br />

GIIc, is chosen to be 10 −7 J/m 2 . <strong>The</strong> reasons for choosing this un-realistically low value<br />

is the same as before. Figure 6.12, shows the contours <strong>of</strong> horizontal displacement (u)<br />

as a function <strong>of</strong> time as the crack is propagated in the continuum. We see at the final<br />

time-step the crack face has entered the MD domain, as indicated by the discontinuity in<br />

displacements (fig.6.12(d)). Figure 6.13 and 6.14 are contours <strong>of</strong> displacement gradients<br />

∂u<br />

∂Y<br />

and ∂v<br />

∂X<br />

at various instances. <strong>The</strong> figures indicate a discontinuity in the gradients to<br />

the displacements especially as the crack impinges on the MD domain. <strong>The</strong> discontinuity<br />

could be an indication <strong>of</strong> difference in material properties between the two zones. <strong>The</strong><br />

redistribution and localization <strong>of</strong> stresses inside the MD domain could also add to the<br />

discontinuity. It is suggested that future simulations should try to incorporate better<br />

material models in the continuum domain such that the continuity <strong>of</strong> gradients can be<br />

achieved and thereby ensuring a reliable and uniform energy density around the coupling<br />

domain. It is also observed that the use <strong>of</strong> non-bonded interactions to transfer forces<br />

109


Figure 6.12: Horizontal displacement (u) contours for a propagating 2D mode II crack<br />

in coupled simulation <strong>of</strong> thermoplastic polymer system at times t = (a) 0.5 fs (b) 25 ps<br />

(c) 30 ps, and (d) 32.5 ps, with displacement (v) plotted along Z-axis and undeformed<br />

coordinates <strong>of</strong> the body in the X-Y plane<br />

from the continuum down to atomistic domain contributes to the dissipation <strong>of</strong> energy<br />

at the MD/continuum interface. <strong>The</strong> coarse grained amorphous polymer system creates<br />

voids inside the MD system during the equilibration process. <strong>The</strong> formation <strong>of</strong> voids<br />

effectively reduces the interaction <strong>of</strong> the IVC atoms with the atoms inside the interior<br />

zone. This concept is not unlike the mean-free path in gases. <strong>The</strong> lower the mean-free<br />

path the better is the momentum transfer between gas particles. Needless to say, the<br />

formation <strong>of</strong> gaps (or alternatively increase in mean-free path) affects the ability <strong>of</strong> the<br />

handshake region to transfer internal forces from the continuum domain to the atomistic<br />

system. It is envisioned that switching to a full atom (non-coarse grained) MD system<br />

will help in mitigating some <strong>of</strong> the issues encountered with the current coupling scheme<br />

by ensuring a better packing density <strong>of</strong> polymer atoms with a reduced amount <strong>of</strong> voids<br />

110


and other related anomalies.<br />

Figure 6.13: Contours <strong>of</strong> displacement gradient ∂u/∂Y for a propagating 2D mode II<br />

crack in coupled simulation <strong>of</strong> thermoplastic polymer system at times t = (a) 0.5 fs<br />

(b) 25 ps (c) 30 ps, and (d) 32.5 ps, with ∂u/∂Y plotted along Z-axis and undeformed<br />

coordinates <strong>of</strong> the body in the X-Y plane<br />

6.2 Free Energy Calculations <strong>of</strong> <strong>The</strong>rmoplastic Polymers<br />

<strong>The</strong> atomistic deformation and the related thermodynamics <strong>of</strong> a material are a function<br />

<strong>of</strong> the energetic as well as entropic contributions <strong>of</strong> the material. In disordered polymers,<br />

configurational entropy plays a significant role in its elastic properties (Buehler & Wong,<br />

2007). Hence, for a complete description <strong>of</strong> the polymer deformation the model must<br />

include the relevant thermodynamics related to the deformation in order to develop a<br />

complete description <strong>of</strong> the deformation process. In this work, we propose a scheme to<br />

calculate the Helmholtz free energy <strong>of</strong> a thermoplastic polymer undergoing mechanical<br />

deformation. Before we proceed to calculation <strong>of</strong> the Helmholtz free energy, we also look<br />

111


Figure 6.14: Contours gradient ∂v/∂X <strong>of</strong> vertical displacement (v) for a propagating 2D<br />

mode II crack in coupled simulation <strong>of</strong> thermoplastic polymer system at times t = (a) 0.5<br />

fs (b) 25 ps (c) 30 ps, and (d) 32.5 ps, with ∂v/∂X plotted along Z-axis and undeformed<br />

coordinates <strong>of</strong> the body in the X-Y plane<br />

at the entropic contribution <strong>of</strong> individual polymer chains <strong>of</strong> the thermoplastic polymer<br />

model used in this research.<br />

6.2.1 Calculation <strong>of</strong> Persistence Length <strong>of</strong> a Polymer<br />

<strong>The</strong> entropic contribution <strong>of</strong> individual polymer chains is quantified using the persistence<br />

length (Binder, 1995). In general it indicates the transition <strong>of</strong> the molecules from energetic<br />

dominated process to an entropic deformation for individual polymer chains. <strong>The</strong> central<br />

idea <strong>of</strong> the following computations is to capture the configurational entropy <strong>of</strong> the polymer<br />

chain. For a single polymer chain the number <strong>of</strong> configurations depend on the flexibility <strong>of</strong><br />

the polymer chain. In other words, if the chain is pliable the configurations the molecule<br />

can exist in increases and by definition it attributes to the increases in entropy <strong>of</strong> the<br />

112


system. A mechanics based approach to determine the flexibility <strong>of</strong> the polymer chain is<br />

to calculate the bending stiffness EI using simulation. <strong>The</strong> persistence length ξp is then<br />

related to the bending stiffness as,(Buehler & Wong, 2007)<br />

ξp = EI<br />

kBT<br />

(6.8)<br />

Where, kB is Boltzmann’s constant and T is the temperature. To determine the bending<br />

stiffness <strong>of</strong> the molecule we fix both ends <strong>of</strong> the polymer chain by restricting all trans-<br />

lational movement <strong>of</strong> the beads at either end <strong>of</strong> the polymer model. A force is applied<br />

at the center <strong>of</strong> mass <strong>of</strong> the polymer chain using steered molecular dynamics (SMD, see<br />

Plimpton (1995)) as illustrated in fig.(6.15). <strong>The</strong> spring constant for SMD was chosen to<br />

be 10 kcal/mol/˚A 2 . As indicated in the figure we use the coarse-grained bead model <strong>of</strong><br />

the thermoplastic polymer used in this research with 4 beads. <strong>The</strong> interaction properties<br />

<strong>of</strong> bond, angle and dihedral interactions switched on. <strong>The</strong> parameters for these force<br />

fields are given in section 6.1. Steered molecular dynamics were performed for 5 rates <strong>of</strong><br />

Figure 6.15: Schematic <strong>of</strong> polymer model for determination <strong>of</strong> bending stiffness EI<br />

spring deformation ( ˙r), ˙r = 0.01, 0.025, 0.05, 0.075 and 0.1 ˚A/fs and the bending stiffness<br />

113


was determined for each rate <strong>of</strong> loading using the following beam bending equation,<br />

EI =<br />

FappL 3<br />

48d<br />

(6.9)<br />

Where, L is the equilibrated length <strong>of</strong> the polymer chain and d is the vertical displacement<br />

<strong>of</strong> the center <strong>of</strong> center <strong>of</strong> mass. <strong>The</strong> bending stiffness were determined for each loading<br />

rate and plotted as shown in fig.(6.16)). A linear curve fit (indicated as EI(L-S) in fig.6.16)<br />

was used to extrapolate the result to ˙r = 0 in order to determine the bending stiffness EI<br />

at zero loading rate. In this manner we can eliminate all inertial effects in the bending<br />

stiffness that can arise from using the MD algorithm.<br />

Figure 6.16: Bending stiffness EI <strong>of</strong> coarse-grained polymer model as a function <strong>of</strong> loading<br />

rate ˙r<br />

<strong>The</strong> equation to curve-fit was EI(L-S)=7.027×10 −28 ˙r+2.433×10 −29 Nm 2 was obtained<br />

from the linear curve-fir shown in fig.(6.16). Substituting for EI = 2.433 × 10 −29 Nm 2 in<br />

eq.(6.8) and setting T=300K, we get ξp = 58.74˚A. This roughly translates to the length<br />

<strong>of</strong> 40 beads (assuming bond length <strong>of</strong> 1.53˚A) in the coarse grained system. Hence, if<br />

114


the unstretched length <strong>of</strong> coarse grained polymer chain is below 58.74˚A we can expect<br />

the energetic contributions <strong>of</strong> the polymer to dominate its elastic strain energy; for chain<br />

lengths above 58.74˚A the entropic contributions cannot be ignored.<br />

6.2.2 Calculation <strong>of</strong> Free energy <strong>of</strong> Polymers undergoing Me-<br />

chanical Deformation<br />

In the previous section, the conditions governing entropic contributions to elasticity from<br />

single chain polymer bead model were derived. However, a similar derivation does not<br />

exist for the MD model <strong>of</strong> the bulk polymer system. In general, entropy driven forces in<br />

MD models are harder to quantify. <strong>The</strong> Helmholtz free energy however, is a state variable<br />

that combines the energetic and entropic contributions <strong>of</strong> the polymer system 1 . In this<br />

section we use thermodynamic integration method that was introduced in sec.(3.4.1)<br />

to determine the free energy change associated with mechanical deformation. In this<br />

work the Helmholtz energy is determined for a volumetric and simple shear deformation<br />

process in a pure thermoplastic polymer system as shown in the schematic in fig.(6.17).<br />

<strong>The</strong> deformed states are realized through the deformation <strong>of</strong> the periodic MD simulation<br />

box. <strong>The</strong> update equations for the box coordinates in dilatational (i.e., pure volumetric)<br />

deformation at step ‘k’ is given by,<br />

x k 1 = (1 + α k )X1<br />

x k 2 = (1 + α k )X2<br />

x k 3 = (1 + α k )X3 (6.10)<br />

Where, x k i is the deformed coordinates <strong>of</strong> the box and Xi is the undeformed coordinate,<br />

α k is the scaling factor to dilate the simulation box. <strong>The</strong> update equations for deviatoric<br />

1 Helmholtz energy, A = U − T S, U is the potential energy, T the temperature and S is entropy<br />

115


Figure 6.17: Schematic <strong>of</strong> the dilatational and deviatoric deformation used to determine<br />

free energy change associated with mechanical deformation<br />

deformation are given by,<br />

x k 1 = X1 + β k (X2 + X3)<br />

x k 2 = X2 + β k (X1 + X3)<br />

x k 3 = X3 + β k (X1 + X2) (6.11)<br />

Again, β k is the scaling factor by which the box is deformed. <strong>The</strong> volumetric and devia-<br />

toric deformation gradients F vol and F dev respectively, can be derived from eq.(6.10) and<br />

116


(6.11), as<br />

F vol,k =<br />

F dev,k =<br />

⎡<br />

⎢ 1 + α<br />

⎢<br />

⎣<br />

k 0<br />

0<br />

1 + α<br />

0<br />

k 0<br />

0 0 1 + αk ⎤<br />

⎥<br />

⎦<br />

⎡<br />

⎤<br />

⎢<br />

⎣<br />

1 β k β k<br />

β k 1 β k<br />

β k β k 1<br />

⎥<br />

⎦<br />

(6.12)<br />

(6.13)<br />

<strong>The</strong> desired strain measure can then be calculated from the deformation gradient. For<br />

small strains, i.e., α k , β k ≪ 1 the small strain tensor for the two cases will simply be<br />

given as,<br />

ɛ vol,k =<br />

ɛ dev,k =<br />

⎡<br />

⎢ α<br />

⎢<br />

⎣<br />

k 0<br />

0<br />

α<br />

0<br />

k 0<br />

0 0 αk ⎤<br />

⎥<br />

⎦<br />

⎡<br />

⎤<br />

⎢<br />

⎣<br />

0 β k β k<br />

β k 0 β k<br />

β k β k 0<br />

⎥<br />

⎦<br />

(6.14)<br />

(6.15)<br />

Five states <strong>of</strong> deformation were used in this computation with values <strong>of</strong> α (1) = 0.0 (un-<br />

deformed configuration), α (2) = 1.25 × 10 −3 , α (3) = 2.5 × 10 −3 , α (4) = 3.75 × 10 −3 and<br />

α (5) = 5.00 × 10 −3 , similarly, β (1) = 0.0 (undeformed configuration), β (2) = 1.25 × 10 −3 ,<br />

β (3) = 2.5 × 10 −3 , β (4) = 3.75 × 10 −3 and β (5) = 5.00 × 10 −3 <strong>The</strong> Helmholtz energy is<br />

determined using the thermodynamic integration method discussed in section 3.4.1. As<br />

described in sec.3.4.1, the non-bonded parameters <strong>of</strong> the thermoplastic polymer model is<br />

ramped from λ = 0 to λ = 1 during simulation time, and the change in the Hamiltonian<br />

117


<strong>of</strong> the system with respect to internal scaling parameter λ is determined. For the example<br />

considered here the Hamiltonian for the system (H) is given as,<br />

H =<br />

N<br />

i=1<br />

p i · p i<br />

mi<br />

+ U(λ) (6.16)<br />

Where, U(λ) denotes a change in potential energy <strong>of</strong> the system as a function <strong>of</strong> the scaling<br />

parameter λ. In this example we assume that the functional form for the potential energy,<br />

including bond stretching and non-bonded interactions, is given by,<br />

U(λ) = <br />

bonds<br />

kl<br />

2 (l − l0) 2 + <br />

<br />

<br />

σ(λ)<br />

4ɛ(λ)<br />

r<br />

i<br />

j>i<br />

12<br />

<br />

6<br />

σ(λ)<br />

−<br />

r<br />

(6.17)<br />

With,ɛ(λ) = ɛ 0 (1 − λ) + ɛ 1 λ and σ(λ) = σ 0 (1 − λ) + σ 1 λ. Substituting the eq.(6.17) in to<br />

eq.(6.16) and taking derivative <strong>of</strong> the resultant equation with respect to λ gives,<br />

∂H<br />

∂λ<br />

<br />

<br />

= 4(ɛ<br />

i j>i<br />

1 − ɛ 0 σ 12 <br />

σ<br />

<br />

6<br />

) −<br />

r r<br />

+6(σ 1 − σ 0 ) ɛ<br />

<br />

σ<br />

12 <br />

σ<br />

<br />

6<br />

2 −<br />

σ r r<br />

(6.18)<br />

<strong>The</strong> values for constants are as follows: kl = 346kJ/mol/˚A 2 , l0 = 1.53˚A, ɛ 0 = 0.0498J/mol,ɛ 1 =<br />

0.498J/mol, σ 0 = 0.95˚A and σ 1 = 2.95˚A. <strong>The</strong> change in Helmholtz energy is then simply<br />

calculated as (see, sec.3.4.1),<br />

∆A =<br />

λ=1<br />

λ=0<br />

<br />

∂H<br />

dλ (6.19)<br />

∂λ λ<br />

Three point Gaussian quadrature is used to determine the integral in eq.(6.19). <strong>The</strong><br />

gaussian quadratue is a numerical tool commonly used to determine the integral <strong>of</strong> a<br />

118


function f(x) with in the domain -1 to 1.<br />

I =<br />

1<br />

Where, NG is the number <strong>of</strong> Gauss points and w i G<br />

Gauss point x i G (xi G<br />

and three-point integration rule,<br />

−1<br />

NG <br />

f(x)dx ≈ w i Gf(x i G) (6.20)<br />

i<br />

is the weight associated with the ith<br />

∈ [−1, 1]). For a change in limits <strong>of</strong> integration from [-1,1] to [0,1]<br />

1<br />

0<br />

f(x)dx ≈ 1<br />

2<br />

3<br />

w i i xG + 1<br />

Gf<br />

2<br />

Hence, eq.(6.19) can be evaluated numerically using the equation,<br />

∆A = 1<br />

2<br />

3<br />

i=1<br />

i=1<br />

w i G<br />

<br />

∂H<br />

∂λ λi<br />

<strong>The</strong> values <strong>of</strong> λi for which the ensemble average <br />

∂H<br />

∂λ λi<br />

(6.21)<br />

(6.22)<br />

is evaluated are listed in Table<br />

6.1. In the determination <strong>of</strong> the change in free energy for each deformation case, three<br />

Table 6.1: Values <strong>of</strong> scaling parameter λ to determine ∆A<br />

i w i G x i G λi<br />

1 0.556 -0.775 0.113<br />

2 0.889 0 0.500<br />

3 0.556 0.775 0.887<br />

sets <strong>of</strong> simulation must be carried out. <strong>The</strong> Langevin thermostat was used to maintain<br />

temperature <strong>of</strong> the MD system at 300K. <strong>The</strong> time step for MD simulations was set at 0.5<br />

fs. <strong>The</strong> algorithm to determine the change in free energy was implemented using the MD<br />

simulation code LAMMPS. <strong>The</strong> steps involved in determining the free energy change are<br />

listed below:<br />

1. For step k: Update deformation state <strong>of</strong> the MD simulation box using the appro-<br />

119


priate coordinate update equations (eq.6.10 or 6.11).<br />

2. Set i = 1 and repeat the following steps from i = 1 to i = 3.<br />

(a) Choose λ = λi (from Table 6.1)<br />

(b) Run MD for 40 ps and determine the time average <strong>of</strong> <br />

∂H<br />

∂λ λi<br />

eq.6.18) for the last 10 ps.<br />

(computed using<br />

3. Evaluate change in free energy ∆A using eq.(6.22) and proceed to step k + 1.<br />

Figures 6.18 and 6.19 are the results to the Helmholtz energy computation for volumetric<br />

and deviatoric deformations respectively.<br />

Figure 6.18: ∆A versus volumetric strain trace(ɛ vol ) = ɛ vol<br />

kk<br />

<strong>The</strong> results to the Helmholtz computations show a “-” due to the fact that the ther-<br />

modynamic system corresponding to λ = 0 has a larger free energy relative to the system<br />

120


Figure 6.19: ∆A versus deviatoric strain ɛdev 12 = ɛdev 13 = ɛdev 23<br />

at λ = 1 (the original thermoplastic polymer). It has been shown using the thermody-<br />

namic integration approach that we are able to clearly quantify the increase in Helmholtz<br />

energy with mechanical deformation. However, as discussed in Chapter 3, absolute val-<br />

ues <strong>of</strong> this increase is computationally difficult to obtain for the amorphous system we<br />

have chosen to simulate. In the view <strong>of</strong> this author, the ability to quantify the change<br />

in Helmholtz energy with respect to its mechanical behavior holds promise for future<br />

research into determining a material model which is able to capture to both the ener-<br />

getic and entropic contributions <strong>of</strong> the polymer to its free energy. Given in appendix A<br />

is the theoretical formalisms to extract a stress-strain law based on the volumetric and<br />

deviatoric response <strong>of</strong> the system interms <strong>of</strong> the hydrostatic stress and deviatoric stress<br />

tensor. It is recommended that for future simulations that all parameters <strong>of</strong> the energy<br />

121


functional eq.3.34 (Chapter 3) are varied as a function <strong>of</strong> λ in order to sample all possible<br />

occupation volumes <strong>of</strong> the phase space.<br />

6.3 Multi-scale Damage model for <strong>The</strong>rmoplastic Nanocompos-<br />

ites<br />

While section 6.1 attempts to couple atomistic simulation to continuum solution during<br />

simulation time in this section, we will look at constitutive modeling <strong>of</strong> the nanoclay<br />

polymer composite introduced in Chapter 3 using a purely atomistic (MD) simulations<br />

approach. <strong>The</strong> primary objective <strong>of</strong> this section will be to characterize and correlate the<br />

stress-strain behavior <strong>of</strong> the MD nano-composite model to the damage model introduced<br />

in Chapter 5. In determining material properties <strong>of</strong> the nanocomposite (and pure poly-<br />

mer) system the simulation scheme employs kinematic deformation (i.e., strain control)<br />

<strong>of</strong> the simulation box followed by proper ensemble averaging (Zhou, 2003) to determine<br />

the mechanical stress that develops in the bulk. In this work the volume averaged Cauchy<br />

stress is the measure <strong>of</strong> internal force that develops in the atomistic model in response to<br />

the applied deformation. It is determined using a time average <strong>of</strong> the following equation,<br />

σαβ = 1<br />

2Ω<br />

<br />

i<br />

j<br />

r α ijf β<br />

ij<br />

(6.23)<br />

Where, α, β are indicial variables that could be either = 1, 2 or 3, r α ij is α component <strong>of</strong><br />

the vector joining atoms i and j, while f α ij is the interaction force between atoms i and<br />

j that is derived from the total potential energy <strong>of</strong> system. Figure 6.20 is a schematic<br />

<strong>of</strong> the two primary deformations used to study stress response <strong>of</strong> the system. While<br />

the name “molecular dynamics” implies a dynamic stress-strain response <strong>of</strong> the system<br />

at extremely fine time-scales, the algorithm can be employed to determine quasi-static<br />

material properties from sufficiently long simulations and time-samplings. Given below<br />

122


Figure 6.20: Schematic <strong>of</strong> the two primary deformations (extension and simple shear)<br />

used to determine material properties and the associated deformation gradient for each<br />

mode <strong>of</strong> deformation. Shown here are extension in x1 direction and shear in the x1 − x2<br />

plane<br />

are the steps that were followed in this research to determine the stress-strain response<br />

<strong>of</strong> the discrete atomistic system,<br />

1. Equilibrate atomistic model: <strong>The</strong> typical steps involved in equilibrating a MD ther-<br />

moplastic polymer model and a MD polymer/nano composite model are discussed<br />

in section 3.3 <strong>of</strong> Chapter 5.<br />

2. For step k, we have the deformation gradient F k . Apply F k to deform MD simula-<br />

tion box. As the simulation box is deformed, remap coordinates <strong>of</strong> atomistic entities<br />

into the deformed box. While orthogonal deformation <strong>of</strong> the box is straightforward,<br />

simple shear deformations require a triclinic box to deform the box to required state.<br />

<strong>The</strong> MD algorithm LAMMPS (Plimpton, 1995) has the ability to run simulation<br />

in a orthogonal and non-orthogonal (triclinic) box settings.<br />

3. <strong>The</strong> time step used in these calculations were 0.5 fs. Run MD simulations for 5 ps,<br />

the volume averaged stress components (6 components) were written to file every 5<br />

123


fs. <strong>The</strong> simulations are run with the Langevin thermostat set at 300K. For all cases<br />

<strong>of</strong> deformation the linear momentum <strong>of</strong> the center <strong>of</strong> mass (COM) is set to zero<br />

in order to prevent rigid translation <strong>of</strong> the atomistic entities. <strong>The</strong> stress averaging<br />

can then be performed as a post-processing <strong>of</strong> the MD data. However, this would<br />

stretch memory requirements.<br />

4. At the end <strong>of</strong> the MD run, set k = k+1 and return to step 2. If we have reached<br />

maximum strain proceed to next step.<br />

5. After the simulations are completed the results are averaged (as discussed above)<br />

to calculate the stress for each strain update. A 13 point moving average <strong>of</strong> the<br />

data is performed to further smooth the stress-strain response.<br />

It should be mentioned that although the kinematics <strong>of</strong> deformation has been defined in<br />

terms <strong>of</strong> the deformation gradient F, this is not a finite deformation model. <strong>The</strong> strain<br />

to failure is assumed to be small enough such that the non-linear terms <strong>of</strong> the finite<br />

deformation strain measures can be ignored in the interest <strong>of</strong> solution tractability.<br />

6.3.1 Simulation results for <strong>The</strong>rmoplastic Polymer Nano-composite<br />

A polymer nano-composite specimen was created and equilibrated using the steps de-<br />

scribed in sec.3.3.2. <strong>The</strong> structure <strong>of</strong> the nanoclay used in this is shown in fig.(6.21),<br />

with an aspect ratio <strong>of</strong> 4. <strong>The</strong> paramaters for bonded interactions for the nanoclay were<br />

obtained from Katti et al. (2005), while the non-bonded parameters were obtained from<br />

the CVFF force field. <strong>The</strong> total potential energy for the nanoclay is given by the following<br />

energy equation,<br />

U(r) = Ebond + Eangle + Etorsion + Enon-bond<br />

124<br />

(6.24)


Figure 6.21: Structure <strong>of</strong> nano-clay with dimensions (unequilibrated) indicated with the<br />

constituent atoms.<br />

Where,<br />

Ebond = <br />

kl(l − l0) 2<br />

bonds<br />

Eangle = <br />

angles<br />

Etorsion = <br />

torsions<br />

kθ(cos θ − cos θ0) 2<br />

Enon-bond = <br />

σ 4ɛ<br />

r<br />

i<br />

j>i<br />

V [1 + cos(nϕ − δ)]<br />

12<br />

−<br />

<br />

σ<br />

<br />

6<br />

r<br />

(6.25)<br />

In order to tabulate the force field parameters we use O1 to indicate the oxygen in the<br />

octahedral sheet and O2 as surface oxygen. Further, C represents the coarse grained<br />

polymer atom. Listed in Tables 6.2 through 6.4 are the bonded interaction parameters.<br />

<strong>The</strong> non-bonded interactions between polypropylene coarse grained atoms and nanoclay<br />

and the internal nanoclay non-bonded interactions are listed in Table 6.5. <strong>The</strong> parame-<br />

125


Table 6.2: Bond parameters for Nanoclay<br />

Bond Atom i Bond Atom j l0(˚A) kl(kcal/mol/˚A 2 )<br />

Al O1 1.955 382.617<br />

O1 H 0.988 827.209<br />

Si O2 1.626 492.387<br />

Si O1 1.62 558.593<br />

Table 6.3: Angle parameters for Nanoclay<br />

Bond Atom i Bond Atom j Bond Atom k θ0( ◦ ) kθ(kcal/mol/rad 2 )<br />

Al O1 H 118 34.933<br />

Si O1 Al 124.5 155.754<br />

O2 Si O1 110.6 97.493<br />

Al O1 Al 109.5 198.27<br />

O1 Al O1 90/180 80.0<br />

Table 6.4: Torsion parameters for Nanoclay<br />

Bond Atom i Bond Atom j Bond Atom k Bond Atom l nϕ V (kcal/mol)<br />

O2 Si O1 Al 0 0.722<br />

Table 6.5: Non-bond parameters for Nanoclay-Polymer hybrid<br />

Bond Atom i Bond Atom j σij(˚A) ɛij(kcal/mol)<br />

O O 2.86 0.228<br />

Si Si 4.05 0.04<br />

Al Al 4.05 0.04<br />

C O 3.33 0.094<br />

C Al 3.96 0.0395<br />

C Si 3.96 0.0395<br />

O Si 3.404 0.0955<br />

O Al 3.404 0.0955<br />

Si Al 4.053 0.04<br />

ters for the various energetic interactions inside the thermoplastic polymer system is still<br />

the same as mentioned in sec.(6.1). <strong>The</strong> volume percentage <strong>of</strong> nano-clay is 4.72% which<br />

is the ratio <strong>of</strong> nanoclay volume to the volume <strong>of</strong> the simulation box. For a 1500 coarse<br />

grained polymer system, the weight fraction <strong>of</strong> the polymer nano-composite modeled in<br />

this work is 6.8%, calculated as WNC/(WNC +WPP) where WNC is the weight if nano-clay<br />

and WP P is weight <strong>of</strong> the polymer. <strong>The</strong> aspect ratio <strong>of</strong> the nano-composite that is, the<br />

126


length along x direction divided by the width (z dimension) is 4.19, which is significantly<br />

lower than conventional polymer nano-clay composites. For the purposes <strong>of</strong> model de-<br />

velopment this model <strong>of</strong> nano-clay could serve as the foundation for future work. <strong>The</strong><br />

polymer nano composite has the nano-clay oriented along the long x1 axis <strong>of</strong> the box (see<br />

fig.6.20). This particular orientation <strong>of</strong> nano-clay gives the RVE orthotropic properties.<br />

Consequently, with a set <strong>of</strong> six strain controlled simulations the entire stiffness matrix can<br />

be numerically characterized. <strong>The</strong> following six simulations were employed to completely<br />

characterize mechanical behavior <strong>of</strong> the polymer nano-composite:<br />

1. Case 1: Extension along x1 axis with the dimensions in x2 and x3 constrained<br />

2. Case 2: Extension along x2 axis with the dimensions in x1 and x3 constrained<br />

3. Case 3: Extension along x3 axis with the dimensions in x1 and x2 constrained<br />

4. Case 4: Simple shear in x2 − x3 plane<br />

5. Case 5: Simple shear in x1 − x3 plane<br />

6. Case 6: Simple shear in x1 − x2 plane<br />

For the sake <strong>of</strong> completeness we revisit some <strong>of</strong> the pertinent equations from Chapter<br />

5. We know that the stiffness matrix [C] is a linear combination <strong>of</strong> the damaged and<br />

undamaged stiffness.<br />

[C] = [C 0 ] + [C 1 ] (6.26)<br />

127


Where, [C 0 ] is the 6 × 6 undamaged orthotropic stiffness matrix,<br />

⎡<br />

[C 0 ⎢<br />

] = ⎢<br />

⎣<br />

2c1 c4 c5 0 0 0<br />

c4 2c2 c6 0 0 0<br />

c5 c6 2c3 0 0 0<br />

0 0 0 2c7 0 0<br />

0 0 0 0 2c8 0<br />

0 0 0 0 0 2c9<br />

⎤<br />

⎥<br />

⎦<br />

(6.27)<br />

Where, the elastic constants ci(i = 1, 2, 3 . . . 9) are to be determined from MD simulations.<br />

<strong>The</strong> damaged stiffness [C 1 ] is given as,<br />

⎡<br />

[C 1 ⎢<br />

] = ⎢<br />

⎣<br />

C 1 11 C 1 12 C 1 13 0 0 0<br />

C 1 12 C 1 22 C 1 23 0 0 0<br />

C 1 13 C 1 23 C 1 33 0 0 0<br />

0 0 0 C 1 44 0 0<br />

0 0 0 0 C 1 55 0<br />

0 0 0 0 0 C 1 66<br />

128<br />

⎤<br />

⎥<br />

⎦<br />

(6.28)


Where,<br />

C 1 11 = 2A1D1 (6.29)<br />

C 1 22 = 2A5D2 (6.30)<br />

C 1 33 = 2A9D3 (6.31)<br />

C 1 12 = A10D1 + A11D2 (6.32)<br />

C 1 13 = A13D1 + A15D3 (6.33)<br />

C 1 23 = A17D2 + A18D3 (6.34)<br />

C 1 44 = 2A22D2 (6.35)<br />

C 1 55 = 2A26D3 (6.36)<br />

C 1 66 = 2A24D2 (6.37)<br />

<strong>The</strong> inelastic (damage) constants Ai are characterized through atomistic MD simulations<br />

in the following sections. <strong>The</strong> scaling factor (α) to scale the stretch (eq.5.81) to account<br />

for the heterogeneities in the MD that are not accounted for in the continuum RVE model<br />

is α = 0.861.<br />

Case 1: Extension along x1 axis with deformations in x2 and x3 constrained<br />

As the polymer model is strained along x1 axis, a fully three dimensional stress state<br />

develops within the material model due to the constraint on the Poisson’s contraction.<br />

Figures 6.22 to 6.24 are the stress-strain curves obtained from MD simulation (simula-<br />

tion results shown as data points) and the nanoscale damage model (NDM). Since, the<br />

dilatational strain in the local x1 direction is the only non-zero component <strong>of</strong> strain and<br />

promotes void nucleation and coalescence, we can characterize constants A1, A10 and A13<br />

using the constitutive equations listed above and the damage definition <strong>of</strong> D1. Given in<br />

Table 6.6 are the results to the analysis. Also, listed is the correlation coefficient <strong>of</strong> the<br />

129


linear fit. Shown in fig.(6.25) is the deformation <strong>of</strong> the atoms in the PNC in response to<br />

case 1 deformation. As can be seen from the deformed configuration <strong>of</strong> the composite we<br />

see micro-voids that have coalesced into larger damage entities.<br />

Table 6.6: Linear Regression and Analysis for polymer nanoclay composite (PNC) defor-<br />

mation Case 1<br />

Property Value (GPa) R 2<br />

C 0 11 1.155 0.989<br />

C 0 12 0.763 0.993<br />

C 0 13 0.585 0.976<br />

2A1 -0.99 0.999<br />

A10 -0.60 0.997<br />

A13 -0.47 0.999<br />

Figure 6.22: σ11 versus ɛ11 for polymer nanoclay composite (PNC) deformation (Case 1)<br />

130


Figure 6.23: σ22 versus ɛ11 for polymer nanoclay composite (PNC) deformation (Case 1)<br />

Figure 6.24: σ33 versus ɛ11 for polymer nanoclay composite (PNC) deformation (Case 1)<br />

131


Figure 6.25: Deformed and undeformed configurations <strong>of</strong> polymer nanoclay composite<br />

(PNC) deformation (Case 1)<br />

Case 2: Extension along x2 axis with deformations in x1 and x3 constrained<br />

Figures 6.26 to 6.28 are the stress-strain curves obtained from MD simulation (simulation<br />

results shown as data points) and the nanoscale damage model (NDM). In this case, the<br />

stretch in local 2-axis direction (also direction <strong>of</strong> applied maximum stretch) debonds the<br />

interface <strong>of</strong> the polymer nanocomposite. <strong>The</strong> constants A11, 2A5 and A17 are found using<br />

the constitutive equations listed above and the damage definition <strong>of</strong> D2. Given in Table<br />

6.7 are the results to the analysis. Shown in fig.(6.29) is the deformation <strong>of</strong> the atoms in<br />

the PNC in response to case 2 deformation with the interface debond clearly indicated.<br />

132


Table 6.7: Linear Regression and Analysis for polymer nanoclay composite (PNC) deformation<br />

Case 2<br />

Property Value (GPa) R 2<br />

C 0 12 0.708 0.979<br />

C 0 22 1.153 0.994<br />

C 0 23 0.563 0.960<br />

A11 -0.548 0.996<br />

2A5 -0.979 0.998<br />

A17 -0.435 0.998<br />

Figure 6.26: σ11 versus ɛ22 for polymer nanoclay composite (PNC) deformation (Case 2)<br />

Figure 6.27: σ22 versus ɛ22 for polymer nanoclay composite (PNC) deformation (Case 2)<br />

133


Figure 6.28: σ33 versus ɛ22 for polymer nanoclay composite (PNC) deformation (Case 2)<br />

Figure 6.29: Deformed and undeformed configurations <strong>of</strong> polymer nanoclay composite<br />

(PNC) deformation (Case 2)<br />

Case 3: Extension along x3 axis with deformation in x1 and x2 constrained<br />

Figures 6.30 to 6.32 are the stress-strain curves obtained from MD simulation and the<br />

nanoscale damage model (NDM). <strong>The</strong> stretch in local 3-axis direction (also direction <strong>of</strong><br />

134


maximum stretch) causes voids to appear in the polymer nanocomposite. <strong>The</strong> constants<br />

A15, A18 and A9 are found using the constitutive equations listed above and the damage<br />

definition <strong>of</strong> D1. Given in Table 6.8 are the coefficients obtained from results to the<br />

analysis, using linear regression.<br />

Table 6.8: Linear Regression and Analysis for polymer nanoclay composite (PNC) deformation<br />

Case 3<br />

Property Value (GPa) R 2<br />

C 0 13 0.733 0.973<br />

C 0 23 0.633 0.973<br />

C 0 33 0.648 0.993<br />

A15 -0.588 0.997<br />

A18 -0.504 0.995<br />

2A9 -0.528 0.994<br />

Figure 6.30: σ11 versus ɛ33 for polymer nanoclay composite (PNC) deformation (Case 3)<br />

Case 4: Simple shear in x2 − x3 plane<br />

In this section we characterize the stiffness and damage coefficients for simple shear in the<br />

x2 −x3 plane. As mentioned in Chapter 5, for the case <strong>of</strong> pure shear in small deformation<br />

the volumetric strain is zero. However for simple shear deformations in the local x1−x2 or<br />

135


Figure 6.31: σ22 versus ɛ33 for polymer nanoclay composite (PNC) deformation (Case 3)<br />

Figure 6.32: σ33 versus ɛ33 for polymer nanoclay composite (PNC) deformation (Case 3)<br />

x2−x3 planes, the RVE is susceptible to interface debond due to the non-zero stretch that<br />

occurs as shown in fig.(5.5). Since the influence function is a uniform radial atomistic<br />

interaction field as set by the Leonard Jones potential, the damage in shear can be<br />

characterized by simply tracking the maximum principle stretch in the RVE. Figure 6.33<br />

shows the result for MD and the NDM curve fit model through linear regression. <strong>The</strong><br />

values for the linear regression model are listed in Table 6.9.<br />

136


Table 6.9: Linear Regression and Analysis for polymer nanoclay composite (PNC) deformation<br />

Case 4<br />

Property Value (GPa) R 2<br />

C 0 44 0.189 0.941<br />

A22 -0.337 0.959<br />

Figure 6.33: σ23 versus γ23 for polymer nanoclay composite (PNC) deformation (Case 4)<br />

Case 5: Simple shear in x1 − x3 plane<br />

While the debond behavior seems to be well correlated to the simple shear deformation<br />

in case 4, the dilatational behavior is seen to have no affect on the stress response <strong>of</strong> the<br />

PNC as shown by the stress-strain curves in fig.(6.34). Hence, the damage coefficients,<br />

A20=A26=0. <strong>The</strong> undamaged stiffness coefficient C 0 55 is determined to be 38.851MPa.<br />

Case 6: Simple shear in x1 − x2 plane<br />

<strong>The</strong> arguments presented earlier for the debond growth in simple shear deformation in<br />

case 4 applies here. However, it is seen that a neat debond separation behavior at the<br />

interface is not observed; Instead, a distinct stick-slip stress-strain response with very<br />

little to no discernible linear elastic region is seen to occur. This observation seems<br />

to indicate that the polymer nano-composite is undergoing shear induced elasto-plastic<br />

137


Figure 6.34: σ13 versus γ13 for polymer nanoclay composite (PNC) deformation (Case 5)<br />

deformation at the macro-scale. It is seen that a similar behavior is observed for the<br />

stress-strain response for case 4 deformation, however the “stick” and “slip” <strong>of</strong> the stress<br />

versus increasing strain is not as pronounced as it is here. This could be attributed to<br />

the lower interaction length in the direction <strong>of</strong> applied shear strain. Further discussion<br />

on this observation is provided in Chapter 7. Figure 6.35 is the result for MD. Linear<br />

regression was not not performed for these results since a linear elastic region could not<br />

be discerned for the computation <strong>of</strong> the necessary stiffness and damage variables.<br />

6.3.2 Simulation results for pure Polypropylene without Nanoscale<br />

Reinforcement<br />

In an effort to quantify the improvements in mechanical properties <strong>of</strong> the polymer nanocom-<br />

posite and to clearly delineate the constituent damage modes for the pure polymer and<br />

polymer nanocomposite system, the pure thermoplastic polymer model was created using<br />

the steps described in section 3.3.1. We assume that the failure in the material model is<br />

due to the dilatational strains in the material that causes voids to nucleate and coalesce<br />

within the model when subject to deformation beyond a critical value. In the constitutive<br />

138


Figure 6.35: σ12 versus γ12 for polymer nanoclay composite (PNC) deformation (Case 6)<br />

modeling <strong>of</strong> pure polymer systems two cases <strong>of</strong> deformation needed to be studied using<br />

MD simulation. Following along the lines <strong>of</strong> the derivation for elastic-damage constitutive<br />

law for polymer nano composites, we propose a similar set <strong>of</strong> equations for the isotropic<br />

stiffness <strong>of</strong> the polypropylene MD model,<br />

Where,<br />

C12 = C 0 12 + C 1 12 (6.38)<br />

C22 = C 0 22 + C 1 22 (6.39)<br />

C23 = C 0 23 + C 1 23 (6.40)<br />

C66 = C 0 66 + C 1 66 (6.41)<br />

C 1 22 = 2B1D2 (6.42)<br />

C 1 12 = B2D2 (6.43)<br />

C 1 23 = B3D2 (6.44)<br />

C 1 66 = 2B4D2 (6.45)<br />

139


<strong>The</strong> damage coefficients Bi for the pure polymer are characterized in the next section.<br />

Note that in the following calculations the principle stretch determined from the defor-<br />

mation gradient is scaled by a factor <strong>of</strong> α = 0.941.<br />

Case 1: Extension along x2 axis with the deformations in x1 and x3 constrained<br />

Figures 6.36 to 6.38 are the stress-strain curves obtained from MD simulation (simulation<br />

results shown as data points) and the nanoscale damage model (NDM). In this case, the<br />

stretch in local 2-axis direction (also direction <strong>of</strong> maximum stretch) creates void(s) to<br />

nucleate with in the polymer system. <strong>The</strong> constants B1, B2 and B3 are found using the<br />

constitutive equations listed above and the damage definition <strong>of</strong> D2. Given in Table 6.10<br />

are the results to the analysis. Shown in fig.(6.39) is the deformation <strong>of</strong> polypropylene<br />

(PP) in response to case 1 deformation with the void formation clearly indicated.<br />

Table 6.10: Linear Regression and Analysis for pure polypropylene (PP) deformation<br />

Case 1<br />

Property Value (GPa) R 2<br />

C 0 12 0.447 0.981<br />

C 0 22 0.559 0.962<br />

C 0 23 0.498 0.92<br />

B1 -0.304 0.994<br />

2B2 -0.405 0.994<br />

B3 -0.355 0.997<br />

Case 2: Simple shear in x1 − x2 plane<br />

Table 6.11: Linear Regression and Analysis for pure polypropylene (PP) deformation<br />

Case 2<br />

Property Value (GPa) R 2<br />

C 0 66 0.109 0.941<br />

B4 -0.757 0.960<br />

140


Figure 6.36: σ11 versus ɛ22 for pure polypropylene (PP) deformation (Case 1)<br />

Figure 6.37: σ22 versus ɛ22 for pure polypropylene (PP) deformation (Case 1)<br />

For the case <strong>of</strong> simple shear in the x1 − x2 plane, we assume that the principal stretch<br />

causes the void to nucleate and grow along a direction inclined 45 ◦ to x2 (see, fig.6.40).<br />

Figure 6.41 shows the result for MD and the NDM curve fit model using linear regression.<br />

<strong>The</strong> values for the linear regression model are listed in Table 6.11. Also, shown in figure<br />

6.42 is the atomistic deformation in simple shear <strong>of</strong> the polymer MD model. It should<br />

be noted that the deformation map did not conclusively show a void opening on surface,<br />

141


Figure 6.38: σ33 versus ɛ22 for pure polypropylene (PP) deformation (Case 1)<br />

Figure 6.39: Deformed and undeformed configurations <strong>of</strong> pure polypropylene (PP) deformation<br />

(Case 1)<br />

indicating perhaps that the void nucleated inside the bulk polymer.<br />

142


Figure 6.40: Schematic <strong>of</strong> void formation in pure polypropylene RVE undergoing simple<br />

shear deformation in x1 − x2 plane with the principal stretch axis x ′ 2 oriented 45 ◦ to the<br />

local x2 axis<br />

Figure 6.41: σ12 versus γ12 for pure polypropylene (PP) deformation (Case 2)<br />

6.3.3 Comparison <strong>of</strong> Stiffness and Strength between Polymer<br />

Nanocomposite and Pure Polymer<br />

<strong>The</strong> Young’s moduli and failure strengths for the cases <strong>of</strong> extensional and shear deforma-<br />

tion can determined for the polymer nano-composite (PNC) and the pure polymer using<br />

the characterization data from atomistic MD simulations. Assuming orthotropic symme-<br />

143


Figure 6.42: Deformed and undeformed configurations <strong>of</strong> pure polypropylene (PP) deformation<br />

(Case 2)<br />

try in material properties for the PNC, we can use the following equation to determine<br />

the Young’s modulus in the local 2 direction <strong>of</strong> the nano-composite, E 0,PNC<br />

2 .<br />

E 0,PNC<br />

2<br />

= C0 11C 0 22C 0 33 + 2C0 23C 0 12C 0 13 − C0 0 2<br />

11C23 − C0 0 2<br />

22C13 − C0 0 2<br />

33C12 C0 22C 0 0 2<br />

33 − C23 (6.46)<br />

Where, C 0 11 = 1.155GPa, C 0 12 = 0.708GPa, C 0 13 = 0.585GPa, C 0 22 = 1.153GPa, C 0 23 =<br />

0.563GPa and C 0 33 = 0.648GPa. Similarly, assuming that the pure polymer has an<br />

isotropic response we can determine the Young’s modulus <strong>of</strong> the pure polymer E 0,PP<br />

is given as,<br />

E 0,PP =<br />

2<br />

C0 22 + C0 12C 0 0 2<br />

11 − 2C12 C0 22 + C0 12<br />

(6.47)<br />

Where, C 0 22 = 0.559GPa and C 0 12 = 0.447GPa. Table 6.12 is a summary <strong>of</strong> the Young’s<br />

modulus and stiffness enhancements <strong>of</strong> the polymer nano-composite relative to the pure<br />

144


polymer system as predicted by the MD algorithm.<br />

Table 6.12: Comparison <strong>of</strong> modulus <strong>of</strong> Polypropylene (PP) and Polymer Nano-composite<br />

(PNC) systems as predicted by Nano-scale Damage Model (NDM)<br />

C 0,PP<br />

22 (GPa) C 0,PNC<br />

22 (GPa) % increase E 0,PP (GPa) E 0,PNC<br />

2 (GPa) % increase<br />

0.559 1.153 106.26 0.162 0.567 250<br />

In order to compare the strengths <strong>of</strong> the polypropylene (σ PP ) and polymer nano-<br />

composite system (σPNC 2 ) in uni-axial extension along the local x2 axis we find the max-<br />

imum value <strong>of</strong> stress that can be obtained from the nano-scale damage model (NDM).<br />

From the results presented in the sections above we arrive at the results tabulated in<br />

Table 6.13.<br />

Table 6.13: Comparison <strong>of</strong> strength <strong>of</strong> Polypropylene (PP) and Polymer Nano-composite<br />

(PNC) systems as predicted by Nano-scale Damage Model (NDM)<br />

σ PP (MPa) σ PNC<br />

2 (MPa) % increase<br />

149.04 189.37 27.06<br />

<strong>The</strong> values for shear modulus enhancements in the x1−x2 plane for the polymer nano-<br />

composite could not be tabulated since a linear stress-strain response was not observed in<br />

the MD simulations. It is important to note that while the MD results <strong>of</strong> pure polymer<br />

indicates a clear drop in stress after damage is initiated (akin to brittle failure), the PNC<br />

response in shear (in both x1 − x2 and x2 − x3 plane) exhibits an elastic-plastic response<br />

indicating a prolonged yield type behavior.<br />

6.3.4 Summary<br />

In this section, the parameters for the proposed damage model from Chapter 5 were char-<br />

acterized using molecular dynamics simulations. It is observed from the linear regression<br />

curve fits to the MD simulations results that while the behavior <strong>of</strong> the polymer models<br />

145


can be well correlated to the damage model for the case <strong>of</strong> pure volumetric deformation,<br />

the shear behavior needs review and further research. However, the current model for<br />

shear failure is consistent with the assumptions that were made for the damage model<br />

(see Chapter 5). Interestingly, it is also observed that while the polypropylene had failed<br />

(clear drop in stress) at about 37% shear strain, the polymer nanocomposite exhibited a<br />

stick slip behavior indicating an elastic-plastic response in stress. This observation could<br />

possibly explain the strength enhancements that are seen in polymer nano-composites.<br />

146


Chapter 7<br />

Discussion<br />

In this dissertation two novel methods <strong>of</strong> modeling polymers and polymer nanocom-<br />

posites were proposed. While the complexities involved in modeling these materials are<br />

numerous, the proposed methods are able to determine an applicable multi-scale model<br />

<strong>of</strong> the material behavior by appealing to fundamental physics and the statistical nature<br />

<strong>of</strong> atomistic interactions at the nanoscale. In this chapter the significant contributions <strong>of</strong><br />

this research with respect to modeling <strong>of</strong> polymer nanocomposites is discussed, followed<br />

by a discussion on avenues for further research and development.<br />

7.1 Significance <strong>of</strong> Research<br />

Listed below are some <strong>of</strong> the salient features and possible applications <strong>of</strong> the modeling<br />

work in this dissertation.<br />

1. Coupling Meshless methods to Atomistic Simulation: In this work, we have ad-<br />

dressed and shown the application <strong>of</strong> meshless methods in coupling continuum to<br />

MD simulations. While traditional concurrent coupling methods have used finite<br />

element (FEA) based analysis <strong>of</strong> the continuum, in this work we have clearly shown<br />

the advantages <strong>of</strong> using a meshless methods. <strong>The</strong>y are,<br />

(a) <strong>The</strong> particles act as a natural transition from the continuum domain to atom-<br />

istic particles. As shown in Chapter 4, we have taken advantage <strong>of</strong> the particle<br />

description in GIMP to create the spatial and temporal connection to atomistic<br />

behavior in the coupling method.<br />

(b) <strong>The</strong> meshless method allows for finite deformation analysis <strong>of</strong> the MD system<br />

147


since it is not as susceptible to mesh entanglement issues as conventional finite<br />

element.<br />

(c) Atomistic interactions with crack growth in the continuum is now easier to<br />

study in coupled simulation due to the ease with which cracks can be modeled<br />

in GIMP, a method inherently independent <strong>of</strong> mesh.<br />

2. Concurrent Coupling <strong>of</strong> Polymers While there exists numerous techniques to con-<br />

currently couple atomistic models <strong>of</strong> regular lattice structures <strong>of</strong> solids to continuum<br />

analysis, a similar analysis methods for polymers are limited. <strong>The</strong> primary appli-<br />

cation <strong>of</strong> this analysis scheme is the ability to scale the MD simulation box to<br />

larger system sizes. Concurrent coupling, enables the user to model larger mate-<br />

rial systems while subscribing to macro-scale loading conditions without tagging on<br />

additional computational expenses. Further, since the modeling technique treats<br />

finite boundary effects more effectively than an equivalent MD simulation, it allows<br />

for the focussed study <strong>of</strong> the polymer in relation to macroscale loading conditions.<br />

3. Damage Modeling <strong>of</strong> Polymer Nanocomposites A simple small-strains phenomeno-<br />

logical model has been proposed to simulate damage behavior in nanocomposites.<br />

<strong>The</strong> damage model uses the exact non-bonded interactions at the atomistic scale to<br />

determine the thermodynamics <strong>of</strong> the polymer nanocomposite as a function <strong>of</strong> its<br />

mechanical behavior. Since the model is able to simulate the atomistic interaction<br />

in the nano composite without resorting to a full MD simulation, the model can<br />

be implemented in a continuum analysis to simulate deformational response in a<br />

nano-composite all the way to failure.<br />

4. Modeling Fiber reinforced Polymer/Nanocomposite matrix <strong>The</strong> primary objective<br />

<strong>of</strong> this research was to explain the strength enhancements in compression and shear<br />

properties <strong>of</strong> nano composites with very little addition <strong>of</strong> nanoclay to the polymer<br />

148


matrix. <strong>The</strong> modeling methodologies proposed in this work can be applied in a<br />

multi-scale analysis <strong>of</strong> the fiber reinforced composite with nano clay inclusions in<br />

the polymer matrix. <strong>The</strong> methodology provides engineers a design scale analysis<br />

and material evaluation theory for structural applications in which weight and cost<br />

cutting measures are <strong>of</strong> primary importance.<br />

7.2 Future Work<br />

While the proposed model is a step towards understanding and accurate simulation <strong>of</strong><br />

nanocomposites, there remains scope for further improvement and extended research<br />

opportunities. In this section specific algorithmic enhancements to the coupled simulation<br />

method and the multi-scale damage model are discussed. It is envisioned these topics<br />

would provide as a road map for future research and development.<br />

7.2.1 Coupled simulation <strong>of</strong> Fiber reinforced PNC<br />

From Chapter 6, it is evident that the coupled simulation technique is a viable alternative<br />

to a million atom MD simulation. As mentioned before, concurrent coupling can mitigate<br />

the finite edge effects that are normally unaccounted for in a periodic cell MD simulation.<br />

We have seen through coupled simulation that that the macroscale loading conditions<br />

creates a multi-axial stress field around the MD domain which in general is hard to<br />

simulate in a regular periodic purely atomistic simulation without artificially constraining<br />

the atomistic domain. Secondly, the use <strong>of</strong> coupled simulation also enables the user to<br />

scale the size <strong>of</strong> the system without the need for additional atoms. With the concurrent<br />

coupling scheme, the continuum domain can be considered as an extension <strong>of</strong> the atomistic<br />

region with the process zone (i.e., location in model where damage develops) restricted<br />

around the MD domain. However, certain challenges remain, some <strong>of</strong> which are listed<br />

below,<br />

149


1. Correct transfer <strong>of</strong> energy through the handshake zone is required to derive mean-<br />

ingful results out <strong>of</strong> coupled simulation. Dissipation <strong>of</strong> strain-energy at the interface<br />

due to incompatible material properties is the primary source <strong>of</strong> inaccuracy for all<br />

concurrent coupling algorithms. In the atomistic model <strong>of</strong> polymers these dissipa-<br />

tive mechanisms are enhanced due to the disordered material structure <strong>of</strong> polymer<br />

molecules and the random walk motion <strong>of</strong> the molecules at finite temperature. A<br />

one to one coupling <strong>of</strong> these systems are inherently difficult to implement. <strong>The</strong><br />

statistical coupling technique proposed in this work can help mitigate some <strong>of</strong> these<br />

issues. However, in the interest <strong>of</strong> improvement it is recommended that a purely pe-<br />

riodic MD simulation is first carried out to determine material properties and then<br />

using these material properties in the simulation <strong>of</strong> the bulk (continuum system).<br />

In this manner the material definition is consistent in both the bulk (continuum)<br />

and the local (MD domain) sense. It further ensures that the deformations and<br />

the associated stress field in the continuum zone will be compatible to those from<br />

atomistic simulation.<br />

2. An important application <strong>of</strong> the analysis scheme is to study behaviour <strong>of</strong> thermo-<br />

plastic composites in relation to macroscale loading. In this work we have seen the<br />

use <strong>of</strong> MD simulation to study the characteristic response <strong>of</strong> polymer (and poly-<br />

mer/nanoclay) atomistic model to various deformation states. In nanocomposites<br />

it is evident that the aspect ratio <strong>of</strong> the nanoclay platelets plays a central role in<br />

reinforcing the polymer system at the nano scale. Most nanoclay platelets have<br />

an aspect ratios in the range <strong>of</strong> about 50:1 to 100:1. In addition to the geometry<br />

distribution <strong>of</strong> the particles, weight fractions <strong>of</strong> nanocomposites are equally impor-<br />

tant in determining the mechanical behaviour <strong>of</strong> the composite. Hence, a realistic<br />

atomistic model will require significantly large number <strong>of</strong> simulation atoms and a<br />

relatively large sized atomistic domain. However it is envisioned that with coupled<br />

150


simulation the computational costs can be considerably reduced by modeling parts<br />

<strong>of</strong> composite with a smeared continuum model that is reliant on the deformation<br />

mechanics <strong>of</strong> the bulk rather than the details at the local scale.<br />

3. It is recommended that for a better description <strong>of</strong> the continuum response the im-<br />

plicit GIMP algorithm be used such that the time-integration errors per continuum<br />

update associated with an explicit time-stepping GIMP are avoided. An implicit<br />

analysis will also allow for higher time-step in the continuum. However, this will<br />

imply that the MD simulations will have to be run for longer periods (assuming<br />

that the MD time-step remains the same). Hence a parallel implementation <strong>of</strong> the<br />

GIMP/MD coupling is recommended to reduce the computational time per coupled<br />

simulation update.<br />

7.2.2 Multi-scale modeling <strong>of</strong> Polymer Nanocomposite<br />

For a given specimen <strong>of</strong> polymer nanocomposite, the nanoclay can exist in a exfoliated or<br />

intercalated morphology as shown in fig.7.1. In general, the configuration <strong>of</strong> the polymers<br />

around the nanoclay and the spatial distribution within the bulk polymer influences the<br />

property enhancements <strong>of</strong> the polymer nanocomposite. Hence, it will be beneficial to<br />

include real nano-scale system parameters such as gallery spacing, nanoclay aspect ratio,<br />

weight fraction <strong>of</strong> the embedded particle etc. in defining the damage entity. Further in<br />

Chapter 6, the nano-scale damage model (NDM) was used to make strength predictions<br />

for the composite and the pure polymer system. While the stiffness and strength enhance-<br />

ments in the PNC with respect to pure polymer is readily observed, the absolute values <strong>of</strong><br />

strength and stiffness predictions for the two systems (pure polymer and polymer nano-<br />

composite) do not compare well with experimental results 1 . <strong>The</strong>se anomalies could be<br />

attributed to system size <strong>of</strong> the MD and the particular choice <strong>of</strong> coarse grained atomistic<br />

1 modulus <strong>of</strong> polypropylene is between 0.8-2 GPa, while strength <strong>of</strong> polypropylene ˜40MPa<br />

151


Figure 7.1: Schematic <strong>of</strong> Nano-clay fully exfoliated and intercalated morphology in Polymer<br />

nano-composite<br />

model. In general, the purely amorphous polymer system will predict lower stiffness and<br />

higher strength since the crystalline nature <strong>of</strong> polypropylene is entirely omitted from the<br />

MD model. Further, while the use <strong>of</strong> coarse-graining expedites MD computations, the<br />

predictions <strong>of</strong> mechanical properties are in general not good, as can be observed from the<br />

results. Stiffness and strength are properties extensive to the MD system hence, simu-<br />

lations are required to determine an optimal MD system size which will be viable both<br />

computationally and physically as a representation <strong>of</strong> the actual polymer system. This<br />

exercise will also enable in the design <strong>of</strong> a reliable NDM for the PNC and pure polymer.<br />

152


Figure 7.2: Schematic <strong>of</strong> randomly oriented Nano-clay in global coordinates X,Y,Z and<br />

local coordinates x1, x2, x3<br />

7.2.3 A Generalized Three-Dimensional Damage Model<br />

In general, with most commonly employed manufacturing techniques it is difficult to<br />

fabricate well-aligned nano-clay nano composites. Hence, the final product will invariably<br />

have some nanoclay platelets that are not aligned to global XYZ directions <strong>of</strong> macro-scale<br />

loading (see, fig.7.2). <strong>The</strong> nanocomposite will hence be under the influence <strong>of</strong> a multi-<br />

axial state <strong>of</strong> stress. <strong>The</strong> RVE model can be extended for a general state <strong>of</strong> deformation<br />

in which the local axis <strong>of</strong> the nanoclay do not coincide with the loading directions <strong>of</strong> the<br />

macro-scale composite.<br />

Consider a point within the macro-composite that is in a general state <strong>of</strong> deformation,<br />

represented by the deformation gradient F ∗ . Let, R be an orthogonal transformation that<br />

153


converts the coordinates (X,Y,Z) in the global frame to local coordinates (x1, x2, x3) <strong>of</strong><br />

the RVE, or mathematically: R : R · X ↦−→ x.<br />

In general, R will have the following properties,<br />

R −1 = R T and R T R = I (7.1)<br />

Where the superscript T denotes transpose and I is the identity tensor. We can then<br />

define the deformation gradient F in the local coordinate frame as follows,<br />

F = RF ∗ R T<br />

or, in indicial notation, Fij = rimrjnF ∗ mn<br />

(7.2)<br />

We then employ the following procedure to determine the effective stretch λ along the<br />

direction <strong>of</strong> vector n={n1, n2, n3} (n is the normal vector to the damage surface, see<br />

Chapter 5). If N is the principle direction vector, λ ∗ the principle stretch values <strong>of</strong> the<br />

deformation gradient F and C is the right deformation tensor (Malvern, 1969), then we<br />

obtain the complete set <strong>of</strong> principle values (λ ∗ 1,λ ∗ 2,λ ∗ 3) solving the following eigen value<br />

problem,<br />

(C − λ ∗2 I) · N = 0 (7.3)<br />

Let Ni be the principle direction vector that corresponds to the eigen value λ ∗ i . Further,<br />

let the components <strong>of</strong> Ni in the local coordinate be given as, Ni =<br />

<br />

N (1)<br />

i , N (2)<br />

i , N (3)<br />

<br />

i .<br />

<strong>The</strong>n the stretch λ can be determined for a general state <strong>of</strong> deformation as (indicial<br />

summation implied),<br />

λ = λiNi · n = λ ∗ 1N1 · n + λ ∗ 2N2 · n + λ ∗ 3N3 · n (7.4)<br />

154


References<br />

Abraham, F., Broughton, J., Bernstein, N., & Kaxiras, E. (1999). Concurrent coupling<br />

<strong>of</strong> length scales: methodology and application. Physical Review B, 60 (4), 2391–2403.<br />

Abraham, F. F., Walkup, R., Gao, H., Duchaineau, M., Diaz De La Rubia, T., & Seager,<br />

M. (2002). Simulating materials failure by using up to one billion atoms and the worlds<br />

fastest computer: Brittle fracture. Proceedings <strong>of</strong> the National Academy <strong>of</strong> Sciences<br />

(U.S.A.), 99 (9), 5783–5787.<br />

Abramson, A. R., Tien, C., & Majumdar, A. (2002). Interface and strain effects on the<br />

thermal conductivity <strong>of</strong> heterostructures: A molecular dynamics study. Journal <strong>of</strong> Heat<br />

Transfer, 124, 963–970.<br />

Adkins, J. E. (1959). Symmetry relations for orthotropic and transversely isotropic materials.<br />

Archive for Rational Mechanics and Analysis, 4 (1), 193–213.<br />

Alivisatos, P., Barbara, P. F., Castleman, W., Chang, J., Dixon, D. A., Klein, M. L.,<br />

McLendon, G. L., Miller, J. S., Ratner, M. A., Rossky, P. J., Stupp, S. I., & Thompson,<br />

M. E. (1998). From molecules to materials: Current trends and future directions.<br />

Advanced Materials, 10 (16), 1297–1336.<br />

Allen, M. P. & Tildesley, D. J. (1987). Computer Simulations <strong>of</strong> Liquids. Oxford, U.K.:<br />

Clarendon Press.<br />

Ayton, G., Bardenhagen, S. G., McMurtry, P., Sulsky, D., & Voth, G. A. (2001). Interfacing<br />

continuum and molecular dynamics: An application to lipid bilayers. Journal<br />

<strong>of</strong> Chemical Physics, 114 (15), 6913–6924.<br />

Bardenhagen, S. G. & Kober, E. (2004). <strong>The</strong> generalized interpolation material point<br />

method. Computer Modeling in Engineering & Sciences, 5, 477–496.<br />

Bathe, K. J. (1982). Finite Element Procedures in Engineering Analysis. New Jersey,<br />

U.S.A.: Prentice Hall.<br />

Berber, S., Kwon, Y., & Tomnek, D. (2000). Unusually high thermal conductivity <strong>of</strong><br />

carbon nanotubes. Physics Review Letters, 84 (20), 4613–4616.<br />

Berta, M., Lindsay, C., Pans, G., & Camino, G. (2006). Effect <strong>of</strong> chemical structure on<br />

combustion and thermal behaviour <strong>of</strong> polyurethane elastomer layered silicate nanocomposites.<br />

Polymer Degradation and Stability, 91 (5).<br />

Binder, K. (1995). Monte Carlo and molecular dynamics simulations in polymer sciences.<br />

New York, U.S.A.: Oxford <strong>University</strong> Press.<br />

Buehler, M. J. & Wong, S. Y. (2007). Entropic elasticity controls nanomechanics <strong>of</strong> single<br />

tropocollagen molecules. Biophysical Journal, 93, 37–43.<br />

155


Buryachenko, V., Roy, A., L., K., Anderson, K., & Chellapilla, S. (2005). Multi-scale<br />

mechanics <strong>of</strong> nanocomposites including interface: Experimental and numerical investigation.<br />

Composites Science and Technology, 65, 2435–2465.<br />

Charati, S. G. & Stern, S. A. (1998). Molecular-dynamics simulation <strong>of</strong> diffusion <strong>of</strong> small<br />

penetrants in polymers. Macromolecules, 31 (16), 5529–5535.<br />

Chen, L., Wong, S.-C., Liu, T., Lu, X., & He, C. (2004). Deformation mechanisms<br />

<strong>of</strong> nanoclay-reinforced maleic anhydride-modified polypropylene. Journal <strong>of</strong> Polymer<br />

Science: Part B: Polymer Physics, 42, 2759–2768.<br />

Courant, R., Friedrichs, K., & Lewy, H. (1928). ber die partiellen differenzengleichungen<br />

der mathematischen physik. Mathematische Annalen, 100 (1), 32–74.<br />

Danielsson, M., Parks, D., & Boyce, M. (2002). Three-dimensional micromechanical<br />

modeling <strong>of</strong> voided polymeric materials. Journal <strong>of</strong> the Mechanics and Physics <strong>of</strong><br />

Solids, 50, 351–379.<br />

Daphalapurkar, N. P., Lu, H., Coker, D., & Komanduri, R. (2007). Simulation <strong>of</strong> dynamic<br />

crack growth using the generalized interpolation material point (gimp) method.<br />

International Journal <strong>of</strong> Fracture, 143, 79–102.<br />

Dong, Y. & Bhattacharyya, D. (2008). Effects <strong>of</strong> clay type, clay/compatibiliser content<br />

and matrix viscosity on the mechanical properties <strong>of</strong> polypropylene/organoclay<br />

nanocomposites. Composites: Part A, 39, 1177–1191.<br />

Drozdov, A. D., Jensen, E. A., & Christiansen, J. d. C. (2010). Nonlinear time-dependent<br />

response <strong>of</strong> polypropylene/nanoclay melts: Experiments and modeling. Computational<br />

Materials Science, 47, 807–816.<br />

Duflot, M., Ni, K., & Hu, P. (2010). Extended finite element methods based twoparameter<br />

fatigue crack growth prediction. AIAA paper, AIAA–2010–2535.<br />

Erpenbeck, J. J. (1984). Shear viscosity <strong>of</strong> the hard-sphere fluid via non-equilibrium<br />

molecular dynamics. Physics Review Letters, 52 (15), 1333–1335.<br />

Frankland, S. J. V., Caglar, A., Brenner, D. W., & Griebel, M. (2002). Molecular<br />

simulation <strong>of</strong> the influence <strong>of</strong> chemical cross-links on the shear strength <strong>of</strong> carbon<br />

nanotube-polymer interfaces. Journal <strong>of</strong> Physical Chemistry, 43 (3), 285–293.<br />

Galgalia, G., Agarwal, S., & Lelea, A. (2004). Effect <strong>of</strong> clay orientation on the tensile<br />

modulus <strong>of</strong> polypropylene nanoclay composites. Polymer, 45, 6059–6069.<br />

Gilmanov, A. & Acharya, S. (2008). A computational strategy for simulating heat transfer<br />

and flow past deformable objects. International Journal <strong>of</strong> Heat and Mass Transfer,<br />

51, 4415–4426.<br />

156


Golebiewski, J. & Galeski, A. (2007). <strong>The</strong>rmal stability <strong>of</strong> nanoclay polypropylene composites<br />

by simultaneous dsc and tga. Composites Science and Technology, 67, 3442–<br />

3447.<br />

Guilkey, J., Harman, T., Xiz, A., Kashiwa, B., & McMurtry, P. (2003). An eulerianlagrangian<br />

approach for large deformation fluid structure interactions problems, part<br />

i: Algorithm development. In Proceedings <strong>of</strong> the Second International Conference on<br />

Fluid Structure Interaction, Cadiz, Spain.<br />

Guo, Y. & <strong>Nair</strong>n, J. A. (2004). Calculation <strong>of</strong> j-integral and stress intensity factors using<br />

the material point method. Computer Modeling in Engineering & Sciences, 6, 295–308.<br />

Hbaieb, K., Wang, Q., Chia, Y., & Cotterell, B. (2007). Modelling stiffness <strong>of</strong> polymer/clay<br />

nanocomposites. Polymer, 48, 901–909.<br />

Heffelfinger, G. S. & van Swol, F. (1994). Diffusion in lennard-jones fluids using dual<br />

control volume grand canonical molecular dynamics simulation (dc-gcmd). Journal <strong>of</strong><br />

Chemical Physics, 100 (10), 7548–7553.<br />

Hess, B. (2002). Determining the shear viscosity <strong>of</strong> model liquids from molecular dynamics<br />

simulations. Physics Review Letters, 116 (1), 209–213.<br />

Hockney, R. W. (1970). <strong>The</strong> potential calculation and some applications. Methods in<br />

Computational Physics, 9, 136–211.<br />

Jones, R. M. (1998). Mechanics <strong>of</strong> Composite Materials. U.S.A.: Taylor & Francis, Inc.<br />

Katti, D. R., Schmidt, S. R., Ghosh, P., & Katti, K. S. (2005). Modeling the response <strong>of</strong><br />

pyrophyllite interlayer to applied stress using steered molecular dynamics. Clays and<br />

Clay Minerals, 53 (2), 171–178.<br />

Katti, K. S., Sikdar, D., Katti, D. R., Ghosh, P., & Verma, D. (2006). Molecular interactions<br />

in intercalated organically modified clay and claypolycaprolactam nanocomposites:<br />

Experiments and modeling. Polymer, 47, 403–414.<br />

Khare, R., Mielke, S., Paci, J., Zhang, S., Schatz, G., & Belytschko, T. (2007). Two<br />

quantum mechanical/molecular mechanical coupling schemes appropriate for fracture<br />

mechanics studies. AIAA paper.<br />

Kojima, Y., Usuki, A., Kawasumi, M., Okada, A., Fukushima, Y., Kurauchi, T., & Kamigaito,<br />

O. (1993). Mechanical properties <strong>of</strong> nylon 6-clay hybrid. Journal <strong>of</strong> Materials<br />

Research, 8 (5), 1185–1189.<br />

Kollman, P. (1993). Free energy calculations: Applications to chemical and biological<br />

phenomena. Chemical Reviews, 93, 2395–2417.<br />

Krishnamoorti, R., Vaia, R. A., & Giannelis, E. P. (1996). Structure and dynamics <strong>of</strong><br />

polymer-layered silicate nanocomposites. Chemistry <strong>of</strong> Materials, 8, 1728–1734.<br />

157


Krueger, R. (2004). Virtual crack closure technique: History, approach and applications.<br />

Applied Mechanics Reviews, 57, 109–143.<br />

Kuppa, V., Menakanit, S., Krishnamoorti, R., & Manias, E. (2003). Simulation insights<br />

on the structure <strong>of</strong> nanoscopically confined poly(ethylene oxide). Journal <strong>of</strong> Polymer<br />

Science: Part B: Polymer Physics, 41, 3285–3298.<br />

Leach, A. R. (1996). Molecular Modeling: Principles and Applications. U.K.: Addison<br />

Wesley Longman Ltd.<br />

Lei, Y., Wu, Q., Clemons, C. M., Yao, F., & Xu, Y. (2007). Influence <strong>of</strong> nanoclay on<br />

properties <strong>of</strong> hdpe/wood composites. Journal <strong>of</strong> Applied Polymer Science, 106, 3958–<br />

3966.<br />

Liew, K. M., He, X. Q., & Wong, C. H. (2004). On the study <strong>of</strong> elastic and plastic properties<br />

<strong>of</strong> multi-walled carbon nanotubes under axial tension using molecular dynamics<br />

simulation. Acta Materialia, 52, 2521–2527.<br />

Liu, W. K., Karpov, E. G., & Park, H. S. (2006). Nano Mechanics and Materials: <strong>The</strong>ory,<br />

Multiscale Methods and Applications. U.K.: J. Wiley & Sons Ltd.<br />

Looa, L. S. & Gleason, K. K. (2004). Investigation <strong>of</strong> polymer and nanoclay orientation<br />

distribution in nylon 6/montmorillonite nanocomposite. Polymer, 45, 5933–5939.<br />

Lu, H., Daphalapurkar, N., Wang, B., Roy, S., & Komanduri, R. (2006). Multiscale simulation<br />

from atomistic to continuum coupling molecular dynamics (md) with material<br />

point method (mpm). Philosophical Magazine, 86 (20), 2971–2994.<br />

Ma, J., Lu, H., & Komanduri, R. (2006). Structured mesh refinement in generalized interpolation<br />

material point (gimp) method for simulation <strong>of</strong> dynamic problems. Computer<br />

Modeling in Engineering & Sciences, 12 (3), 213–227.<br />

Ma, J., Lu, H., Wang, B., Hornung, R., Wissink, A., & Komanduri, R. (2006). Multiscale<br />

simulation using generalized interpolation material point (gimp) method and molecular<br />

dynamics (md). Computer Modeling in Engineering & Sciences, 14 (2), 101–117.<br />

Ma, J., Lu, H., Wang, B., Roy, S., Hornung, R., Wissink, A., & Komanduri, R. (2005).<br />

Multiscale simulations using generalized interpolation material point (gimp) method<br />

and samrai parallel processing. Computer Modeling in Engineering & Sciences, 8 (2),<br />

135–152.<br />

Malvern, L. E. (1969). Introduction to the Mechanics <strong>of</strong> a Continuous Medium. New<br />

Jersey, U.S.A.: Prentice-Hall.<br />

Mani, G., Fan, Q., Ugbolue, S. C., & Yang, Y. (2005). Morphological studies <strong>of</strong> polypropylenenanoclay<br />

composites. Journal <strong>of</strong> Applied Polymer Science, 97, 218–226.<br />

158


Maruyama, S. (2002). A molecular dynamics simulation <strong>of</strong> heat conduction in finite<br />

length swnts. Physica B, 323, 1895–1898.<br />

Mavrantzas, V. G. & <strong>The</strong>odorou, D. N. (1998). Atomistic simulation <strong>of</strong> polymer melt<br />

elasticity: Calculation <strong>of</strong> the free energy <strong>of</strong> an oriented polymer melt. Macromolecules,<br />

31, 6310–6332.<br />

Minisini, B. & Tsobnang, F. (2005). Molecular dynamics study <strong>of</strong> specific interactions in<br />

grafted polypropylene organomodified clay nanocomposite. Composites: Part A, 36,<br />

539–544.<br />

Muller-Plathe, F. (1991). Calculation <strong>of</strong> the free energy for gas absorption in amorphous<br />

polypropylene. Macromolecules, 24, 6475–6479.<br />

<strong>Nair</strong>, A. & Roy, S. (2009). Implicit time integration in the generalized interpolation<br />

material point method for finite deformation hyper elasticity. Mechanics <strong>of</strong> Advanced<br />

Materials and Structures, In review.<br />

<strong>Nair</strong>, S. V., Goettler, L. A., & Lysek, B. A. (2002). Toughness <strong>of</strong> nanoscale and multiscale<br />

polyamide-6,6 composites. Polymer Engineering and Science, 42 (9), 1872–1882.<br />

<strong>Nair</strong>n, J. A. (2003). Material point method calculations with explicit cracks. Computer<br />

Modeling in Engineering & Sciences, 4, 694–664.<br />

Ogata, S., Lidorikis, E., Shimojo, F., Nakano, A., Vashishta, P., & Kalia, R. (2001). Hybrid<br />

finite-element/molecular-dynamics/electronic-density-functional approach to materials<br />

simulations on parallel computers. Computer Physics Communications, 138,<br />

143–154.<br />

Pant, P. V. K. & Boyd, R. H. (1993). Molecular-dynamics simulation <strong>of</strong> diffusion <strong>of</strong> small<br />

penetrants in polymers. Macromolecules, 26 (4), 679–686.<br />

Park, J. H. & Jana, S. C. (2003). Mechanism <strong>of</strong> exfoliation <strong>of</strong> nanoclay particles in<br />

epoxy-clay nanocomposites. Macromolecules, 36, 2758–2768.<br />

Plimpton, S. J. (1995). Fast parallel algorithms for short-range molecular dynamics.<br />

Journal <strong>of</strong> Computational Physics, 117, 1–19.<br />

Qian, D., Liu, W. K., & Zheng, Q. (2008). Concurrent quantum/continuum coupling<br />

analysis <strong>of</strong> nanostructures. Computational Methods in Applied Mechanics and Engineering.,<br />

197, 3291–3323.<br />

Reddy, J. N. (2006). An Introduction to the Finite Element Method. New York, U.S.A.:<br />

Mc-Graw Hill Co.<br />

Rey-Castro, C. & Vega, L. F. (2006). Transport properties <strong>of</strong> the ionic liquid 1-ethyl-3methylimidazolium<br />

chloride from equilibrium molecular dynamics simulation. the effect<br />

<strong>of</strong> temperature. Journal <strong>of</strong> the Physical Chemistry B, 110, 14426–14435.<br />

159


Riddick, J., Frankland, S., & Gates, T. (2006a). Modeling <strong>of</strong> interface behavior in carbon<br />

nanotube composites. AIAA paper, AIAA–2006–1677.<br />

Riddick, J., Frankland, S., & Gates, T. (2006b). Multiscale analysis <strong>of</strong> delamination <strong>of</strong><br />

carbon fiber-epoxy laminates with carbon nanotubes. AIAA paper, AIAA–2006–1676.<br />

Rountree, C. L., Kalia, R. K., Lidorikis, E., Nakano, A. Brutzel, L., & Vashishta, P.<br />

(2002). Crack propagation and fracture in ceramic films-million atom molecular dynamics<br />

simulations on parallel computers. Annual Review <strong>of</strong> Materials Research, 32,<br />

377–400.<br />

Roy, S., Hussain, F., Narasimhan, K., Vengadassalam, K., & Lu, H. (2007). Eglass/polypropylene<br />

nanocomposites: Manufacture, characterization, thermal and mechanical<br />

properties. Polymer & Polymer Composites, 15 (2), 91–102.<br />

Saether, E., Yamakov, V., & Glaessgen, E. H. (2009). An embedded statistical method<br />

for coupling molecular dynamics and finite element analyses. International Journal for<br />

Numerical Methods in Engineering, 78, 1292–1391.<br />

Schroder, J. & Neff, P. (2003). Invariant formulation <strong>of</strong> hyperelastic transverse isotropy<br />

based on polyconvex free energy functions. International Journal <strong>of</strong> Solids and Structures,<br />

40, 401–445.<br />

Sheng, N., Boyce, M., Parks, D., Rutledge, G., Abes, J., & Cohen, R. (2004). Multiscale<br />

micromechanical modeling <strong>of</strong> polymer/clay nanocomposites and the effective clay<br />

particle. Polymer, 45, 487–506.<br />

Sinsawat, A., Anderson, K. L., Vaia, R. A., & Farmer, B. L. (2003). Influence <strong>of</strong> polymer<br />

matrix composition and architecture on polymer nanocomposite formation: Coarsegrained<br />

molecular dynamics simulation. Journal <strong>of</strong> Polymer Science: Part B: Polymer<br />

Physics, 41, 3272–3284.<br />

Smith, J. S., Bedrov, D., & Smith, G. D. (2003). A molecular dynamics simulation study<br />

<strong>of</strong> nanoparticle interactions in a model polymer-nanoparticle composite. Composites<br />

Science and Technology, 63, 1599–1605.<br />

Spencer, A. J. M. (1971). <strong>The</strong>ory <strong>of</strong> Invariants in Continuum Physics. New York, U.S.A.:<br />

Academic Press.<br />

Srinath, G. & Gnanamoorthy, R. (2005). Effect <strong>of</strong> nanoclay reinforcement on tensile and<br />

tribo behaviour <strong>of</strong> nylon 6. Journal <strong>of</strong> Materials science, 40, 2897–2901.<br />

Subramaniyan, A. K. & Sun, C. T. (2006a). Enhancing compressive strength <strong>of</strong> unidirectional<br />

polymeric composites using nanoclay. Composites: Part A, 37, 2257–2268.<br />

Subramaniyan, A. K. & Sun, C. T. (2006b). Enhancing compressive strength <strong>of</strong> unidirectional<br />

polymeric composites using nanoclay. Composites: Part A, 37, 2257–2268.<br />

160


Sulsky, D., Chen, Z., & Schreyer, H. (1994). A particle method for history-dependent<br />

materials. Computational Methods in Applied Mechanics and Engineering, 118, 179–<br />

196.<br />

Sulsky, D. & Schreyer, H. (1996). Axisymmetric form <strong>of</strong> the material point method<br />

with applications to upsetting and taylor impact problems. Computational Methods in<br />

Applied Mechanics and Engineering, 139, 409–429.<br />

Sulsky, D., Zhou, S., & Schreyer, H. (1995). Application <strong>of</strong> particle-in-cell method to<br />

solid mechanics. Computer Physics Communications, 87, 236–252.<br />

Sun, Q., Schork, F. J., & Deng, Y. (2007). Water-based polymer/clay nanocomposite<br />

suspension for improving water and moisture barrier in coating. Composites Science<br />

and Technologye, 67 (9), 1823–1829.<br />

Suter, J. L., Coveney, P. V., Greenwell, C. H., & Thyveetil, M. (1995). Monte carlo<br />

simulation <strong>of</strong> interlayer molecular structure in swelling clay minerals 1. methodology.<br />

Clays and Clay Minerals, 43 (3), 285–293.<br />

Suter, J. L., Coveney, P. V., Greenwell, C. H., & Thyveetil, M. (2007). Large-scale<br />

molecular dynamics study <strong>of</strong> montmorillonite clay: Emergence <strong>of</strong> undulatory fluctuations<br />

and determination <strong>of</strong> material properties. Journal <strong>of</strong> Physical Chemistry, 111,<br />

8248–8259.<br />

Swope, W. C., Anderson, H. C., Berens, P. H., & Wilson, K. R. (1982). A computer<br />

simulation method for the calculation <strong>of</strong> equilibrium constants for the formation <strong>of</strong><br />

physical clusters <strong>of</strong> molecules: application to small water clusters. Journal <strong>of</strong> Chemical<br />

Physics, 76, 637–649.<br />

Talreja, R. (1990). Internal variable damage mechanics <strong>of</strong> composite materials. Yielding,<br />

Damage and Failure <strong>of</strong> Anisotropic Solids, EGFS (Edited by J. P. Boehler), 509–533.<br />

Tan, H. & <strong>Nair</strong>n, J. A. (2002). Hierarchical, adaptive, material point method for dynamic<br />

energy release rate calculations. Computer Methods in Applied Mechanics and<br />

Engineering, 191, 2095–2109.<br />

Tang, Y. W., Szalai, I., & Chan, K. (2001). Transport properties <strong>of</strong> the ionic liquid 1ethyl-3-methylimidazolium<br />

chloride from equilibrium molecular dynamics simulation.<br />

the effect <strong>of</strong> temperature. Journal <strong>of</strong> the Physical Chemistry A, 105, 9616–9623.<br />

Timmerman, J. F., Hayes, B. S., & Seferis, J. C. (2002). Nanoclay reinforcement effects<br />

on the cryogenic microcracking <strong>of</strong> carbon fiber/epoxy composites. Composites Science<br />

and Technology, 62, 1249–1258.<br />

Toth, R., Coslanich, A., Ferrone, M., Fermeglia, M., Pricl, S., Miertus, S., & Chiellini,<br />

E. (2004). Computer simulation <strong>of</strong> polypropylene/organoclay nanocomposites: characterization<br />

<strong>of</strong> atomic scale structure and prediction <strong>of</strong> binding energy. Polymer, 45,<br />

8075–8083.<br />

161


Truesdell, C. T. & Noll, W. (1992). <strong>The</strong> Non-linear Field <strong>The</strong>ories <strong>of</strong> Mechanics. Berlin,<br />

Germany: Springer-Verlag.<br />

Tsai, J. & Sun, C. T. (2004). Effect <strong>of</strong> platelet dispersion on the load transfer efficiency<br />

in nanoclay composites. Journal <strong>of</strong> Composite Materials, 38 (7), 901–909.<br />

Vaia, R. A. & Giannelis, E. P. (1997). Lattice model <strong>of</strong> polymer melt intercalation in<br />

organically-modified layered silicates. Macromolecules, 30, 7990–7999.<br />

Valavala, P., Clancy, T., Odegard, G., & Gates, T. (2007). Nonlinear multiscale modeling<br />

<strong>of</strong> polymer materials. International Journal <strong>of</strong> Solids and Structures, 44, 1161–1179.<br />

Vashishta, P., Nakano, A., Kalia, R. K., & Ebbsjio, I. (1996). Crack propagation and<br />

fracture in ceramic films-million atom molecular dynamics simulations on parallel computers.<br />

Materials Science and Engineering B, 37, 56–71.<br />

Verlet, L. (1967). Computer experiments on classical fluids. i. thermodynamical prperties<br />

<strong>of</strong> lennard-jones molecules. Physical Review Letters, 165, 201–214.<br />

Volz, S. G. & Chen, G. (1999). Molecular dynamics simulation <strong>of</strong> thermal conductivity<br />

<strong>of</strong> silicon nanowires. Applied Physics Letters, 75 (14), 2056–2058.<br />

Wagner, G. & Liu, W. (2003). Coupling <strong>of</strong> atomistic and continuum simulations using a<br />

bridging scale decomposition. Journal <strong>of</strong> Computational Physics, 190, 249–274.<br />

Wei, C., Srivastava, D., & Cho, K. (2002). <strong>The</strong>rmal expansion and diffusion coefficients<br />

<strong>of</strong> carbon nanotube-polymer composites. Nano Letters, 2 (6), 647–650.<br />

Wieckowski, Z. (2004). <strong>The</strong> material point method in large strain engineering problems.<br />

Computer Methods in Applied Mechanics and Engineering, 193, 4417–4438.<br />

Xiao, S. & Belytschko, T. (2004). A bridging domain method for coupling continua with<br />

molecular dynamics. Computer Methods in Applied Mechanics and Engineering, 193,<br />

1645–1669.<br />

Yalcin, B., Ergungor, Z., Konishi, Y., Cakmak, M., & Batur, C. (2008). Molecular origins<br />

<strong>of</strong> toughening mechanism in uniaxially stretched nylon 6 films with clay nanoparticles.<br />

Polymer, 49, 1635–1650.<br />

Yang, I.-K. & Tsai, P.-H. (2006). Intercalation and viscoelasticity <strong>of</strong> poly(ether-blockamide)<br />

copolymer/montmorillonite nanocomposites: Effect <strong>of</strong> surfactant. Polymer, 47,<br />

5131–5140.<br />

Yang, J.-L., Zhang, Z., Schlarb, A. K., & Friedrich, K. (2006). On the characterization<br />

<strong>of</strong> tensile creep resistance <strong>of</strong> polyamide 66 nanocomposites. part i. experimental results<br />

and general discussions. Polymer, 47, 2791–2801.<br />

162


York, A. R., Sulsky, D., & Schreyer, H. L. (2000). Fluid-membrane interaction based on<br />

the material point method. International Journal for Numerical Methods in Engineering,<br />

48, 901–924.<br />

Yuan, Q., Deshmanea, C., Pesacreta, T., & Misra, R. (2008). Nanoparticle effects on<br />

spherulitic structure and phase formation in polypropylene crystallized at moderately<br />

elevated pressures:the influence on fracture resistance. Materials Science and Engineering<br />

A, 480, 181–188.<br />

Yuan, Q. & Misra, R. (2006). Impact fracture behavior <strong>of</strong> clayreinforced polypropylene<br />

nanocomposites. Polymer, 47, 4421–4433.<br />

Yuge, T., Ito, N., & Shimizu, A. (2005). Nonequilibrium molecular dynamics simulation<br />

<strong>of</strong> electric conduction. Journal <strong>of</strong> the Physical Society <strong>of</strong> Japan, 74 (7), 1895–1898.<br />

Zeng, Q. H., Yua, A. B., & Lu, G. (2008). Multiscale modeling and simulation <strong>of</strong> polymer<br />

nanocomposites. Progress in Polymer Science, 33, 191–269.<br />

Zhang, D., Zou, Q., VanderHeyden, W. B., & Ma, X. (2008). Material point method<br />

applied to multiphase flows. Journal <strong>of</strong> Computational Physics, 227, 3159–3173.<br />

Zhang, H., Wang, K., & Chen, Z. (2009). Material point method for dynamic analysis <strong>of</strong><br />

saturated porous media under external contact/impact <strong>of</strong> solid bodies. Computational<br />

Methods in Applied Mechanics and Engineering, 198, 1456–1472.<br />

Zhao, H. & Aluru, N. R. (2008). Molecular dynamics simulation <strong>of</strong> bulk silicon under<br />

strain. Interaction and Multiscale Mechanics, 1 (2), 303–315.<br />

Zhou, M. (2003). A new look at the atomic level virial stress: on continuum-molecular<br />

system equivalence. Proceedings <strong>of</strong> the Royal Society A, 459, 2347–2392.<br />

Zhu, L. & Narh, K. (2004). Numerical simulation <strong>of</strong> the tensile modulus <strong>of</strong> nanoclayfilled<br />

polymer composites. Journal <strong>of</strong> Polymer Science: Part B: Polymer Physics,<br />

42 (7), 2391–2406.<br />

163


5),<br />

Appendix A<br />

Derivation <strong>of</strong> Stress-Strain law based on Dilatational and<br />

Deviatoric response <strong>of</strong> a material<br />

Following the Claussius-Duhem inequality (Malvern, 1969) we can know (see, Chapter<br />

σij = ρ ∂ψ<br />

ɛij<br />

(A.1)<br />

Where, ρ is the current density <strong>of</strong> the system, ψ is the specific Helmholtz free energy.<br />

For the small strain tensor we can define the volumetric (dilatational) (ɛii) and deviatoric<br />

(¯ɛij) strains as follows,<br />

ɛii = tr(ɛ) = ɛ11 + ɛ22 + ɛ33 (A.2)<br />

¯ɛij = ɛij − 1<br />

3 δijɛkk<br />

(A.3)<br />

Where, i, j, k are indicial variables and repeated indices imply summation (eq.A.2),<br />

δij is the Kronecker delta function. Assuming that the energy can be partitioned in to a<br />

pure volumetric and deviatoric parts we have,<br />

<br />

∂ψ ∂ɛkk<br />

σij = ρ<br />

∂ɛkk ∂ɛij<br />

+ ∂ψ<br />

∂¯ɛlm<br />

Differenciating eq.(A.3) with respect to ɛij we get,<br />

∂¯ɛlm<br />

∂ɛij<br />

= ɛlm<br />

ɛij<br />

− 1<br />

3 δlm<br />

∂ɛkk<br />

∂ɛij<br />

<br />

∂¯ɛlm<br />

∂ɛij<br />

= δilδjm − 1<br />

3 δijδlm<br />

Where the identity ∂ɛkk<br />

∂ɛij = δij has been used. Substituting eq.(A.5) in eq.(A.4),<br />

164<br />

(A.4)<br />

(A.5)


∂ψ<br />

σij = ρ δij +<br />

∂ɛkk<br />

∂ψ<br />

∂¯ɛlm<br />

<br />

∂ψ<br />

= ρ δij +<br />

∂ɛkk<br />

∂ψ<br />

∂¯ɛij<br />

<br />

δilδjm − 1<br />

3 δijδlm<br />

<br />

− 1<br />

3 δij<br />

<br />

∂ψ<br />

∂¯ɛll<br />

(A.6)<br />

(A.7)<br />

We see that the last term in the above equation is zero since, trace(¯ɛ) = 0. Hence the<br />

constitutive law for the material can be derived as,<br />

<br />

∂ψ<br />

σij = ρ δij +<br />

∂ɛkk<br />

∂ψ<br />

<br />

∂¯ɛij<br />

(A.8)<br />

<strong>The</strong> eq.(A.8) is not applicable for fully anisotropic material behavior. <strong>The</strong> volumetric<br />

(σii) and deviatoric (¯σij) stress tensor can now be derived using equation A.8,<br />

σii =<br />

<br />

∂ψ<br />

ρ δii +<br />

∂ɛkk<br />

∂ψ<br />

<br />

∂¯ɛii<br />

= 3ρ ∂ψ<br />

∂ɛkk<br />

¯σij = σij − 1<br />

3 δijσkk<br />

= ρ ∂ψ<br />

∂¯ɛij<br />

Hence, in this manner the material stress-strain response can be deduced.<br />

165<br />

(A.9)<br />

(A.10)<br />

(A.11)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!