138
MULTISCALE MODELING OF CONTACT PLASTICITY AND NANOINDENTATION IN NANOSTRUCTURED FCC METALS A Dissertation Presented by Virginie Dupont to The Faculty of the Graduate College of The University of Vermont In Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy Specializing in Mechanical Engineering October, 2008

MULTISCALE MODELING OF CONTACT PLASTICITY AND ...pages.erau.edu/~dupontv/Dissertation.pdfatom-method potentials for aluminum were used in order to study the effect of the potential

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • MULTISCALE MODELING OF CONTACT PLASTICITY AND

    NANOINDENTATION IN NANOSTRUCTURED FCC METALS

    A Dissertation Presented

    by

    Virginie Dupont

    to

    The Faculty of the Graduate College

    of

    The University of Vermont

    In Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy Specializing in Mechanical Engineering

    October, 2008

  • Accepted by the Faculty of the Graduate College, The University of Vermont, in partial fulfillment of the requirements for the degree of Doctor of Philosophy, specializing in Mechanical Engineering.

    Dissertation Examination Committee: Advisor Frédéric Sansoz, Ph.D. Dryver Huston, Ph.D. Yves Dubief, Ph.D. Chairperson Dennis Clougherty, Ph.D. Vice President for Research Frances E. Carr, Ph.D. and Dean of Graduate Studies Date: September 5th, 2008

  • ABSTRACT

    Nanocrystalline thin films are materials with a grain size less than 100 nm which

    are commonly used to fabricate microscale electro-mechanical devices. At such small scale, nanoindentation is the only standard experimental technique to study the mechanical properties of thin films. However, it is unclear if the continuum laws commonly used in nanoindentation analysis of polycrystalline materials are still valid for nano-grained metals. It is therefore critical to better understand the behavior of nanocrystalline materials under nanoscale contact. This dissertation summarizes the results of atomistic simulations aimed at modeling the nanoindentation of nanocrystalline metal thin films for which the grain size is smaller than the indenter diameter.

    The nanoindentation of aluminum thin films was first studied using the Quasicontinuum method, which is a concurrent multiscale model where regions of small gradients of deformations are represented as a continuum medium by finite elements, and regions of high gradients of deformation are fully-treated atomistically. Two embedded-atom-method potentials for aluminum were used in order to study the effect of the potential on the nanoindentation behavior. The aim is to better understand the effects of a grain boundary network on the plasticity and the underlying mechanisms from an atomistic perspective. Our results show that a grain boundary network is the primary medium of plasticity at the nanoscale, via shear banding that causes flow serration. We also show that although the dislocation mechanisms are the same, the mechanisms involving grain boundaries are different depending on the interatomic potential.

    In a second part, abnormal grain growth in aluminum thin films under nanoindentation is studied using both the Quasicontinuum method and parallel molecular dynamics simulations. The effects of the potential, the nature of the indenter and of its size on the grain growth under nanoindentation are investigated. Our results show that the potential used, which can be related to the purity of the material, can reduce grain growth. We also show that the size and material used for the indenter both have significant effects on grain growth. More specifically, grain growth under the indenter is found to occur via atom diffusion if the indenter is of the same material as the thin film.

    Finally, the sample size effects were studied using parallel molecular dynamics simulations on nickel thin films and nanowires. Single crystals with different sizes are modeled in order to investigate the effects of the free boundaries as well as of the thickness of the samples. It is shown that the yield point and the incipient plasticity mechanisms are similar for all simulations. However, the hardness of the nanowires is found to decrease with the nanowire size during nanoindentation, due to the interaction of prismatic loops and dislocations with the free boundaries.

    This dissertation has shed light on the plastic deformation mechanisms under nanoscale contact. The results obtained will help the scientific community gain a better understanding of the behavior of nanomaterials, which will lead to the fabrication of more reliable nanodevices.

  • ii

    ACKNOWLEDGMENTS

    I would like to thank my advisor, Frederic Sansoz, for allowing me to work

    with him and for his guidance and support during my graduate career at the University of

    Vermont. The subject we worked on was very interesting, and he continued to push me to

    achieve my goals and challenged me by sharing our thoughts on the projects.

    I would also like to thanks Bertrand Rollin, for helping me through this

    endeavor, research-wise, physically and psychologically. He helped through the most

    difficult parts of the PhD.

    I would like to extend my thanks to Drs. Yves Dubief, Dryver Huston and

    Dennis Clougherty for taking time out of their busy schedules and agreeing to be on my

    committee.

    Thanks to Steve Plimpton and Ellad Tadmor for the codes they developed and

    shared, and for their support to all my questions (Tadmor; Plimpton). Thanks to Jim

    Lawson for his support with the VACC.

    Finally, thanks to my family back in France for their support during this

    difficult period of my life. Thanks to Karen Bernard and Michelle Mayette, for their help

    throughout my studies and their unwavering good mood. Thanks to Ida Russin and Mike

    Cook for helping me the many times I went to see them at the Grad College. Finally,

    thanks to all my friends: Lucie, Sophie, Aurélie L, Aurélie G, Cécile, Amélie, Marie-

    Agnès, Montse, Xabier, Fahmi, Carl, Ana, Chris, Benji, Nirav, Ben and those I am

    forgetting.

  • iii

    TABLE OF CONTENTS

    ACKNOWLEDGMENTS................................................................................................... ii

    LIST OF TABLES............................................................................................................ vii

    LIST OF FIGURES..........................................................................................................viii

    CHAPTER 1: INTRODUCTION........................................................................................ 1

    1.1. Motivations and Objectives ................................................................................ 1

    1.2. State of Knowledge............................................................................................. 8

    1.2.1. Contact Plasticity in Nanocrystalline Thin Films....................................... 8

    1.2.2. Grain Growth Mechanisms at Atomic Scale ............................................ 12

    1.2.2.1. Thermodynamically-activated Grain Growth............................. 12

    1.2.2.2. Stress-assisted Grain Growth...................................................... 13

    1.2.3. Size Effects in Nanosized Structures........................................................ 15

    1.3. Plan of the Dissertation..................................................................................... 17

    CHAPTER 2: NUMERICAL METHODS....................................................................... 18

    2.1. The Quasicontinuum Method ........................................................................... 18

    2.2. Parallel Molecular Dynamics............................................................................ 20

    2.3. Modeling of Spherical/Cylindrical Contact ...................................................... 23

    2.4. Interatomic Potentials ....................................................................................... 24

    2.4.1. Embedded Atom Method Potentials ......................................................... 24

    2.4.2. Tersoff Potential ....................................................................................... 25

    2.4.3. Morse Potential ......................................................................................... 27

    2.5. Calculation of Stresses...................................................................................... 27

  • iv

    2.6. Voronoi Construction........................................................................................ 28

    2.7. Tools for the Visualization of Defects, Dislocations and Grain Boundaries in

    Atomistic Simulations....................................................................................................... 29

    2.7.1. Centro-symmetry Parameter ..................................................................... 29

    2.7.2. Ackland Parameter.................................................................................... 30

    2.8. Validity of the 3D Models ................................................................................ 31

    CHAPTER 3: EFFECTS OF A GRAIN BOUNDARY NETWORK ON INCIPIENT PLASTICITY DURING NANOSCALE CONTACT...................................................... 34

    3.1. Objectives ......................................................................................................... 34

    3.2. Model ................................................................................................................ 35

    3.3. Characterization of the EAM Potentials ........................................................... 38

    3.4. Force and Contact Pressure Calculations.......................................................... 40

    3.5. Determination of the Yield Point...................................................................... 41

    3.6. Shear Localization Mechanisms ....................................................................... 44

    3.7. Effects of the Interatomic Potential on Plasticity in a 7 nm-grain-size Model. 46

    3.7.1. Effects of EAM Potential on GB Structure and Energy at Equilibrium... 46

    3.7.2. Mechanical Response Under Nanoscale Contact ..................................... 49

    3.7.3. Incipient Mechanisms of Plasticity at Yield Point (δ < 20 Å).................. 53

    3.7.4. GB-mediated Plasticity Mechanisms (δ > 20 Å)...................................... 55

    3.8. Discussion ......................................................................................................... 58

    3.8.1. Similarities Between Simulation and Experiments in Nanocrystalline

    Metal Indentation .............................................................................................................. 58

  • v

    3.8.2. Effects of Interatomic Potential on GB-mediated Plasticity..................... 60

    3.8.3. Incidence of Impurity on Flow Stress and GB-mediated Plasticity in

    Nanocrystalline Metals ..................................................................................................... 61

    3.9. Conclusions....................................................................................................... 63

    CHAPTER 4: FUNDAMENTAL MECHANISMS OF GRAIN GROWTH DURING THIN FILM NANOINDENTATION............................................................................... 64

    4.1. Objectives ......................................................................................................... 64

    4.2. Models............................................................................................................... 65

    4.2.1. Quasicontinuum Model ............................................................................ 65

    4.2.2. Molecular Dynamics Model ..................................................................... 65

    4.3. Stress-assisted Grain Growth............................................................................ 67

    4.3.1. Grain Growth at 0K .................................................................................. 67

    4.3.2. Grain Growth at 300K .............................................................................. 71

    4.4. Grain Growth under Spherical Contact............................................................. 75

    4.5. Discussion ......................................................................................................... 80

    4.5.1. Stress-assisted Grain Growth.................................................................... 80

    4.5.2. Thermodynamically-activated Grain Growth........................................... 81

    4.6. Conclusions....................................................................................................... 85

    CHAPTER 5: STUDY OF SIZE EFFECTS ON SINGLE CRYSTAL PLASTICITY IN THIN FILMS AND NANOWIRES.................................................................................. 87

    5.1. Objectives ......................................................................................................... 87

    5.2. Model ................................................................................................................ 87

  • vi

    5.3. Results............................................................................................................... 91

    5.3.1. Elastic Behavior and Limit of Elasticity................................................... 91

    5.3.2. Plastic Behavior in Thin Films ................................................................. 93

    5.3.3. Plastic Behavior in Nanowires.................................................................. 96

    5.4. Discussion ....................................................................................................... 101

    5.5. Conclusions..................................................................................................... 104

    CHAPTER 6: CONCLUSIONS......................................................................................106

    REFERENCES................................................................................................................108

  • vii

    LIST OF TABLES

    Table 3.1: Stacking fault energy (γSF), unstable stacking fault energy (γUSF), unstable twinning fault energy (γUTF) and GB energy (γGB) for three Σ tilt symmetric grain boundaries calculated from quasicontinuum method on the two EAM Al potentials investigated, and reference values from first-principles simulations for pure Al and Al with impurities. All units of energy are in mJ/m². ............................................................ 40

    Table 3.2: Constitutive parameters extracted from force-displacement nanoindentation curves obtained by quasicontinuum simulation in nanocrystalline Al with a mean grain size of 7 nm. The parameters δf and δmax represent the depth of the residual impression after unloading and the maximum penetration depth, respectively. ................................. 51

    Table 4.1: Molecular simulations performed on an aluminum thin film. All the tips are rigid. .................................................................................................................................. 66

    Table 5.1: Young’s modulus and mechanical characteristics at yield point in Ni thin films and nanowires from molecular dynamics simulations of spherical indentation. .............. 93

    Table 5.2: Number of dislocations nucleated and absorbed by free surfaces for each nanowire............................................................................................................................ 99

  • viii

    LIST OF FIGURES

    Fig. 1.1: (a) Scanning Electron Microscope (SEM) picture of a made-to-measure nanoindentation probe for the group’s Atomic Force Microscope (AFM); (b) AFM picture of nanoindentation tests performed on a thin film. Pictures courtesy of Travis Gang, Helix 2008. ............................................................................................................... 2

    Fig. 1.2: In situ nanoindentation experiments of Al thin film. (a) Load as a function of displacement. (b) Load as a function of displacement for the leading portion of the loading curve. (c) Initial image of the Al grain. Note that it is free of dislocations. (d) Pictures corresponding to event 1 on the loading curve. Note the appearance of dislocations in the grain, characterized by the darker color. (e) Pictures corresponding to event 2 on the loading curve. Note the appearance of more dislocations in the grain. Pictures courtesy of Minor et al. (Minor et al., 2006)......................................................... 3

    Fig. 1.3: TEM observation during an in situ nanoindentation on nanocrystalline aluminium. (a) no grains in strong diffraction condition under the tip area indicated by the white arrow; (b) a grain with size about 10 nm has rotated into strong diffraction condition; (c) a group of grains in bright contrast; (d) the size of the group has become larger with increasing load. Picture courtesy of Jin et al. (Jin et al., 2004). ....................... 4

    Fig. 1.4: A comparison of (a) number of fraction and (b) volume fraction grain size distribution after 30 min of indenter dwell time for indents made at room temperature and -190°C. The arrow in (a) indicates the presence of the large grains after 30 min dwell time at the low temperature. The presence of large grains is more evident in the volume distribution. Picture courtesy of Zhang et al. (Zhang et al., 2005b). .................................. 5

    Fig. 1.5: Elastic moduli of ZnO nanowires ENW and GNW obtained from nanoindentation measurement and fitted as functions of the nanowire radius. Graph courtesy of Stan et al. (Stan et al., 2007). ............................................................................................................... 6

    Fig. 1.6: A SEM image of a Ni 20-µm-diameter nanopillar after application of a 4% compression strain. The arrow points at the steps left on the surface by the escaping dislocations. Picture courtesy of Uchic et al. (Uchic et al., 2004). ..................................... 7

    Fig. 1.7: Thermodynamically-activated grain growth processes. (a) and (b) curvature driven grain growth; (c) and (d) atom diffusion. .............................................................. 13

    Fig. 1.8: Stress-assisted grain growth. (a) Original configuration; (b) grain rotation and coalescence or (c) shear-coupled motion. Grains boundaries are represented by thick continuous lines. Thin lines represent the crystal orientation........................................... 14

    Fig. 2.1: Schematics of the Voronoi construction of a 2D model..................................... 29

  • ix

    Fig. 2.2: Evolution of the contact pressure as a function of the indentation depth for both models. The arrows indicate the yield points.................................................................... 33

    Fig. 2.3: Indentation step 250 ps after the yield point (a) for the 40 nm ä 12 nm ä 40 nm model and (b) the 60 nm ä 12 nm ä 60 nm model. The atoms in perfect FCC lattice have been removed. ................................................................................................................... 33

    Fig. 3.1: Quasicontinuum model of a 7 nm-grain size Al thin film indented by a 15 nm radius cylindrical indenter. (a) Full view of both finite element domain and atomistic region. (b) Close-up view of full atomistic zone near the contact region in unrelaxed configuration. .................................................................................................................... 36

    Fig. 3.2: Generalized stacking and planar fault energy curves obtained by quasicontinuum method with the Mishin-Farkas EAM potential for Al..................................................... 39

    Fig. 3.3: Effect of nanosized grains on the nanoindentation response of Al substrates from molecular static simulation using the Al-VC potential. The indenter radius is 15 nm. Serrated plastic flow clearly appears in the two nanocrystalline Al substrates under nanoindentation................................................................................................................. 42

    Fig. 3.4: Contact pressure versus penetration depth plot for a 7-nm polycrystal with the Al VC potential, along with the corresponding theoretical fitting.................................... 43

    Fig. 3.5: Thin shear band formation in 5-nm-grain-size nanocrystalline Al after 2.5-nm-deep indentation. (a) Partial view of the contact interface and location of the grain cluster associated with the shear band. (b) Enlarged view of the shear plane. A mechanical twin nucleated at the triple junction in the prolongation of the shear place is clearly visible in grain 2. (c) Magnitude and direction of atomic displacements between two loading increments represented by arrows. The shear band results from sliding of aligned grain boundaries (grains 3 and 4) and intragranular partial slip (grain 2). ................................ 45

    Fig. 3.6: Statistics of misorientation angle and GB structure in simulated nanocrystalline Al films after force relaxation as a function of interatomic potential. (a), (b) Distribution of misorientation angles (ψ + ψ’) between grains. (c), (d) Degree of symmetry of the GB structure from perfectly-symmetrical tilt GB (STGB, ψ – ψ’ ~ 0) to highly-asymmetrical tilt GB (ATGB, ψ – ψ’ ~ 180°). ....................................................................................... 48

    Fig. 3.7: Atomic energies (in eV) calculated after relaxation of a cluster of 6 nano-grains in the contact zone. (a) Schematic representing the GB distribution and corresponding grain number as indicated in figure 2. (b) Voter-Chen EAM potential. (c) Mishin-Farkas EAM potential................................................................................................................... 49

    Fig. 3.8: Contact pressure – displacement curves predicted by quasicontinuum simulation using Al-VC and Al-MF interatomic potentials. The curve for the Al-MF potential has been shifted to the right for clarity. The dashed curves correspond to the contact response

  • x

    of an isotropic elastic surface deformed by a perfectly-rigid, wedge-like cylinder, obtained by continuum theory. ......................................................................................... 50

    Fig. 3.9: Evolution of contact pressure as a function of penetration depth for shallow indentation using Al-VC potential in a 7 nm-grain-size simulation. Close-up views represent atomic details of deformation in the contact region. (a) Nucleation of the very first dislocation. (b) Nucleation and evolution of new dislocations. (c) Drop in curve corresponding to a sudden increase in contact area, just before the yield point. (d) Nanocrystal after the yield point....................................................................................... 52

    Fig. 3.10: Evolution of contact pressure as a function of penetration depth using Al-MF potential, for shallow indentation in a 7 nm-grain-size simulation. Close-up views represent atomic details of deformation in the contact region. (a) Nucleation of the very first dislocation. (b) Nucleation and evolution of the structure. (c) Nanostructure before the yield point. (d) Nanostructure after the yield point..................................................... 53

    Fig. 3.11: Effects of interatomic potential on the structure of a low-angle GB (ψ + ψ’ ~ 164°) before indentation (δ = 0 Å) and at final indentation (δ = 80 Å). (a) Al-VC potential; (b) Al-MF potential. Continuous lines indicate the [110] crystal lattice direction. ........................................................................................................................... 56

    Fig. 3.12 : Effects of interatomic potential on the structure of a high-angle GB (ψ + ψ’ ~ 117°) before indentation (δ = 0 Å) and at mid indentation (δ = 40 Å). (a) Al-VC potential; (b) Al-MF potential. Continuous lines indicate the [110] crystal lattice direction........... 57

    Fig. 4.1: Polycrystalline 3D model of an aluminum thin film deformed by a spherical tip of aluminum. The radius of the tip is 9 nm, and the mean grain size is 7 nm. The crystallographic orientations of the tip are as indicated. The model is made of 3 million atoms. ................................................................................................................................ 66

    Fig. 4.2: Evolution of the GB structure at the interface between grains 3 and 5 for (a) Al-VC potential and (b) Al-MF. ............................................................................................ 68

    Fig. 4.3: Atomic-level shear stress with respect to the orientation of the interface between grains 3 and 5 at different stages of the boundary motion (in units of GPa). The boundaries of grains 3 and 5 have been highlighted for clarity. ....................................... 69

    Fig. 4.4: Evolution of the misorientation angle at the interface between grains 3 and 5. Open symbols are used when the GB migrates, while plain symbols indicate that the GB is not moving..................................................................................................................... 70

    Fig. 4.5: Evolution of the microstructure of the sample indented by a 9 nm-radius Al indenter. General view (a) after relaxation and before indentation, ε = 0% and (b) at maximum indentation ε = 94%; (c) close-up view of grains 1, 2, 5, 6, 7 and 8 for ε = 0%; (d) close-up view of grains 1, 2, 5, 6, 7 and 8 for ε = 94%. The grain boundaries and

  • xi

    lattice defects are colored in white, and the atoms in FCC lattice are colored according to the grain they belonged to at the beginning of the simulation.......................................... 72

    Fig. 4.6: Evolution of the microstructure of grains 1, 5, 6, 7 and 8 from the samples indented by a 4, 9 and 15 nm-radius Al indenter under a 75% indentation strain. (a) With a 4 nm-radius Al indenter; (b) With a 9 nm-radius indenter; (c) With a 15 nm-radius indenter. The atoms belonging to grain boundaries or lattice defects are colored in white, and the other atoms are colored according to the grain they belonged to at the beginning of the simulation. .............................................................................................................. 74

    Fig. 4.7: Close-up view of the region below the tip for the nanocrystalline thin films indented by (a) and (b) a 4 nm-radius Al tip, (c) and (d) a 9 nm-radius Al tip and (e) and (f) a 15 nm-radius Al tip. The left column represents the Ackland parameter, and the right column shows the same view with atoms in grain boundaries or lattice defects in white and other atoms colored according to the grain they belonged to at the beginning of the simulation.......................................................................................................................... 76

    Fig. 4.8 : Microstructure of the thin films (a) at 0% indentation strain for the virtual tip; (b) the polycrystalline thin film indented by a 9 nm-radius virtual tip after a 75% indentation strain; (c) a polycrystalline thin film indented by a 9 nm-radius Al tip after a 75% indentation strain; (d) a single crystalline thin film indented by a 9 nm-radius Al tip after a 75% indentation strain. .......................................................................................... 77

    Fig. 4.9: (a) Evolution of Ly and Lz for the simulation with the aluminum tip with a radius of 9 nm and (b) Evolution of the energies below the tip for the polycrystal indented by a 9 nm-radius Al and virtual tip, and the single crystal indented by a 9 nm-radius Al tip. .... 79

    Fig. 4.10 : Atom diffusion processes. (a) Direct interchange; (b) Rotation; (c) By vacancy formation........................................................................................................................... 82

    Fig. 4.11 : 7 nm-grain size Al thin film indented by a diamond tip under a 50% indentation strain............................................................................................................... 84

    Fig. 5.1: Atomistic models of spherical indentation for thin films and nanowires. The radius of the virtual indenter is 18 nm. (a) 30 nm-thick film. (b) 30 nm-diameter nanowire. (c) 12 nm-thick film. (d) 12 nm-diameter nanowire. The crystallographic orientation of the films and nanowires is as indicated. Free surfaces appear in dark-gray, while atoms in FCC lattice are colored in light-gray........................................................ 89

    Fig. 5.2: Simulated force-displacement nanoindentation curves for Ni thin films and nanowires indented by a spherical tip. Eq. (5.2) fitted to the 30 nm-thick film model is also represented................................................................................................................. 93

  • xii

    Fig. 5.3: Evolution of the mean contact pressure as a function of penetration depth in (a) thin films and (b) nanowires. The numbers indicate the occurrence of dislocation absorption by free surfaces in the nanowires. ................................................................... 94

    Fig. 5.4 : Atomic-level representation of plastic deformation in Ni thin films during spherical indentation. (a) 12 nm-thick film at yield point. (b) 12 nm-thick film at maximum indentation depth. (c) 30 nm-thick film at yield point. (d) 30 nm-thick film at maximum indentation depth. Dashed lines represent the trajectory of prismatic dislocation loops. X = ]111[ ; Y = ]211[ ; Z = ]101[ . ....................................................... 95

    Fig. 5.5: Atomic-level representation of plastic deformation in a 12 nm-diameter Ni nanowire during spherical indentation. (a,b) Yield point. (c,d) Maximum indentation depth. X = ]111[ ; Y = ]211[ ; Z = ]101[ . .......................................................................... 97

    Fig. 5.6: Atomic-level representation of plastic deformation in a 30 nm-diameter Ni nanowire during spherical indentation. (a,b) Yield point. (c,d) Maximum indentation depth. Dashed lines represent the trajectory of prismatic dislocation loops escaping at free surfaces. X = ]111[ ; Y = ]211[ ; Z = ]101[ . ...................................................................... 98

    Fig. 5.7: Dislocation mechanisms in the nanowires. (a) Local von Mises strain at the yield point indicating a homogeneous nucleation of the first dislocation in the 12-nm nanowire; (b) Absorption of the first half-dislocation loop in the 12-nm nanowire; (c) Creation of the first four-sided prismatic loop in the 30-nm nanowire, and comparison with the size of the 12-nm nanowire. ................................................................................................... 100

  • 1

    CHAPTER 1: INTRODUCTION

    1.1. Motivations and Objectives

    Nanocrystalline materials are materials with a grain size less than 100 nm. They

    are commonly used to fabricate microscale electro-mechanical systems (MEMS).

    However, MEMS technology could be substantially improved if we could broaden our

    knowledge about the mechanical behavior of nanomaterials. Indeed, it is now well-

    established in FCC metals and alloys that a marked transition in plasticity mechanism,

    accompanied by a continuous change in mechanical behavior, operates with a reduction

    of grain size from the microcrystalline to the nanocrystalline regime (Schiotz and

    Jacobsen, 2003; Trelewicz and Schuh, 2007; Chang and Chang, 2007). Several

    experimental techniques (Hemker and Sharpe, 2007) have been used to study the

    mechanical behavior of thin films, such as tensile testing (Bagdahn et al., 2003; Tsuchiya

    et al., 1996; Sharpe et al., 1997; Chasiotis and Knauss, 2002; Chasiotis and Knauss, 2000;

    Cho et al., 2005), or membrane deflection experiment (MDE) (Espinosa et al., 2003a;

    Espinosa et al., 2003b). Small-scale contact experiments spanning from simple micro-

    hardness testing to fully instrumented nanoindentation have been used for some time to

    characterize the nature of yield phenomena and the influence of grain size on hardness

    and strengthening in nanocrystalline metals (Nieman et al., 1989; Elsherik et al., 1992;

  • 2

    Fougere et al., 1995; Qin et al., 1995; Farhat et al., 1996; Sanders et al., 1997; Malow et

    al., 1998). Nanoscale contact probes are particularly well-suited for the studies of the

    plasticity transition in nanograined metals, because they can be highly sensitive to the

    heterogeneous nature of plastic deformation in very confined volumes of materials, and

    they provide quantitative insights into the mechanisms governing incipient plasticity at

    reduced length scales. A picture of an indenter tip that was made-to-measure for the

    group’s Atomic Force Microscope (AFM) is shown in Fig. 1.1a, and resulting indentation

    tests performed on a nickel thin film are shown in Fig. 1.1b.

    Fig. 1.1: (a) Scanning Electron Microscope (SEM) picture of a made-to-measure nanoindentation

    probe for the group’s Atomic Force Microscope (AFM); (b) AFM picture of nanoindentation tests

    performed on a thin film. Pictures courtesy of Travis Gang, Helix 2008.

  • 3

    Fig. 1.2: In situ nanoindentation experiments of Al thin film. (a) Load as a function of

    displacement. (b) Load as a function of displacement for the leading portion of the loading curve.

    (c) Initial image of the Al grain. Note that it is free of dislocations. (d) Pictures corresponding to

    event 1 on the loading curve. Note the appearance of dislocations in the grain, characterized by

    the darker color. (e) Pictures corresponding to event 2 on the loading curve. Note the appearance

    of more dislocations in the grain. Pictures courtesy of Minor et al. (Minor et al., 2006).

    Recent experiments on aluminum thin films have shown that the onset of

    plasticity in nanomaterials is more complex than was previously thought. Minor et al.

    (Minor et al., 2006) performed in situ transmission electron microscopy (TEM)

    experiments of nanoindentation on a polycrystalline aluminum film. They show that

    contrary to previous conclusions on the onset of plasticity in single crystals, dislocation

  • 4

    activity could occur before the yield point was attained in the film. In Fig. 1.2, they show

    that the two very small events marked 1 and 2 correspond to the emission of dislocations,

    as can be observed in Fig. 1.2d and e when the grain becomes darker. The nucleation of

    the first dislocation in a single crystal is believed to correspond to the onset of plasticity,

    but this experiment shows that polycrystalline materials behave differently. It is then

    worth investigating what corresponds to the onset of plasticity for nanocrystalline

    materials and what are the mechanisms associated with plasticity.

    Fig. 1.3: TEM observation during an in situ nanoindentation on nanocrystalline aluminium: (a) no

    grains in strong diffraction condition under the tip area indicated by the white arrow; (b) a grain

    with size about 10 nm has rotated into strong diffraction condition; (c) a group of grains in bright

    contrast; (d) the size of the group has become larger with increasing load. Picture courtesy of Jin

    et al. (Jin et al., 2004).

  • 5

    Fig. 1.4: A comparison of (a) number of fraction and (b) volume fraction grain size distribution

    after 30 min of indenter dwell time for indents made at room temperature and -190°C. The arrow

    in (a) indicates the presence of the large grains after 30 min dwell time at the low temperature.

    The presence of large grains is more evident in the volume distribution. Picture courtesy of Zhang

    et al. (Zhang et al., 2005b).

    Another interesting phenomenon that was observed during the nanoindentation

    of nanocrystalline FCC metals is abnormal grain growth. It was observed in aluminum by

    Jin et al. (Jin et al., 2004) during a TEM experiment of nanoindentation. Fig. 1.3 shows

    the TEM images of several steps during nanoindentation of nanocrystalline aluminum.

    The brightness is indicative of the orientation of the grains, so that Fig. 1.3 shows that a

    bigger grain is forming under the indenter. The final size of the bright spot is five times

    the original grain size after 3 seconds of indentation. Abnormal grain growth was also

    observed by Zhang et al. (Zhang et al., 2005b) in nanocrystalline copper. They show that

    grain growth is faster at cryogenic temperatures than at room temperature, and that the

  • 6

    purity of the material has an influence on grain growth. Their result in Fig. 1.4 shows that

    for the same time under the indenter, the sample at cryogenic temperature forms grains

    much bigger than the sample at room temperature, which is particularly striking in the

    volume fraction plot. It is now well established that the strength of a material depends on

    its grain size (Schiotz and Jacobsen, 2003; Hall, 1951; Petch, 1953), which renders the

    process of grain growth undesirable in most nanotechnological applications. It is then

    important to study this phenomenon in order to learn how to control it.

    Fig. 1.5: Elastic moduli of ZnO nanowires ENW and GNW obtained from nanoindentation

    measurement and fitted as functions of the nanowire radius. Graph courtesy of Stan et al. (Stan et

    al., 2007).

    The final interest of nanosized materials lies in the nanowires and nanopillars.

    Those cylindrical structures have been shown to present amazing size effects under

    microcompression, corresponding to an increase of the elastic moduli as their radius

  • 7

    decreases (Fig. 1.5) (Stan et al., 2007). The free boundaries introduced by the shape of

    the cylinder allow dislocations to leave the volume, as demonstrated in Fig. 1.6 with the

    lines left on the surface of the wire by the escaping dislocations (Uchic et al., 2004). This

    process changes the properties known to apply to bulk materials. It is therefore crucial to

    understand the mechanisms involved in the plasticity of nanosized metals.

    Fig. 1.6: A SEM image of a Ni 20-µm-diameter nanopillar after application of a 4% compression

    strain. The arrow points at the steps left on the surface by the escaping dislocations. Picture

    courtesy of Uchic et al. (Uchic et al., 2004).

    The objectives of this dissertation are three fold. The first objective is to

    examine the onset of plasticity and the underlying mechanisms of plasticity when a

    nanocrystalline grain boundary network is involved during contact. Second, the plasticity

    phenomenon of grain growth is further studied. Finally, the size effects in nanowires are

    analyzed, along with the plasticity mechanisms in such structures. Those studies will be

    performed using atomistic simulations of nanoindentation.

  • 8

    1.2. State of Knowledge

    1.2.1. Contact Plasticity in Nanocrystalline Thin Films

    In past work, nanoscale contact experiments in nanocrystalline metals have

    revealed two microstructure length scales producing different modes of plastic

    deformation. First, particular focus has been placed on examining how dislocations

    interact with surrounding grain boundaries (GBs) by performing nanoindentations at the

    center of single nanograins, that is, by forcing the contact area to be much smaller than

    the grain size. Yang and Vehoff (Yang and Vehoff, 2007) have observed that the

    dislocations, which nucleate below the indenter, only interact directly with the

    neighboring interfaces for grain sizes below 900 nm. At such scale, the point of elastic

    instability is clearly defined by a “pop-in” event whose width is strongly correlated to the

    size of the indented grain. The smaller the grain size, the smaller the pop-in width and the

    harder the material. For grain sizes comparable to the contact area, however, Minor et al.

    (Minor et al., 2006) have revealed using in-situ transmission electron microscopy (TEM)

    nanoindentation that significant dislocation activity takes place in ultrafine-grained Al

    thin films before the first obvious jump in displacement in the load-depth nanoindentation

    curves. This result challenged the prevailing notion that the first pop-in event

    corresponding to the onset of plasticity during nanoindentation occurs in a dislocation-

    free crystal.

    The second microstructure length scale, at which past nanoscale experiments

    have been performed, corresponds to contact areas much larger than the mean grain size.

  • 9

    In this case, it is the collective deformation of the nanocrystalline GB network that

    dominates the plastic behavior. All experimental evidence shows that the pile-up of

    deformation left around residual impressions varies dramatically from homogeneous at

    large grain size (> 20 nm) to inhomogeneous with intense plastic deformation in highly-

    localized shear bands for very small grain sizes (< 20 nm) (Malow et al., 1998; Trelewicz

    and Schuh, 2007; Andrievski et al., 2000; Van Vliet et al., 2003). Since shear band

    propagation is routinely observed in bulk amorphous metals under nanoindentation (Kim

    et al., 2002; Jana et al., 2004; Shi and Falk, 2005; Su and Anand, 2006; Antoniou et al.,

    2007; Lund and Schuh, 2004), it was suggested that the plastic deformation of

    nanocrystalline FCC metals with the finest grain size is comparable to the shear

    localization processes observed in bulk metallic glasses (Hartford et al., 1998). This

    assumption has recently been confirmed by the nanoindentation study of Trelewicz and

    Schuh (Trelewicz and Schuh, 2007) in nanocrystalline Ni-W alloys.

    Contact plasticity at the former length scale has been well-documented using

    atomistic modeling, because the underlying mechanisms are clearly defined by slip

    events and slip-GB interactions. A predictive understanding of the atomic mechanisms

    leading to shear localization under an indenter at very small grain sizes, however, has

    proved elusive, primarily for two reasons:

    • It is commonly acknowledged that the GB networks play a critical role in the process

    of plasticity in nanocrystalline metals. But GBs can also be involved in simultaneous,

    yet different plasticity mechanisms at very small grain sizes. These mechanisms

    include sources and sinks for lattice dislocations (Yamakov et al., 2002; Van

  • 10

    Swygenhoven et al., 2002; Tschopp and McDowell, 2008), GB sliding (Van Vliet et

    al., 2003; Schiotz et al., 1998; Schiotz et al., 1999; Hasnaoui et al., 2002), grain

    rotation-induced shear band formation and grain coalescence (Shimokawa et al.,

    2006; Hasnaoui et al., 2002; Wei et al., 2002; Fan et al., 2006; Joshi and Ramesh,

    2008; Wang et al., 2008; Yang et al., 2008; Haslam et al., 2001), and GB migration

    coupled to shear deformation (Haslam et al., 2001; Gianola et al., 2008a; Gianola et

    al., 2008b; Gianola et al., 2006a; Gianola et al., 2006b; Farkas et al., 2006; Gutkin et

    al., 2008). From past nanoindentation experiments, there is no clear consensus

    whether one or several of these atomic mechanisms truly dominate the onset of

    plasticity for the finest grain size. In particular, stress-assisted grain growth under an

    indenter has been observed at the onset of plasticity of nano-grained Al and Cu with

    a mean grain size of 20 nm (Jin et al., 2004; Zhang et al., 2005b; Gai et al., 2007). In

    a second group of nanoindentation experiments, however, neither grain growth nor

    GB migration processes have been observed in nano-Cu and NiAl with a mean grain

    size of 14 nm and 10 nm, respectively. Instead, it was concluded that the onset of

    plasticity was related to the nucleation of lattice dislocations from GBs (Chen et al.,

    2003a; Li and Ngan, 2005). Another study in plated Cu thin films with a 10 nm-grain

    size reported some evidence of void formation at GBs and triple junctions as a

    consequence of GB sliding during indentation (Chang and Chang, 2007).

    • Earlier attempts made to model the nanoindentation of nanocrystalline metals by

    atomistic simulations have employed a spherical repulsive force to model virtual tips

    varying from 30 Å to 98 Å in diameter (Feichtinger et al., 2003; Lilleodden et al.,

  • 11

    2003; Ma and Yang, 2003; Jang and Farkas, 2004; Hasnaoui et al., 2004; Saraev and

    Miller, 2005; Kim et al., 2006; Jang and Farkas, 2007). As such, contact areas were,

    to a large extent, smaller than the grain size, and the plastic zone produced by these

    tips was only limited to one or two grains. In contrast, some recent studies

    (Szlufarska et al., 2005) have shown that it is critically important to simulate

    nanoindentation tips with more realistic sizes, in order to put into perspective the

    collective plastic processes of the GB networks in nanocrystalline materials.

    Szulfarska et al. (Szlufarska et al., 2005) have simulated the nanoindentation of

    normally-brittle nanocrystalline ceramics with a four to one ratio between tip

    diameter and grain size, which revealed unusual GB-mediated plastic behavior.

    Furthermore, past simulations have used different embedded-atom-method (EAM)

    potentials, from which predictions of stacking fault energies can lead to strong

    differences within the same metal (Zimmerman et al., 2000). The impact of the

    interatomic potential on collective plastic processes, however, has never been fully

    characterized.

  • 12

    1.2.2. Grain Growth Mechanisms at Atomic Scale

    Intense grain refinement can promote drastic changes of plasticity mechanism in

    bulk materials and thin films (Schiotz et al., 1998; Lu et al., 2000; Yamakov et al., 2002;

    Chen et al., 2003b; Schiotz and Jacobsen, 2003; Kumar et al., 2003; Hasnaoui et al.,

    2003; Huang et al., 2006). In metals containing nanosized grains (

  • 13

    thermodynamically driven mechanism is GB atom diffusion (Smigelskas and Kirkendall,

    1947; Huntington and Seitz, 1942), during which the atoms jump in the crystal into point

    vacancies, creating a new vacancy in the process (Fig. 1.7 (c) and (d)). The availability of

    point vacancies follows an Arrhenius equation, so that the rate of atom diffusion

    increases with temperature.

    Fig. 1.7: Thermodynamically-activated grain growth processes. (a) and (b) curvature driven grain

    growth; (c) and (d) atom diffusion.

    1.2.2.2. Stress-assisted Grain Growth

    The last two mechanisms are stress-assisted, and explain why grain growth is

    still possible at cryogenic temperatures (Zhang et al., 2005b). The first one is rotation-

  • 14

    induced grain coalescence (Haslam et al., 2001), during which one grain rotates and its

    orientation comes to match the orientation of a neighbor grain, thus forming a single

    bigger grain (Fig. 1.8b). This process is often associated with GB sliding (Cahn et al.,

    2006). The final mechanism of grain growth is called shear-coupled motion (Fig. 1.8c).

    In this mechanism, the normal motion of grain boundaries results from a shear stress

    applied tangentially to them and causing tangential motion, or coupled motion (Cahn and

    Taylor, 2004; Suzuki and Mishin, 2005; Sansoz and Molinari, 2005). It was shown that

    the GB structure greatly influences the behavior of the GB between the mechanisms of

    grain sliding or shear-coupled motion (Sansoz and Molinari, 2005).

    Fig. 1.8: Stress-assisted grain growth. (a) Original configuration; (b) grain rotation and

    coalescence or (c) shear-coupled motion. Grains boundaries are represented by thick continuous

    lines. Thin lines represent the crystal orientation.

  • 15

    The complexity and the variety of mechanisms of grain growth make it a

    difficult phenomenon to study and to observe experimentally. Atomistic simulations offer

    the advantage of investigating an atomistically the phenomenon given the atomic

    potential as a single input. This method was shown to be very successful in characterizing

    plastic deformation mechanisms in nanocrystalline metals as shown in Haslam et al.

    (Haslam et al., 2001) and in Chapter 4 of this dissertation.

    1.2.3. Size Effects in Nanosized Structures

    One-dimensional metal nanowires (Tian et al., 2003) are the building blocks for

    nanoscale research in a vast variety of disciplines that range from biology, to electro-

    mechanics and photonics (Mock et al., 2002; Husain et al., 2003; Bauer et al., 2004;

    Barrelet et al., 2004). These nanomaterials have recently stimulated the interest of the

    mechanics community, because all experimental evidence shows a strong influence of the

    sample dimension on the mechanical properties of metals at nanometer scale (Dimiduk et

    al., 2005; Uchic et al., 2004; Greer and Nix, 2006; Greer et al., 2005; Wu et al., 2005;

    Volkert and Lilleodden, 2006). While the strength and ductility of metals in macroscopic

    samples are predominantly determined by the relevant microstructure length scale (e.g.

    grain size), which is often small relative to the sample size, a distinctive behavior of

    crystal plasticity emerges in metal nanostructures, where the material strength

    significantly increases as the deformation length scale (diameter or volume) decreases. A

    micro-plasticity mechanism has been proposed to account for the size scale dependence

    of small metallic samples based on dislocation starvation, in which the density of mobile

  • 16

    dislocations created from pre-existing dislocation sources is counter-balanced by the

    density of dislocations escaping the crystal at free surfaces (Greer et al., 2005; Greer and

    Nix, 2006; Tang et al., 2007). The in-situ TEM compression experiments of Shan et al.

    (Shan et al., 2008) have recently confirmed this mode of deformation in Ni nanopillars as

    small as 150 nm in diameter. Nevertheless, it remains crucial to characterize the influence

    of sample size on dislocation activity at even smaller length-scale (< 100 nm) in order to

    achieve meaningful results in the crystal plasticity of metal nanowires.

    Nanoindentation technique via pillar compression method (Dimiduk et al., 2005;

    Uchic et al., 2004; Greer and Nix, 2006; Greer et al., 2005; Volkert and Lilleodden, 2006;

    Shan et al., 2008) has enabled rapid progress in the experimental investigation of

    nanomechanical properties and their size dependence in metals at the micron and

    submicron scales. However, the nanopillar compression method has never been applied

    to samples less than 100 nm in diameter due to complications in preparation and

    mechanical testing at such a small scale. By contrast, the use of nanoindentation tips to

    probe the radial elastic modulus and hardness of sub-100 nm nanowires has been found

    very successful in the past (Stan et al., 2007; Lucas et al., 2008; Feng et al., 2006; Lee et

    al., 2006; Li et al., 2003; Tao and Li, 2008; Zhang et al., 2008; Liang et al., 2005; Bansal

    et al., 2005; Fang and Chang, 2004). While a fundamental understanding of dislocation

    activity during metal nanopillar compression has already been supplemented by atomistic

    simulations (Rabkin et al., 2007; Rabkin and Srolovitz, 2007; Afanasyev and Sansoz,

    2007; Zhu et al., 2008; Cao and Ma, 2008), the atomic mechanisms of plasticity and

    related size effects for metal nanowires deformed by nanoindentation remain elusive.

  • 17

    1.3. Plan of the Dissertation

    The numerical methods used for this research, including both quasicontinuum

    method and molecular dynamics as well as a description of the interatomic potentials, are

    presented in Chapter 2. Chapter 3 shows the study on the effects of a grain boundary

    network on the incipient plasticity of nanocrystalline Al deformed by a cylindrical

    contact. The effects of different factors on grain growth under an indenter are

    investigated in Chapter 4. Chapter 5 presents the study on the size effects on contact-

    induced plasticity in Ni nanowires and thin films. Finally, the major conclusions of this

    dissertation are summarized in Chapter 6.

  • 18

    CHAPTER 2: NUMERICAL METHODS

    Two different molecular simulation techniques have been used for the

    calculations: The Quasicontinuum Method (Miller and Tadmor, 2002), which is

    multiscale atomistic/finite element simulation technique, and parallel three-dimensional

    molecular dynamics simulations. Both methods and additional numerical tools used for

    this study are presented in the following.

    2.1. The Quasicontinuum Method

    A complete description of the method can be found in the article written by the

    developers of the Quasicontinuum (QC) method, Ronald E. Miller and E.B. Tadmor

    (Miller and Tadmor, 2002).

    The Quasicontinuum Method is a multiscale atomistic/finite element simulation

    technique. It combines the advantages of continuum finite element simulation methods

    with those of molecular dynamics. At the atomistic scale, finite element methods do not

    represent accurately the behavior of a material because they are based on the hypothesis

    that the material is a continuum, whereas it is actually made up of discrete particles. On

    the other hand, molecular dynamics simulations allow a very accurate representation of

    the material by representing all the atoms. But even with today’s supercomputers’

  • 19

    capacities, the number of atoms is limited and the simulated sample has small

    dimensions. This is usually taken care of with periodic boundary conditions, but in the

    case of nanoindentation, this implies that an infinite number of indenters indent the

    surface at the same time. The Quasicontinuum Method combines finite elements where

    deformations are small with an atomistic representation in high deformation areas. This

    allows the user to model larger models with an accurate representation of the material’s

    behavior where needed.

    A typical mesh is constituted of atomistic zones (non local) and finite element

    zones (local). The regions that sustain plastic deformations are modeled atomically,

    whereas the rest of the mesh is modeled by finite elements. Each node in the model is

    called “repatom” for representative atom. Each repatom can represent just itself (non

    local zone as well as some atoms of the interface), or more than itself (local zone). The

    total energy of the system is computed as follows:

    exact

    N

    tot EEnErep

    ≈= ∑=1α

    αα (2.1)

    where Nrep is the total number of repatoms in the system, nα is the number of real atoms

    the repatom is representing (nα = 1 for non local atoms), and Eα is the energy of each

    repatom. This formulation allows having the same calculation on both local and non-local

    regions, so that there is no discontinuity at the interface. The minimum energy is

    calculated at each step using a conjugate gradient method then a new set of forces is

    applied, the minimum energy found again and so on. The conjugate gradient method does

  • 20

    not take into account the effects of the temperature, so all calculations performed with

    QC are performed at 0K.

    The QC method can apply a “nonlocality criterion” to the model in order to

    verify whether atoms should be local or non-local. A cutoff rnl is defined, and the applied

    criterion will be:

    ελλ

  • 21

    ( ) ( )

    ii

    j kkji

    jji

    ii

    vdtrd

    rrrFrrFdtvd

    m

    rr

    rrrrrr

    =

    ++= ∑∑∑ ...,,, 32 (2.3)

    where mi is the mass of atom i, irr and iv

    r are its position and velocity vectors, F2 is a

    force function describing pairwise interactions between atoms, F3 describes three-body

    interactions, and many-body interactions can be added. The Verlet integration algorithm

    was chosen in order to calculate the atoms’ positions.

    The Verlet integration algorithm calculates the position of the atoms at the next

    time step from the positions at the previous and current time steps, without using the

    velocity. It is derived by writing two Taylor expansions of the position vector in different

    time directions:

    ( )433

    32

    2

    2)(

    61)(

    21)()()( tOtt

    dt

    rdtt

    dt

    rdtt

    dtrd

    trttr iiiii ∆+∆+∆+∆+=∆+rrr

    rr (2.4)

    ( )4333

    22

    2)(

    61)(

    21)()()( tOtt

    dtrdtt

    dtrdtt

    dtrdtrttr iiiii ∆+∆−∆+∆−=∆−

    rrrrr (2.5)

    When adding the two expansions, we have:

    ( )422

    2)()(2)()( tOtt

    dt

    rdtrttrttr iiii ∆+∆+=∆−+∆+

    rrrr (2.6)

    or

    2)()()(2)( ttattrtrttr iii ∆+∆−−=∆+rrrr (2.7)

    with a(t) the acceleration.

  • 22

    This offers the advantage that the first and third-order terms from the Taylor expansion

    cancel out, thus making the Verlet integrator more accurate than integrations by simple

    Taylor expansion alone.

    By subtracting (2.4) and (2.5), and reducing the precision, we have:

    ( )2)(2)()( tOttdtrd

    ttrttr iii ∆+∆=∆−−∆+r

    rr (2.8)

    or, if we reorganize:

    t

    ttrttrt

    dtrd iii

    ∆∆−−∆+

    =2

    )()()(

    rrr

    (2.9)

    The first step has slightly different equations, as we cannot use )( tri ∆−r . Instead,

    we use the initial conditions:

    ( )32

    2)0(

    21)0()0()( tO

    dt

    rddtrd

    rtr iiii ∆+++≈∆rr

    rr (2.10)

    The calculations were performed using an NVT integration (constant number of

    atoms, constant volume of the simulation box, and constant temperature). This was

    achieved by using a Nose/Hoover temperature thermostat (Hoover, 1985). In this model,

    the equations are given by:

    ⎟⎟⎠

    ⎞⎜⎜⎝

    ⎛−=′

    ⋅−=′

    ′=′

    1

    )(

    0

    2TT

    pFpmp

    tr

    T

    ii

    ii

    νς

    ς rrr

    rr

    (2.11)

  • 23

    with T0 the desired constant temperature of the simulation, νT the thermostating rate, and

    ζ the thermodynamic friction coefficient. This constant dynamically modifies the

    velocities of each atom such that the temperature of the model tends to T0.

    The actual algorithm used in LAMMPS for solving the equations is described in

    the paper written by the developer of the code, S. Plimpton (Plimpton, 1995).

    2.3. Modeling of Spherical/Cylindrical Contact

    Different types of contacts were investigated during this study. The tips used

    with the Quasicontinuum Method were cylindrical, with their axis perpendicular to the

    indentation direction, and rigid. They were made of aluminum, using the same potential

    as for the thin film.

    In molecular dynamics, the indenters used were spherical, with a radius R. They

    were modeled following three distinct methods. The first method, similar to the one used

    in the Quasicontinuum Method, was to model the indenter with the same material as the

    film and to keep it rigid. The second method consisted in modeling the indenter with

    carbon atoms in the diamond structure, and keeping it fixed. The last method was to use a

    virtual indenter. Similar to past atomistic works (Lilleodden et al., 2003), a repulsive

    force is applied such that:

    2)()( RrkrF −−= (2.12)

  • 24

    with k a specified force constant (k = 10 N/m²), R the radius of the indenter, and r the

    distance between the atom and the center of the indenter. This type of indenter removes

    the adhesion and friction forces that are applied by the real indenters.

    2.4. Interatomic Potentials

    2.4.1. Embedded Atom Method Potentials

    Interactions amongst atoms for numerical applications are represented using an

    interatomic potential. The Embedded Atom Method (EAM) potentials (Daw and Baskes,

    1983; Daw and Baskes, 1984) accurately represent defects, surfaces and impurities in

    FCC metals.

    In this method, the total energy of a monoatomic system is represented as

    (Mishin et al., 1999):

    ( ) ( )∑ ∑+=ij i

    iijtot FrVE ρ21

    (2.13)

    where ( )ijrV is a pair potential as a function of the distance ijr between atoms i and j ,

    and F is the “embedding energy” as a function of the host “density” iρ induced at site i

    by all other atoms in the system. The latter is given by:

    ( )∑≠

    =ij

    iji rρρ (2.14)

  • 25

    ( )ijrρ being the “atomic density” function. The second term in equation (2.13) is volume

    dependent and represents, in an approximate manner, many-body interactions in the

    system.

    When creating a new potential, one needs to make sure it will represent very

    closely the real interactions between atoms. EAM potentials can be fitted to experimental

    data for the values of equilibrium lattice parameter 0a , the cohesive energy 0E , three

    elastic constants and the vacancy formation energy fvE . This basic set of properties can

    often be complemented by other data such as planar fault energy or phonon frequencies.

    But those parameters are usually not enough in order to create a reliable potential. The

    created potential is thus fitted to ab-initio calculations for the given material. Ab-initio

    calculations can represent the electronic structure of the atoms, and use laws of quantum

    physics instead of empirical potentials. The aim of the creator of a potential is thus to fit

    the created curve very closely with the ab-initio curve.

    2.4.2. Tersoff Potential

    A Tersoff potential was used in order to represent the interactions between

    carbon atoms for a diamond indenter. This potential is a 3-body potential, and the energy

    of the system of atoms is computed as:

    ∑ ∑≠

    =i ij

    ijVE 21 (2.15)

    ( ) ( ) ( )[ ]ijAijijRijCij rfbrfrfV += (2.16)

  • 26

    ( )

    DRr

    DRrDR

    DRr

    DRrrfC

    +>

    +

  • 27

    2.4.3. Morse Potential

    A Morse potential was used in order to represent the interactions between carbon

    atoms from the indenter and aluminum atoms from the film. The Morse potential

    computes pairwise interactions with the formula:

    ( ) ( )[ ] crrrr rreeDE

  • 28

    where α and β are the Cartesian coordinates, ωi0 is the undeformed atomic volume of the

    ith atom, Det[Fiαβ] is the determinant of the deformation gradient, j is the interatomic

    potential, and rij is the distance between ith and jth atoms. Note that the kinetics terms have

    been eliminated in (2.24) as compared to (Lilleodden et al., 2003). In this equation, the

    use of the determinant of the deformation gradient has been shown to provide improved

    accuracy for the calculation of deformed atomic volumes. Furthermore, the principal

    shear stress, τ, was calculated for each atom using the components of the atomic-level

    stress tensor iσ~ using the formula in (Johnson et al., 1971).

    2.6. Voronoi Construction

    In order to model polycrystalline thin films, grains have to be created. One

    method often used to create those grains is called the Voronoi method (Voronoi, 1908).

    The Voronoi construction enables the user to create a 2D or 3D grain boundary network

    that is considered to be representative of a natural grain boundary (GB) network.

    The user must start by randomly placing reference points at a specified mean

    distance from each other in the surface or volume studied. Each reference point will be

    the center of a grain, for which a random orientation can be assigned. Each grain will be

    composed of atoms that are closer to the reference point of the grain than to the other

    reference points in the surface/volume. This creates grain boundaries that are orthogonal

    to the lines joining the reference point to neighbor reference points. Fig. 2.1 below is an

    illustration of the result of the procedure for a 2D case.

  • 29

    Fig. 2.1: Schematics of the Voronoi construction of a 2D model

    2.7. Tools for the Visualization of Defects, Dislocations and

    Grain Boundaries in Atomistic Simulations

    2.7.1. Centro-symmetry Parameter

    In solid-state systems, the centro-symmetry parameter is a useful measure of the

    local lattice disorder around an atom and can be used to characterize whether the atom is

    part of a perfect lattice, a local defect (e.g. a dislocation or stacking fault) or at a surface.

    This parameter is computed using the following formula (Kelchner et al., 1998):

    ∑=

    ++=6,1

    26

    iii RRPrr

    (2.25)

    where the 12 nearest neighbors are found and iR and 6+iR are the vectors from the

    central atom to the opposite pair of nearest neighbors. This formula was projected in the

  • 30

    plane for use with the QC method. An atom in perfect FCC lattice will then have a

    centro-symmetry parameter of zero. The values for other configurations depend on the

    material chosen. For aluminum, those values are 32.8 Å2 for a surface atom, 8.2 Å2 for

    atoms in an intrinsic stacking fault, and 2.05 Å2 for atoms halfway between fcc and hcp

    sites (in a partial dislocation). The values were 16.4 Å2, 4.1 Å2 and 1.025 Å2 respectively

    for the QC method.

    The centro-symmetry parameter is well adapted for calculations with no

    temperature involved, because at higher temperatures, the distance between atoms is

    constantly changing. An average of the centro-symmetry parameter is then needed in

    order to facilitate the visualization.

    The color scheme used is this study is the following, unless otherwise indicated:

    atoms in a perfect FCC lattice are colored in grey or are not shown at all, those with a

    HCP structure or representing a stacking fault are in blue, and all other non-coordinated

    atoms are in green or red.

    2.7.2. Ackland Parameter

    In contrast to the centro-symmetry parameter, the method using the formulation

    by Ackland (Ackland and Jones, 2006) is stable against temperature boost, because it is

    based not on the distance between the atoms, but the angles. Therefore, statistical

    fluctuations are averaged out a little more. This parameter classifies atoms depending on

    the closest crystallographic structure it belongs to (BCC, FCC, HCP or unknown).

  • 31

    The procedure (Ackland and Jones, 2006) first calculates the mean squared

    separation ∑=

    =6,1

    220 6

    jijrr for the nearest six particles to atom i . It then finds the closest

    neighbors that verify 202 55.1 rrij < . For each of the neighbor pairs found, it calculates the

    bond angle cosines jikθcos . The procedure then relies on a table given by the author, that

    separates the [-1;1] range of possible values for a cosine into 8 ranges. Depending on the

    crystallographic structure around the atom, there are a certain number of bonds within the

    ( ) 2100 −NN closest neighbors that should fall into each range of cosine. The number of

    cosines that should fall into the first range (from -1.0 to -0.945) is called 0χ , and the

    number of cosines that should fall into the last range (from 0.795 to 1.0) is called 7χ . By

    comparing the values obtained for each iχ with the given values of the table, one is able

    to determine to which category atom i can be assigned. The color scheme used in the rest

    of the study is the same than for the centro-symmetry parameter.

    2.8. Validity of the 3D Models

    Periodic boundary conditions are needed for the molecular dynamics 3D models.

    In order to verify that the periodicity of the model does not influence the results, two

    nickel thin films of thickness 12 nm were modeled. The first one has dimensions of 40

    nm ä 12 nm ä 40 nm, and the second one has dimensions of 60 nm ä 12 nm ä 60 nm. A

    virtual indenter of radius 9 nm applies a repulsive force on the surface. The EAM

    interatomic potential for Ni from Mishin et al. (Mishin et al., 1999) is used. Both thin

  • 32

    films have the same orientation, which is ]211[ in the indentation direction, ]111[ and

    ]101[ in the directions normal to indentation. The bottom two layers of the films were

    constrained in the indentation direction. Both simulations were performed at 300K, with a

    time step of 5 fs. The indenter was displaced at a rate of 1 m/s, and the atomic positions

    were recorded at 50 ps intervals up to 1250 ps (250,000 time steps).

    The evolution of the contact pressure is presented in Fig. 2.2, and the snapshots

    at the step 250 ps after the yield point are presented in Fig. 2.3. Fig. 2.2 shows that the

    elastic regime for both samples is very similar. The yield point is attained for a lower

    value of indentation depth for the larger sample, but the values of the yield point are very

    close: 29.3 GPa for the small sample, and 30.3 GPa for the larger sample, or a difference

    of only 3%. The mechanisms of plasticity are also similar, with dislocation loops and a

    prismatic loop (Li et al., 2002) (Fig. 2.3). We can then say that our smaller system (40 nm

    ä 12 nm ä 40 nm) not affected by the periodic boundary conditions. This will save some

    computation time, as the larger system has over 4 million atoms, and the smaller one only

    1.7 million.

  • 33

    Fig. 2.2: Evolution of the contact pressure as a function of the indentation depth for both models.

    The arrows indicate the yield points.

    Fig. 2.3: Indentation step 250 ps after the yield point (a) for the 40 nm ä 12 nm ä 40 nm model

    and (b) the 60 nm ä 12 nm ä 60 nm model. The atoms in perfect FCC lattice have been removed.

  • 34

    CHAPTER 3: EFFECTS OF A GRAIN BOUNDARY

    NETWORK ON INCIPIENT PLASTICITY DURING

    NANOSCALE CONTACT1

    3.1. Objectives

    In this chapter, the effects of a nanocrystalline grain boundary network on the

    plasticity of thin films are studied in aluminum, along with the mechanisms associated

    with plasticity in polycrystalline thin films. Single crystalline and polycrystalline

    simulations with different grain sizes and different potentials are compared. We first

    investigate how to define the yield point for a polycrystalline simulation. We then study

    the mechanism of shear localization in polycrystals, and finally compare the grain

    boundary mechanisms of plasticity for different interatomic potentials for Al.

    1 The results presented in this chapter also appeared in the following journal articles: - V. Dupont and F. Sansoz, Quasicontinuum Study of Incipient Plasticity under Nanoscale Contact in Nanocrystalline Aluminum, Acta Materialia, in press, doi:10.1016/j.actamat.2008.08.014. - F. Sansoz and V. Dupont, Atomic Mechanism of Shear Localization during Indentation of a Nanostructured Metal, Materials Science and Engineering C 27 (2007) 1509.

  • 35

    3.2. Model

    Multiscale computer models of wedge-like cylindrical nanoindentations in 200

    nm-thick Al films were created using the quasicontinuum method. In this study, the

    region subjected to small deformation gradients outside the plastic zone was treated by

    finite elements with an atomistically-informed elastic behavior, while the contact region

    at the interface between the indenter and film surface was fully represented by individual

    atoms. The film dimensions were 400 nm × 200 nm × 0.286 nm, and the size of the full

    atomistic zone was 50 nm × 25 nm × 0.286 nm, as indicated in Fig. 3.1a for a

    nanocrystalline thin film with a grain size of 7 nm. Plane-strain contact was modeled by

    displacing a single crystal Al cylinder with a radius of 15 nm along the direction normal

    to the top surface of the film. The indenter was oriented along the crystallographic

    directions shown in Fig. 3.1b and kept completely rigid during the simulation. The single

    crystalline model had a ]011[ out-of-plane orientation and a [111] surface.

    The polycrystalline structure of the film was constructed as follows. Reference

    atoms were placed randomly in the sample at an average distance equal to a pre-defined

    grain size. GBs were created by a Voronoi construction, which was based on a

    constrained-Delaunay connectivity scheme. Starting from the reference atom, all atoms in

    the grains were added using the Bravais lattice vectors. The mean grain sizes studied

    were 5 and 7 nm. Each grain was assigned a common tilt axis along the ]011[ direction,

    and random in-plane orientation. To avoid discontinuities in the energy state during force

    minimization, the continuum/atomistic frontier was modeled as a single crystal interface

  • 36

    whose crystallographic orientations are shown in Fig. 3.1a. We note that no significant

    atomistic activity was found near this interface, indicating that the plastic deformation

    was limited to the polycrystalline region during the simulations.

    Fig. 3.1: Quasicontinuum model of a 7 nm-grain size Al thin film indented by a 15 nm radius

    cylindrical indenter. (a) Full view of both finite element domain and atomistic region. (b) Close-

    up view of full atomistic zone near the contact region in unrelaxed configuration.

    The bottom of the film was fixed along each direction, while both sides of the

    model were left free. A spacing of 10 Å was initially imposed between the tip apex and

    the film surface. Periodic boundary conditions were imposed along the out-of-plane

    direction in the entire model. Therefore, this study focuses on a randomly oriented 2D

    columnar microstructure. A caveat here is that the plastic deformation processes in this

    type of microstructure may differ from those found in fully 3D polycrystalline structures.

  • 37

    The total energy was minimized by conjugate gradient method until the addition

    of out-of-balance forces over the entire system was found less than 10-3 eV/Å. The

    sample was first relaxed under zero pressure condition in order to obtain the lowest state

    of energy. After relaxation, the atoms of the indenter were displaced by increments of 0.9

    Å until a total displacement of 90 Å, corresponding to an indentation depth of 80 Å

    (increments of 0.4 Å with a final indentation depth of 30 Å for the single crystal). Energy

    minimization was performed between each loading step. The centro-symmetry parameter

    was calculated after each relaxation step to analyze the presence of planar defects in the

    lattice and the structure of the GB network during deformation.

    The semi-empirical EAM potentials for Al by Voter and Chen (Voter and Chen,

    1987) and Mishin and Farkas (Mishin et al., 1999) were used. For brevity in the

    following, these two potentials are referred to as Al-VC and Al-MF potentials,

    respectively. For each potential, we adopted the quasicontinuum procedures used

    previously by Sansoz and Molinari (Sansoz and Molinari, 2005; Sansoz and Molinari,

    2004) to calculate the generalized planar and stacking fault energy curves and the GB

    energy of three Σ tilt bicrystals, including Σ3(112), Σ9(221) and Σ11(113)

    symmetric tilt GBs.

  • 38

    3.3. Characterization of the EAM Potentials

    The generalized stacking and planar fault energy curves for the Al-MF potential

    are shown in Fig. 3.2. The unstable stacking fault energy (γUSF), stacking fault energy

    (γSF) and unstable twinning fault energy (γUTF) are also indicated in this figure. As pointed

    out by Van Swygenhoven et al. (Van Swygenhoven et al., 2004), the ratios γSF/γUSF, and

    γUTF/γUSF are most important in assessing the activation energy required to predict the

    nucleation of stacking faults, full dislocations or twins in the material. More specifically,

    if γSF/ γUSF is close to 1, the activation energy to create a trailing partial is low, which

    favors the nucleation of full dislocations. In contrast, if this ratio is closer to 0, the

    activation energy is too high, which decreases the propensity to nucleate trailing partials,

    leaving only stacking faults after propagation of the leading partial dislocations. The

    same reasoning can be made for twinning, which is more likely to occur if the ratio

    γUTF/ γUSF is close to 1. The energy values obtained for both potentials are summarized in

    Table 3.1, along with reference values from first principles simulations for pure Al (Murr,

    1975; Bernstein and Tadmor, 2004) and Al with either H or Ge solute impurities (Lu et

    al., 2002; Qi and Mishra, 2007). In this table, we find that the calculated energy values

    are significantly smaller for the Al-VC potential than the Al-MF potential, which is

    consistent with the predicted values in the literature (Mishin et al., 1999). Similarly, the

    first-principle values for γSF and γUSF are found to be lower when adding solute impurities

    to Al. Therefore, our finding is the predicted tendency that the stacking and planar fault

    energy values calculated from the Al-MF potential are consistent with the ab-initio values

  • 39

    for pure Al, while the Al-VC results seem to be in better agreement with the energy

    values for Al with impurities. In addition, we find that all the ratios γSF/ γUSF and

    γUTF/ γUSF are similar and equal to 0.81-0.86 and 1.30-1.32, respectively, which suggests

    the same slip and twinning behavior regardless of the interatomic potential for Al.

    Fig. 3.2: Generalized stacking and planar fault energy curves obtained by quasicontinuum method

    with the Mishin-Farkas EAM potential for Al.

  • 40

    Table 3.1: Stacking fault energy (γSF), unstable stacking fault energy (γUSF), unstable twinning

    fault energy (γUTF) and GB energy (γGB) for three Σ tilt symmetric grain boundaries calculated

    from quasicontinuum method on the two EAM Al potentials investigated, and reference values

    from first-principles simulations for pure Al and Al with impurities. All units of energy are in

    mJ/m².

    Present atomistic study First-principles simulations Energy

    EAM Al-VCa EAM Al-MFb Pure Al Al with impurities

    γUSF 93.04 166.71 175-224c 97d

    γSF 75.41 144.22 120-166c,e 73d, 82f

    γUTF 120.99 220.72 207g -

    γSF/γUSF 0.81 0.86 0.73-0.9 0.75 γUTF/γUSF 1.30 1.32 0.92-1.18 -

    γGB − Σ9(221) 302 454 408h -

    γGB − Σ11(113) 96 151 190-206i,j 96j

    γGB − Σ3(112) 318 355 426i -

    a(Voter and Chen, 1987) ; b(Mishin et al., 1999) ; c(Ogata et al., 2002; Hartford et al., 1998; Lu et al., 2000) ; dAl + 14.3 at. % solute H impurities (Lu et al., 2002) ; e(Murr, 1975; Rautioaho, 1982; Westmacott and Peck, 1971); fAl + 3.3 at. % solute Ge impurities (Qi and Mishra, 2007); g(Bernstein and Tadmor, 2004); h(Inoue et al., 2007); i(Wright and Atlas, 1994); jAl + 9 at. % substitution Ga impurities (Thomson et al., 2000).

    3.4. Force and Contact Pressure Calculations

    The force applied by the indenter was calculated using the formula:

    ∑∈

    =Zi

    iPP , (3.1)

    where Z represents all atoms of the film belonging to the contact zone and Pi is the out-

    of-balance force on atom i in this zone, projected along the direction of indentation. The

    contact zone was computed after each loading step by only including atoms at the

    indenter-film interface within a separation distance from the tip equal to the potential

  • 41

    cutoff radius. The cutoff radii were 5.555 Å and 6.287 Å for the Al-VC and Al-MF

    potentials, respectively. The mean contact pressure was determined at each step as:

    perioza

    PH×

    =2

    , (3.2)

    where a is the contact length, defined as half the width of the projected contact area, and

    zperio is the thickness of our sample in the out-of-plane direction.

    3.5. Determination of the Yield Point

    The onset of plasticity in single crystals is characterized by a pop-in event

    corresponding to the homogeneous nucleation of dislocations in the crystal. This is

    visible on the force – displacement plot because it corresponds to a sudden drop in the

    force (Fig. 3.3), which corresponds to the yield point. In a nanocrystalline thin film,

    however, there is no visible drop on the plot, and rather, we find flow serration leading to

    a significant softening effect. The flow serration has been found to correspond to plastic

    shear localization through the formation of shear bands (Trelewicz and Schuh, 2007;

    Lund and Schuh, 2004). As can be seen on Fig. 3.3, the yield point for a polycrystalline

    material cannot be found from the force – displacement curve directly.

  • 42

    Fig. 3.3: Effect of nanosized grains on the nanoindentation response of Al substrates from

    molecular static simulation using the Al-VC potential. The indenter radius is 15 nm. Serrated

    plastic flow clearly appears in the two nanocrystalline Al substrates under nanoindentation.

    The following criterion was used to quantitatively assess the onset of plasticity as

    obtained by the simulations. We computed the theoretical pressure – displacement curve

    corresponding to an isotropic elastic surface in contact with a rigid cylinder by using the

    following equation from continuum theory (Johnson, 1985):

    ( )[ ])1()2ln(21

    2

    ννπ

    νδ −−−= atE

    F , (3.3)

    where F is the total linear force in N/m ( = P/zperio), δ is the indentation depth, t is the film

    thickness along the direction of indentation, E and ν are the Young’s modulus and

  • 43

    Poisson’s ratio of the film, respectively. Substituting this into (3.3) gives the mean

    contact pressure for an elastic surface He:

    ( )δ

    ννν

    π×

    ⎥⎦

    ⎤⎢⎣

    ⎡−

    −⎟⎠⎞

    ⎜⎝⎛−

    =

    12ln212 2ata

    EH e ,(3.4)

    Fig. 3.4: Contact pressure versus penetration depth plot for a 7-nm polycrystal with the Al VC

    potential, along with the corresponding theoretical fitting.

    Since the above equation depends on the contact length a, which changes with

    the penetration δ, the parameter He was re-evaluated at each loading step. Assuming ν =

    0.345 for polycrystalline Al (Meyers and Chawla, 1999), the isotropic Young’s modulus

    was determined by fitting the elastic contact pressure He from (3.4) to the first portion of

    the curve obtained by quasicontinuum simulation (Fig. 3.4). Using the Al-VC potential,

  • 44

    we see that the yield point is around 2.3 GPa, and that the curve shows significant

    softening effects.

    3.6. Shear Localization Mechanisms

    Shear banding was found to occur in all polycrystalline simulations, and is best

    illustrated by a 5 nm-grain-size simulation using the Al-VC potential (Fig. 3.5). At the

    onset of plasticity, significant grain boundary sliding is found to occur. This behavior

    results in significant rotational deformation of the grains with limited intragranular slip.

    During this process, the grain boundary structure is significantly changed and, in some

    cases, several grain boundaries tend to be aligned (Fig. 3.5a and b). The bands are formed

    by the sliding of aligned interfaces separating the grains (grains 3 and 4 in Fig. 3.5c).

    When the shear plane encounters a triple junction and is stopped by a grain that is not in

    its alignment, the shear band follows its path by intragranular slip in the prolongation of

    the shear plane. For example, a stacking fault left behind a partial dislocation can be seen

    in grain 2 in the prolongation of the shear plane in Fig. 3.5c. Subsequently, the newly

    created stacking faults are found to nucleate mechanical twins, which grow under the

    applied shear stress. Mechanical twinning has also been observed in nanocrystalline Al

    under indentation by Chen et al. (Chen et al., 2003b). This result suggests therefore that

    our simulation is in excellent agreement with the experimental data.

  • 45

    Fig. 3.5: Thin shear band formation in 5-nm-grain-size nanocrystalline Al after 2.5-nm-deep

    indentation. (a) Partial view of the contact interface and location of the grain cluster associated

    with the shear band. (b) Enlarged view of the shear plane. A mechanical twin nucleated at the

    triple junction in the prolongation of the shear place is clearly visible in grain 2. (c) Magnitude

    and direction of atomic displacements between two loading increments represented by arrows.

    The shear band results from sliding of aligned grain boundaries (grains 3 and 4) and intragranular

    partial slip (grain 2).

    Shear localization in nanocrystalline metals has been found to occur through

    collective grain activity initiated by grain boundary sliding. It has been shown by

    Hasnaoui and co-workers (Hasnaoui et al., 2002a) on uniaxial compression that three

    mechanisms contribute to the formation of local shear planes by cooperative grain

    activity in nanocrystalline metals: (i) GB sliding to form a single shear plane consisting

    of a number of collinear GBs, (ii) continuity of the shear plane by intragranular slip and

    (iii) the coalescence via reorientation of neighboring grains that have an initially low

  • 46

    angle GB. The present study clearly indicates that during indentation the first two

    mechanisms also co-exist to form shear bands. We believe however that the mechanism

    of grain coalescence via reorientation should occur at larger depth of indentation after