HETEROGENEOUS DEFORMATION IN POLYCRYSTALLINE COMMERCIAL PURITY TITANIUM By Yiyi Yang A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Materials Science and Engineering 2011 ABSTRACT HETEROGENEOUS DEFORMATION IN POLYCRYSTALLINE COMMERCIAL PURITY TITANIUM By Yiyi Yang Heterogeneous deformation is commonly viewed as a precursor to damage nucleation. Only if the heterogeneous deformation is reliably modeled can location where cracks form be properly predicted. In this research project, commercial purity titanium (hexagonal close-packed (hcp) metal) has been used as a model material to study different scales of heterogeneous deformation in four-point bending specimens and fatigued-to-crack specimens. A series of comprehensive experimental and modeling methods have been conducted to characterize the heterogeneous deformation within different microstructure patches. Heterogeneous deformation within a microstructure patch was quantified by a new combined technique using atomic force microscopy (AFM), backscattered electron (BSE) imaging, and electron-backscattered diffraction (EBSD). Local shear distribution maps derived from z-displacement data of slip steps and/or twins measured by AFM were directly compared to results of a crystal plasticity finite element (CPFE) simulation that incorporates a phenomenological model of the deformation processes to evaluate the ability of the CPFE model to match the experimental observations. The CPFE model successfully predicted most types of active dislocation slip systems within the grains at the correct magnitudes, but the spatial distribution of the strains within grains differed between the measurements and the simulation. Since the difficulty of directly measuring the critical resolved shear stress (CRSS) of different deformation systems in   titanium is a major cause of the deficiency of the match between CPFE modeling results and experimental measurements, nano-indentation technique is introduced to help study the values of CRSS and saturation resistance of different deformation systems for commercial purity titanium. Arrays of conical nano-indentations were performed on patches of grains. AFM scans and EBSD measurements on the patches indicate that the surface pile-up topography was strongly crystallographically dependent. By optimizing the modeling results of the nano-indentation behavior in different grains, a more reliable set of values for CRSS and saturation resistance has been determined using a non-linear optimization procedure. On a larger scale, a novel approach including BSE imaging and discrete Fourier transformation (DFT) processing is established to determine plastic zone shape and to measure plastic strain distribution within the plastic zone around a fatigue crack tip. The measured plastic zone size is in good correlation with the theoretically predicted zone sizes. This approach also holds promise to measure local heterogeneous deformation in other kinds of deformed samples. This work has helped identity the needs for future experiments focused on the observation of dislocation activity near grain boundaries using ECCI, and a statistical study on the relationship of heterogeneous deformation in neighboring grains and damage nucleation at grain boundaries. Future simulation work should focus on developing a more comprehensive CPFE model, considering the information of slip/twin transfer and grain boundaries resistance, in order to simulate heterogeneous deformation within grain patches more reliably.   Copyright by YIYI YANG 2011   ACKNOWLEDGEMENTS I would like to express my deepest gratitude to my advisor, Dr. Martin Crimp, and other committee members, Dr. Thomas Bieler, Dr. Carl Boehlert, Dr. James Lucas, and Dr. Viktor Astakhov. Their wisdom, knowledge, and commitment to the highest standards inspired and motivated me. I would like to thank Dr. Philip Eisenlohr and many other collaborators in Max-Planck-Institut für Eisenforschung GmbH for their invaluable discussions and suggestions. I would also like to show my special gratitude to Dr. Baokang Bi for his assistance with the AFM work, Dr. Zambaldi for the help on the simulation work of nano-indentation, Dr. Rachel Tomlinson, Dr. Eann Patterson, and Mark Dawson for their help with BSE-DFT work, and all of those who supported me in any respect during the completion of the project. Most especially, I owe my deepest gratitude to my parents and Zhongping who have been always supporting, encouraging and believing in me. Most of the research work in this dissertation has been conducted with funding from The National Science Foundation (grant #DMR-0710570) and Deutsche Forschungsgemeinschaft (grant #EL-681/2-1).   v   TABLE OF CONTENTS LIST OF TABLES ............................................................................................................ ix LIST OF FIGURES ........................................................................................................... x LIST OF SYMBOLS AND ABBREVIATIONS ................................................................ xvi CHAPTER ONE INTRODUCTION ..............................................................................................................1 1.1 Importance of and Challenges for Measuring and Simulating Heterogeneous Deformation and Predicting Micro-crack Initiation ..................................................1 1.2 Basic Approaches to Understanding Heterogeneous Deformation ........................3 1.2.1 Slip Transfer Criteria ......................................................................................3 1.2.2 Geometrically Necessary Dislocations...........................................................6 1.3 Experimental Characterizations and Simulations on Heterogeneous Deformation 9 1.3.1 Experimental Characterizations .....................................................................9 1.3.1.1 Single crystals and bi-crystals...........................................................10 1.3.1.2 Poly-crystals......................................................................................14 1.3.2 Simulations of Heterogeneous Deformation ................................................18 1.4 Current Understanding of the Relationship between Heterogeneous Deformation and Damage Nucleation........................................................................................23 1.5 Studies on Commercial Purity Titanium and Other Hexagonal Materials .............27 1.5 Motivations for This Study.....................................................................................32 1.6 Overview ...............................................................................................................34 CHAPTER TWO MATERIAL, EXPERIMENTAL DETAILS, AND SIMULATION FRAMEWORK ..............37 2.1 Material .................................................................................................................37 2.1.1 Material A .....................................................................................................37 2.1.2 Material B .....................................................................................................40 2.2 Sample Preparation ..............................................................................................40 2.3 Experiment Details................................................................................................45 2.3.1 Scanning Electron Microscopy.....................................................................45 2.3.1.1 Electron Imaging ...............................................................................45 2.3.1.2 EBSD/OIM measurement .................................................................45 2.3.2 Four-point Bending Testing..........................................................................49 2.3.3 Atomic Force Microscopy.............................................................................49 2.3.4 Nano-indentation..........................................................................................52 2.3.5 Fatigue Testing and Thermoelastic Stress Analysis Measurement .............55 2.4 Slip/twin Plane Trace Analysis..............................................................................56 2.5 Constitutive Law in Crystal Plasticity Finite Element Method ...............................57   vi   CHAPTER THREE MEASURING AND CPFEM SIMULATING MICRO-SCALE HETEROGENEOUS DEFORMATION .............................................................................................................60 3.1 Microstructure in a Four-point Bending Sample....................................................61 3.1.1 Slip Lines and Twins ....................................................................................61 3.1.2 Grain Boundary Ledge Formation................................................................61 3.1.3 A Selected Microstructure Patch for Further Studies ...................................64 3.2 Quantitative Measurement of Heterogeneous Deformation within a Grain Patch 71 3.2.1 AFM Measurement and Accuracy Check Using T1 Twins...........................71 3.2.2 Generation of Local Shear Distribution Maps ..............................................75 3.2.3 Local Shear Distribution Evolution at Different Strain Level ........................77 3.3 Crystal Plasticity Finite Element Modeling ............................................................81 3.3.1 Constitutive Law...........................................................................................81 3.3.2 Mesh Generation and Boundary Condition ..................................................82 3.3.3 Modeling Results..........................................................................................85 3.4 Discussion.............................................................................................................88 3.4.1 Comparison of Experimental Observations and CPFE Simulations ............88 3.4.2 Effect of Modeled Area Size in Simulation Results ......................................90 3.4.3 Effect of Rim Crystal Orientation in Simulation Results ...............................91 3.5 Conclusions ..........................................................................................................93 CHAPTER FOUR CALIBRATING PARAMETERS FOR CPFE MODELS USING NANO-INDENTATION .95 4.1 Importance and Advantages of Measuring CRSS Using Nano-indentation..........96 4.2 Grain Orientation Dependent Behavior during Nano-indentation .........................98 4.2.1 Relationship between Hardness and Grain Orientation ...............................98 4.2.2 The Relationship between Pile-up Topographies and Grain Orientation ...101 4.3 CPFEM Simulation of Indentation and CRSS Optimization................................105 4.3.1 Constitutive law ..........................................................................................105 4.3.2 Mesh Generation and Boundary Conditions ..............................................107 4.3.3 Crystal Plasticity Parameter Optimization Procedures ..............................107 4.3.4 Modeling Results........................................................................................109 4.4 Discussion...........................................................................................................112 4.4.1 Automated Identified Constitutive Parameter Values ................................112 4.4.2 Comparison of Simulation Results using Two Sets of CRSS ....................115 4.5 Conclusions ........................................................................................................117 CHAPTER FIVE MEASURING MESO-SCALE HETEROGENEOUS DEFORMATION IN COMMERCIAL PURITY TITANIUM ......................................................................................................118 5.1 Observation of Microstructure around a Fatigue Crack Tip using BSE Images..119 5.2 Image Analysis Technique Using DFT................................................................122 5.2.1 Basic Information of DFT ...........................................................................122 5.2.2 DFT Image Data Processing......................................................................122 5.3 Quantifying the Relation between FWHM and Strain Using In-situ Tensile Test 127 5.4 Residual Strain Map around a Crack Tip in the Fatigue Specimen ....................131   vii   5.5 Discussion...........................................................................................................131 5.5.1 The Influence of Brightness, Contrast, Magnification, and Calibration Procedures on the Strain Map ............................................................................131 5.5.2 Comparison between the Plastic Zone Size Measured by BSE-DFT Method and Those Proposed by Other Works.................................................................137 5.5.3 The Application of BSE-DFT Method .........................................................143 5.6 Conclusions ........................................................................................................144 CHAPTER SIX BROAD IMPACTS OF THE STUDY ............................................................................146 CHAPTER SEVEN CONCLUSIONS AND FUTURE WORK.......................................................................149 7.1 Conclusions ........................................................................................................149 7.2 Future work .........................................................................................................151 APPENDIX A ................................................................................................................154 APPENDIX B ................................................................................................................157 APPENDIX C................................................................................................................171 REFERENCES .............................................................................................................174   viii   LIST OF TABLES Table 1.1 CRSS of different slip modes reported for α-titanium. Table 1.2 Deformation twinning modes in α-titanium. Table 3.1 Deformation systems in grains within the solid-line box in Figure 3.3(a) and 3.4(a)-(b) at global strain level 1.5%, 3%, and 6%. Table 3.2 Hardening parameters used to describe evolution of the critical resolved shear stress. Table 4.1 The indentation axis in crystal coordinate system, the angle between c-axis and indentation axis (η) of selected indented grains in specimen S0-2, and the maximum residual depth (hres) of the indents in the center of these grains. Table 4.2 Optimized constitutive parameters for prismatic, basal and pyramidal slip systems. The other constant parameters for the three systems in the simulations were chosen as: m = 20, h0 = 200 MPa, a = 2.   ix   LIST OF FIGURES Figure 1.1 Three-dimensional slip transfer geometry at grain boundaries. Figure 1.2 If each grain of a poly-crystal deforms in a uniform manner, overlap and voids appear (b). This can be corrected by introducing geometrically necessary dislocations, shown in (c) and (d). Figure 1.3 (a) Orientations of the two crystals in relation to the compression direction Z. GND distribution in a 10% (b) and a 30% (c) strained aluminum bicrystal. Figure 1.4 A schematic drawing of the setup of 3D-XRD. Figure 1.5 Two vertical XRD scans (line 1 and 2) compared with the corresponding EBSD scan data. All curves display the total misorientation relative to the base orientation of the pillar plotted as a function of distance from the pillar base. Figure 1.6 (a) SACP composite from the deformed grain (~1% strain) shown in (b). The ‘‘×’’ indicates the electro optic axis at zero sample tilt. (b) Channeling contrast image of the deformation defects in a grain of TiAl with the calculated {111} plane traces. Arrowheads indicate bands of dislocations piled up in the neighboring grain at points of intersection of deformation twins with the grain boundary. Figure 1.7 The various mesh types used in Ref. [61]: (a) real 3D mesh (31,780 nodes), (b) real 3D mesh (130,818 nodes), (c) quasi-3D mesh (14, 076 nodes). Figure 1.8 (a) Backscattered electron image of γ-γ TiAl grain boundary damage initiation due to twin interaction with the grain boundary; (b) 2D schematic of crack nucleation at a concentrated band of slip or twinning. Figure 1.9 Slip systems for α-titanium: (a) prismatic slip, (b) basal slip, (c) pyramidal slip, and (d) pyramidal slip. The arrows only represent the slip directions, not the magnitudes of the dislocation slip. Figure 1.10 Twinning systems for α-titanium: (a) T1 twinning, (b) T2 twinning, (c) C1 twinning, and (d) C2 twinning. The arrows only represent the twinning directions, not the magnitudes of the twinning. Figure 2.1 {0001} and {10 10} pole figures of (a) specimen S0, and (b) specimen S45.   x   Figure 2.2 Specimen orientations in Material A plate (left), and the size of specimens for four-point bending tests and nano-indentation tests (right). Figure 2.3 {0001} and {10 10} pole figures of Material B. Figure 2.4 Size and shape of specimens for in-situ tensile tests cut from Material B. Figure 2.5 Size and shape of specimens for fatigue tests cut from Material B plate. Figure 2.6 Backscattered electron image of a microstructure patch with deposited fine tungsten cross patterns. Figure 2.7 A typical setup of an EBSD/OIM system in SEM. Figure 2.8 Deformation stage for in-situ tensile test. Figure 2.9 Four-point bending stage with a bended specimen loaded. Figure 2.10 Schematic drawing (a) and an SEM image (b) showing the size and shape of the AFM probe tips used in this project. Figure 2.11 BSE image of the indented microstructure patch in specimen S0-2. Figure 2.12 BSE image of the indented microstructure patch in specimen S0-3. Figure 3.1 A typical microstructure patch from S45-1 at 1.5% strain. (a) shows a grain with dense slip lines, while (b) shows a grain with much less slip lines. Twins also shows in grains, as illustrated in (c). Figure 3.2 Secondary electron images show ledges (circled) along grain boundaries in S0-1under 8% global strain; (b) was taken of the same microstructural patch with the whole sample rotated 180 degrees from (a). Circled boundaries showing reverse contrast in two images, indicating these are ledges and not cracks. Figure 3.3 AFM “Section” scans shown ledges with different topography at grain boundaries in S0-2 under 1.5% global strain, indicating that heterogeneous deformation occurs in small deformation. Figure 3.4 Microstructure patch of specimen S45-1 at ~1.5% tensile strain. The patch within the solid outlined box was quantitatively analyzed using AFM-based method, and the patch within the dashed outlined box was modeled by CPFEM. Inverse pole figures of the grain orientation changes within the solid outlined box before and after deformation are shown in (b) and (c), respectively. Directions of slip lines and twins in each grain are highlighted by white lines. These activated deformation systems were identified using trace analysis based on EBSD data.   xi   Figure 3.5 The same microstructure patch in S45-1 specimen as shown in Figure 3.4(a) following further deformed to ~3% and ~6% tensile strains, shown in (a) and (b) respectively. The patch within the solid outlined box was quantitatively analyzed using AFM-based method. The growth of five T1 twins in grain 2 with the increase of the global strain can be easily observed. Furthermore, secondary deformation systems, highlighted by white solid lines in (b), were observed in several hard grains, such as grain 2, at 6% strain level. Figure 3.6 BSE image illustrating the area characterized using AFM and CPFEM at 1.5% global strain. Different activated deformation systems characterized using trace analysis are denoted by color lines. Figure 3.7 Examples of surface topography in grain 2 (left) and grain 3 (right) measured by AFM after deformation. Line section profiles reveal that much larger and more homogeneous surface steps result from twins (left) than from slip lines (right). Dashed line is interpreted as the undistorted surface inclination and serves as the basis for evaluating the overall height change along the section. Figure 3.8 Schematic representation of a cross-section through a twinned portion of a grain. The twin trace on the sample surface is normal to the sketch (i.e., parallel to the Y direction). The surface displacement h due to twinning is determined from an AFM line section in the X direction. tx and ttrue are the twin width in the X direction (projected twin width) and the twin thickness along the twin plane normal direction, respectively. Figure 3.9 Schematic drawing showing the parameters associated with calculation of the local average shear caused by the activated deformation system with the slip plane normal n and Burgers’ vector direction b in sample coordinate system. h is the surface height change caused by this deformation system. Xm(n-1), Xmn, and Xm(n+1) are reference grid increments in the x direction. Figure 3.10 The calculated local shear deviation at 1.5% global strain based on the measurements using both horizontal and vertical AFM “Section” lines for 50 randomly chosen tiles. The measurements agree within about 10% in both directions for all tiles. Figure 3.11 The AFM-based experimentally measured sum of all individual local shear distribution map of the highly characterized microstructure patch shown in Fig. 3.4(a) at global strain level of (a) 1.5%, (b) 3%, and (3) 6%. Figure 3.12 Individual local shear associated with different types of deformation systems in the microstructure patch shown in Fig. 3.4(a). The experimental measurements show that, in the early deformation stage, the local shear   xii   increases almost proportional to the overall tensile strain in all grains, while during further deformation (3% to 6%), the local shear in grains with pyramidal slip shows a much larger increase rate than that in grains with prism or basal slip. Figure 3.13 (a) – (c) Three different sizes of 3D meshes of the grain patch corresponding to the microstructure within the dashed box in Fig. 3.4(a). Figure 3.14 (a) – (c) The summed individual shears for the top surface resulting from simulated tensile loading to 1.5% engineering strain using the three meshes shown in Figure 3.13(a) – (c), respectively. Figure 3.15 The comparison of the summed individual shears distribution on the frontside and back-side of the meshes shown in Figure 3.13(b). Figure 3.16 Comparison of experimental measured (shown in row (a)) and FEM simulated (using mesh (c)), shown in row (b)) individual local shear associated with different types of deformation systems in the microstructure patch shown in Figure 3.4(a) at 1.5% strain. FEM simulation results illustrate that the spatial distribution of local shear caused by prismatic and basal slip agrees well with experimental measurement. Most of the local shear caused by T1 twinning is observed in grain 2. Simulation does not successfully simulate the pyramidal slip in grain 10 where AFM measurement reveals around 0.01 local shear. Figure 3.17 CPFE model illustrating how the surrounding orientation (indicated with Euler angles) affects local shear distributions at 2.2% global strain. Figure 4.1 Material pile-up (left) and sink-in (right) increases or decreases the contact height, hc, during indentation to the maximum indentation depth, hmax. Figure 4.2 The array of indents applied on the microstructure in specimen S0-2. Indentions were spaced by 20 µm × 16 µm. A semi-transparent OIM map with inverse pole figure coloring is superimposed on a backscattered electron image. Figure 4.3 Color-scaled map of the maximum indentation depth (unit: nm) for the indentation array shown in Figure 4.2. Each tile represents one indentation in the middle of this tile. Figure 4.4 Force-displacement curves from indentations conducted in the middle of soft grains 5 and 10, and hard grain 2. The maximum load is 6 mN. The lower inclination of the curve during loading stage for grain 2 indicates that grain 2 is more resistant to plastic deformation.   xiii   Figure 4.5 Surface topography scanned by AFM for indents in soft grain G7 (left) and hard grain G2 (right) indicating the grain orientation dependent behavior of pile-ups. Figure 4.6 Pile-up topographies measured by AFM positioned in the inverse pole figure (IPF) according to the corresponding indentation axes. The indents are rotated according to the convention defined in [117]. The indents in the upper half of the IPF are also displayed in the lower half by a mirroring operation (improper rotation) with respect of (1 0 -1 0) plane and marked there with an asterisk. Figure 4.7 A deformed (indented) finite element mesh of hexahedral elements. Figure 4.8 Experimental, CPFE simulated surface topographies, as well as the Δzi, j map, of indents 6-a and 13-o, which were selected for the identification of the constitutive parameters. The size of each map is 2.5 µm × 2.5 µm. Figure 4.9 Experimental, CPFE simulated surface topographies, as well as the Δzi, j map, of indents 6-o and 13-i, which were selected for the validation of the set of optimized parameters. The size of each map is 2.5 µm × 2.5 µm. Figure 4.10 Twenty-two simulated pile-up topographies measured by AFM positioned in the IPF according to their indentation axes. The pile-ups are shown in gray scale, and the regions below zero are shown in blue. The contour lines are starting at -0.04 µm and 0.04 µm for below and above zero regions, and drawn every 0.04 µm. Similar trends of pile-up features were found in CPFE simulations as that shown in the experimental measurements. Figure 5.1 The around-crack-tip region was divided by a 15 × 15 (225 in total) grid as illustrated, and each BSE image was taken from the middle part of each tile. Figure 5.2 Upper array (shown in Figure 5.1) of backscattered electron images around the crack tip at a center-to center spacing of 187 µm. The images in nearcrack-tip region show more lattice rotation and contrast variation, indicating more residual plastic deformation. Figure 5.3 (a) BSE images h5, h8, h11, and h14 from Figure 5.2; (b) their corresponding DFTs and (c) a horizontal line profile across the center of each DFT and showing a decrease in the center peak width with increasing distance from the crack tip to the image location. Figure 5.4 Schematic plots showing the process of fitting line profiles and measuring the full width at half the maximum of the profile. Line profiles were first   xiv   smoothed using an adjacent-averaging five-point window. The left half of a smoothed profile was then fitted to an exponential curve. Figure 5.5 The full width at half maximum of the FFT center peak versus the distance from the crack tip at (a) θ = 0° (b) θ = 45° Figure 5.6 Engineering stress-strain curve obtained for the commercially pure titanium; the arrows indicate the strain increments at which the in-situ uniaxial tensile test was interrupted to collect BSE images. Figure 5.7 BSE images of a grain patch at different global strain levels during in-situ tensile test indicating the evolution of microstructure features. Figure 5.8 The full width of half the maximum of the DFT as a function of the engineering strain based on images recorded at increments of plastic strain during an in-situ tensile test. Figure 5.9 Map of plastic strain around the crack tip (the area shown in Figure 5.1) based on the FWHM of the DFTs of BSE images. Figure 5.10 BSE images from the same microstructure patch and their corresponding DFT line profiles showing that the change of brightness does not affect the FWHM of the line profiles. Figure 5.11 BSE images from the same microstructure patch and their corresponding DFT line profiles showing that the change of contrast does not affect the FWHM of the line profiles. Figure 5.12 Phase difference and magnitude of the TSA signal as a function of distance along a line through the tip of the fatigue crack in the direction of crack growth. The extent of the plastic zone found from the phase difference map is superimposed as a black area. Figure 5.13 Map of plastic strain around the crack tip based on the FWHM of the DFTs of images from the region shown in Figure 5.1. The white line shows the measurement of plastic zone using TSA data. The green and blue solid lines indicate the theoretical estimations of the plastic zone size based on Dugdale’s and Irwin’s approaches, respectively.       xv   LIST OF SYMBOLS AND ABBREVIATIONS   3D-XRD Three-dimensional X-ray diffraction AFM Atomic force microscopy b Burgers vector of a given dislocation type BSE Backscattered electron CPFE Crystal plasticity finite element CRSS Critical resolved shear stress DFT Discrete Fourier transformation ez Specimen surface normal EBSD Electron-backscattered diffraction ECCI Electron-channeling contrast imaging fip Fracture initiation parameter F Deformation gradient e F p Elastic part of deformation gradient F Plastic part of deformation gradient FWHM Full width at half maximum fcc Face-centered cubic FEM Finite element method FFT Fast Fourier transformation FIB Focused ion beam Gtopo The topographic contribution to the objective fuction GND Geometrically necessary dislocation h Surface step height   xvi   hcp Hexagonal close-packed KI Applied stress intensity factor p L Plastic velocity gradient m Dislocation slip direction m’ Slip transfer parameter MPODM Multiple point over-deterministic method n Dislcation slip plane normal direction N Number of displaced twin planes OIM Orientation imaging microscopy r Plastic zone radius R Rotation matrix s Shear resistance of a given dislocation type S Schmid matrix for a given dislocation type SACP Selected-area channel pattern SE Secondary electron SEM Scanning electron microscopy t Direction of the traction force acting on a grain boundary tx Projected twin thickness ttrue True twin thickness TEM Transmission electron microscopy TSA Thermal-elastic stress analysis Xmn AFM section line length z Dislocation line direction of a given dislocation type α A given dislocation type   xvii   α Dislocation tensor Δzi, j Height difference between the measured and simulated topography Δε Cyclic strain ΔKI Cyclic stress intensity factor ε Applied strain φ1 ,Φ, φ2 Euler angles γ Slip rate of a given slip system κ Angle between the two Burgers vectors of activated slip systems in two neighboring grains τ Resolved shear stress on a given deformation system ρ Dislocation density of a given dislocation type σy Yield stress of a material ωm , ωn Complex roots of unity Ψ Angle between the two slip plane normals of activated slip systems in two neighboring grains                     xviii   CHAPTER ONE INTRODUCTION 1.1 Importance of and Challenges for Measuring and Simulating Heterogeneous Deformation and Predicting Micro-crack Initiation Heterogeneous deformation and local damage nucleation in metals have been crucial topics for decades. Simulations based on continuum fracture mechanics, assuming an average value of both mechanical and crystallographic properties over whole bulk materials, have been used to model macroscopic properties of bulk materials, such as ductility, fatigue life, and fracture toughness [1-3]. Heterogeneous deformation is commonly viewed as a precursor to damage nucleation. Only if the heterogeneous deformation is reliably modeled can locations where cracks form in the first place then can be properly predicted. Furthermore, from an engineering point of view, knowledge of the micro-crack initiation locations can help to develop processing routes, such as rolling and extrusion, to strengthen metals and alloys by avoiding the development of these locations. Although many experiments and simulations of localized damage nucleation in polycrystalline materials during plastic deformation have been conducted in past decades [4-7], few fundamental breakthroughs in predicting nucleation locations have been achieved due to the complex behavior of grain boundaries. It is generally recognized that cracks prefer to nucleate at discontinuous interfaces, such as grain boundaries [8, 9] and phase boundaries [10, 11], and that the plastic deformation in one grain must transfer to another through grain boundaries in polycrystalline materials   1   during deformation. During this process, some boundaries may initiate micro-cracks easily — these boundaries are called “weak” boundaries, while those where damage nucleation is relatively difficult are referred to as “strong” boundaries. Several factors, however, can influence the behavior of grain boundaries, adding confusion to the distinction between strong versus weak boundaries. The first factor is the orientations of the two grains on the opposite sides of the grain boundary, which strongly influences the transfer of dislocations across the boundary [12], and the kind and amount of dislocations subsequently left in or near the grain boundary. This factor also has influence on the activated deformation systems in these grains. Details concerning this factor will be provided in the following sections. The boundary energy is another factor that should be considered. Miura et al., have shown that the higher the grain boundary energy, the more easily the grain boundary fractures [13]. The boundary energy is, however, closely related to the grain boundary structure, especially the cohesive properties. For example, low Σ boundaries may have more resistance to damage nucleation due to their lower boundary energy and lower ability to absorb residual dislocations than random boundaries [14]. Finally, yet importantly, the strain gradient history in two adjacent grains also plays a crucial role in determining if a grain boundary tends to crack. Recent research in finite element models is seeking ways to effectively model strain gradients [7, 15, 16]. The same microstructure patch can display different strain distributions after experiencing two different sets of deformation processes in simulations, which may directly leads to the crack initiation at different grain boundaries. To date, there has been little work combining all of these factors, leaving general rules for predicting damage nucleation sites in poly-crystals undeveloped.   2   1.2 Basic Approaches to Understanding Heterogeneous Deformation Heterogeneous deformation usually results from two facts [17-19]. One is that some “soft” grains are much more easily deformed than other grains, which means that one or more deformation systems in the “soft” grains can be more easily activated than those in “hard” grains under certain stress states, leading to large strain differences between soft and hard grains. Secondly, equilibrium across the grain boundary requires an additional stress field along the boundary, resulting from the various constraints along the grain boundary of a given grain from different neighboring grains during deformation process. Consequently, these influences from neighboring grains can lead to heterogeneous deformation and associated strain gradients within the given grain. Two basic theories have been reviewed in this section to help better understand the heterogeneity within metallic materials during deformation processes. 1.2.1 Slip Transfer Criteria An important approach to understanding the existence of heterogeneous deformation in polycrystalline materials is the “slip transfer criteria”. During continuous plastic deformation, strain needs to be transferred from one grain to another. This transfer process often leaves residual dislocations in the grain boundaries and changes the directions of Burgers vectors and slip planes of the original dislocations. To identify which slip systems in neighboring grains will be activated during slip transfer processes, the geometry of slip directions and planes, in relation to the misorientation between two grains, has been studied by several authors. Luster et al. [12] presented a simple parameter to describe the slip transfer event at grain boundaries:   3   (1.1) where, as shown in Figure 1.1, κ is the angle of a Burgers’ vector in one grain with respect to the Burger’s vector in the neighboring grain, and Ψ is the angle of a slip plane normal with respect of the normal in the other grain. In addition, the angle between the intersection line of one slip plane at grain boundary and the other intersection line of the other slip plane in a neighboring grain is defined as θ. Luster’s slip transfer criteria are built on three general rules for slip transfer in poly-crystals [20, 22]: (1) cosθ must be at a maximum; (2) the magnitude of Burgers vector of residual dislocations in grain boundary must be at a minimum; (3) the resolved shear stress on the outgoing slip system must be at a maximum. For example, if there are two slip systems that are perfectly aligned, the parameter equals one, which means that the grain boundary is transparent to dislocations (no heterogeneity existing near this boundary). The parameter, m’, becomes zero when the slip planes or the slip directions of two grains are perpendicular, which indicates that the edge dislocations cannot transfer across the grain boundary (cross slip may still occur when the two plane normals are perpendicular to each other). Based on the m’ slip transfer criteria, Ashmawi and Zikry [21] further identified the critical piling-up dislocation density near the boundary for slip transfer into the adjacent grain. Instead of using Luster’s parameter, they use the slip transfer criteria from Clark [22]: (1.2)   4   Figure 1.1 Three-dimensional slip transfer geometry at grain boundaries [6].     5   which provides a framework to predict how heterogeneous deformation evolves as a function of grain orientation and structure. Similar to m’, the smaller the parameter, ε, is, the more difficult it is for dislocations to transfer, and thus the more dislocations will pile up near the boundary (or transfer across with residual dislocation left in the boundary), producing a larger heterogeneous stress field. These simple rules provided some understanding of the heterogeneous strain between adjacent grains near grain boundaries. They are, however, not sufficient to describe heterogeneous deformation, since: (1) the process of transfer in the early stage of plastic deformation affects subsequent deformation between two grains [21, 23]; and (2) it over-simplifies as they only consider one slip system activated in each grain [20]. Also, it is still not clear how the heterogeneous strain is correlated to the damage nucleation. For example, in random boundaries, heterogeneous strain may result in large tensile stresses between two grains, which may lead to damage nucleation [19, 21, 24]. Strongly cohesive boundaries, however, may force heterogeneous strain in one or both grains in order to maintain compatibility [25]. 1.2.2 Geometrically Necessary Dislocations Another common approach to quantitatively measure heterogeneous deformation is to measure the local distribution of the density of geometrically necessary dislocations (GNDs). Because of the heterogeneous deformation between grains during global deformation processes in poly-crystalline specimens, as shown in Figure 1.2, GNDs are necessary to avoid the generation of overlaps and voids at grain boundaries in order to avoid micro-damage in the early stage of deformation. Ashby [26] and Busso et al. [27]   6       Figure 1.2 If each grain of a poly-crystal deforms in a uniform manner, overlap and voids appear (b). This can be corrected by introducing geometrically necessary dislocations, shown in (c) and (d). [26]   7   showed that the density of GNDs resulting from the existence of plastic strain gradients in a given grain is completely geometrically determined. Nye [28] showed that there exits a precise relationship between the dislocation tensor (α ), in other words, the field of GNDs at a given point in the material and the dislocation density in its neighborhood, expressed by S α = ∑ ρ (s)b(s) ⊗ z (s) (1.3) s=1 (s) where ρ (s) is the dislocation density of dislocation type s, b type s, and z (s) is the Burgers vector of the is the dislocation line direction of type S. S represents the set of all possible dislocations that could exist in the given crystal. Because there are too many combinations of Burgers vectors and line directions that can give an arbitrary α tensor, infinite sets of combinations can form the same lattice rotation. Usually, to narrow down the number of solutions, certain assumptions can be added to the general Equation 1.3 according to the nature of the materials of interest. In one of the Nye’s papers [28], only nine dislocation types with Burgers vector (s) b // <100> and z (s) // <100> were considered, which includes three pure screw and six pure edge dislocations. In this case, Equation 1.3 can be written as a second-order linear system form: 9 α ij = ∑ ρ (s)bi(s) z (s) j (1.4) s=1 (s) from which the unique or finite set(s) of solution of the dislocation density ρ can be solved. Sun et al. [29] only considered {111}<110> slip systems for face-centered cubic   8   (fcc) metals, which contain 12 pure screw dislocations and 24 pure edge dislocations. They solved for the dislocation density using Equation 1.3 by applying fcc dislocation tensor deconstruction method [30]. In other works [31, 32], only edge dislocation types were considered in order to simplify the calculation processes. According to Ashby, the field of GNDs is totally lattice curvature determined, therefore, the dislocation tensor (α ) can be experimentally determined by measuring the local lattice rotations using transmission electron microscopy (TEM), electronbackscattered diffraction (EBSD), or X-ray diffraction. By comparing the variation of GND density between and/or within different grains, the heterogeneity of plastic deformation in these grains can be observed. The relationships between micro damage initiation and the GND density, however, are still unknown. First of all, most of the GND density calculations are based on certain simplification assumptions, which may lead to the difference between calculated results and the real microstructure environment. Also, a great change in GND density across a grain boundary or within a grain does not necessary mean that damage will initiate at this spot. A grain may generate large amounts of GNDs at a local compression stress state, which generally does not tend to crack during deformation. 1.3 Experimental Characterizations and Simulations on Heterogeneous Deformation 1.3.1 Experimental Characterizations Based on the theories discussed in Section 1.2, in order to fully understand the heterogeneous deformation in different microstructure environments, including single,   9   bi-, and poly- crystals, a large number of experimental studies have been conducted over the decades. 1.3.1.1 Single crystals and bi-crystals Single-crystal materials, such as silicon and niobium, are used in different kinds of industries. A number of studies [33-36] on single-crystals have been conducted to study heterogeneous deformation processes within a single grain due to different global and/or local stress constraint. Also, since single crystals fully eliminate grain boundary effects, many previous studies [37-39] have used single crystals as model materials to study the relation between the activation of deformation systems and grain orientations relative to uniaxial stress states, such as tension and compression. Quantitative experimental measurements of heterogeneous deformation can be traced back to Livingston and Chalmers’ work on cubic bi-crystals in 1950s [20]. Bicrystal specimens serve as another simplified prototype to model heterogeneous deformation, because they have relatively simple stress states near the boundary and the orientation of the grain boundary with respect to the applied stress can be easily controlled. In bi-crystals, there is usually only one slip system in the interior part of the grains under global uniaxial stress states. It has been shown, however, by Livingston et al. that macroscopic continuity at the boundary will in general require the activation of at least four slip systems near the grain boundary in a bi-crystal, even under uniaxial stress states, because of the change in the grain boundary shape due to dislocation slip. This result indicates that the local stress state in the near-grain-boundary region can be much more complicated than in the interior parts of the grains, even in bicrystals. In other words, heterogeneous deformation exists in these bi-crystal tensile   10   specimens. To further quantity the dislocation activity in single crystals and bi-crystals during deformation processes, the GND density is often calculated based on the theory discussed in previous section using lattice rotation data measured by EBSD and/or Xray diffraction. EBSDs are Kikuchi patterns that form as the backscattered electrons from a stationary electron beam diffract when they exit the crystal. EBSD provides a convenient way to observe the surface heterogeneous deformation (especially local crystal lattice rotations [40-42] and texture changes [43, 44]) on the surface of crystalline materials. Advanced EBSD equipment has the advantage of being able to collect information from small areas under the electron beam, with a resolution in the range of hundreds of nanometers. However, the quantification becomes difficult at large strain levels. Sun et al. [29] have attempted to quantify the density of GNDs within an aluminum bi-crystal and at the grain boundary, by extracting lattice curvatures using EBSD and employing the geometrical relationship between strain gradients and GND density. Their results show that GND density and distribution are highly dependent on applied plastic strain level and grain boundary structure, as illustrated in Figure 1.3. The recently developed three-dimensional X-ray diffraction (3D-XRD) technique is able to characterize local orientations and stress states in bulk materials in a nondestructive manner. The basic setup of the 3D-XRD is illustrated in Figure 1.4. By analyzing the streak direction of a specific Laue reflection from a small volume (about 1 µm × 1 µm ×1 µm) of the material, the Burgers vectors associated with local GNDs might be identified [31]. 3D-XRD studies on Cu single-crystal micro-pillars [42], and Ni   11       Figure 1.3 (a) Orientations of the two crystals in relation to the compression direction Z. GND distribution in a 10% (b) and a 30% (c) strained aluminum bi-crystal [29].   12   Figure 1.4 A schematic drawing of the setup of 3D-XRD [45]. (For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this dissertation.)       13   bi-crystals [31] have quantitatively characterized the strain heterogeneity of local lattice rotation within/between grains. The in-situ 3D-XRD quantitatively measured the total lattice rotation of Cu single crystal micro-pillar oriented for double slip activation during compression testing [42], which was demonstrated by an EBSD scan from the same crystal, as shown in Figure 1.5. This work showed that lattice rotation occurred heterogeneously along the vertical axis of pillars during the compression test. Although single- and bi- crystals provide some basic understanding of grain lattice rotation and dislocation transfer across grain boundaries, many observed deformation behaviors in poly-crystals could not be simplified satisfactorily to the bicrystal deformation processes. Bi-crystals generally activate only one slip system with the highest value of Schmid factor in the interior part of grains, where grains behave like single crystals; in contrast, poly-crystals may require more activated slip systems during plastic deformation processes to keep the grain boundaries cohesive according to the von Mises criteria [35]. Also, poly-crystals can activate even more slip systems near grain boundaries than in the interior of the grains [20], because of the complex stress states and constraints near the boundaries. 1.3.1.2 Poly-crystals Based on the studies on plastic deformation of single- and bi- crystals, a number of extensive experimental studies of plastic heterogeneities have been carried out on cubic or near-cubic system poly-crystalline metals in recent years. Scanning electron microscopy (SEM) has been used to observe and measure the plastic deformation to assess how the local lattice curvature and dilation associated   14   Figure 1.5 Two vertical XRD scans (line 1 and 2) compared with the corresponding EBSD scan data. All curves display the total misorientation relative to the base orientation of the pillar plotted as a function of distance from the pillar base [42].   15   with GNDs affect the lattice interaction with incident electron beams. Selected-area channel patterns (SACPs), formed from the variation in intensity of backscattered electrons resulting from rocking the electron beam about a point on the surface of a single crystal or grain, are well known to deteriorate with increases in plastic strain [4648]. Many studies have quantitatively elucidated the dislocation density and lattice rotation behavior near crack tips in a poly-crystalline superalloy [49, 50], and an aluminum alloy [51] during plastic deformation using SACPs. One of the further developments of the application of SACPs is electronchanneling contrast imaging (ECCI), which allows the direct observation of dislocations and other micro-defects near sample surfaces without destruction of bulk samples [52]. This technique can be conducted by aligning the optic axis of the microscope to the edge of a strong channeling band using SACPs to set up “two-beam” conditions similar to that in TEM. This non-destructive technique is ideal for observing the heterogeneous behavior during the early stages of plastic deformation in crystalline materials [53, 54]. As shown in Figure 1.6, Simkin et al. [55] revealed dislocation pile-ups in a given grain near grain boundary at which twins in the neighboring grain intersected with this grain boundary. This observation is solid experimental evidence of the dislocation transfer across grain boundary in poly-crystal environment. Also, ECCI allows quantitative measurements of dislocation density [56] from a much larger area (compared with TEM) in different grains in bulk samples, which can reveal the heterogeneous deformation in these grains. These two techniques, however, are usually limited to the study of rather course grained polycrystalline materials or single crystals, because a clearly defined SACP requires the capability to rock the beam only in an area of the tens to hundreds of   16   Figure 1.6 (a) SACP composite from a deformed grain (~1% strain) shown in (b). The ‘‘×’’ indicates the electro optic axis at zero sample tilt. (b) Channeling contrast image of the deformation defects in a grain of TiAl with the calculated {111} plane traces. Arrowheads indicate bands of dislocations piled up in the neighboring grain at points of intersection of deformation twins with the grain boundary. The bright contrast regions, pointed by the black arrows in (b), near the grain boundary represents high density of dislocations [55].   17   microns in diameter, due to spherical aberration and focusing errors. Other experimental studies, such as Delaire et al. [19], have shown that slip systems with high Schmid factors (calculated based on the grain orientation collected by EBSD and the global stress state) tend to extend across grains, while slip systems with moderate Schmid factors had slip traces that extended part way into the grain interior in polycrystalline columnar-grained pure copper specimen under uniaxial tensile state. Another study has focused on the micro-scale distribution of GNDs during plastic deformation. According to Barabash, et al. [57], in the near-surface zone of severely deformed polycrystalline Ti after friction stir processing, the GND density shows a heterogeneous distribution between the thermal mechanical affected zone and the heat affected zone, between different grains in a given zone, and even within a particular grain. The study of hexagonal close-packed (hcp) metals, however, is limited, which will be discussed in Section 1.5. 1.3.2 Simulations of Heterogeneous Deformation The crystal plasticity finite element (CPFE) method is often used to simulate the plastic deformation processes that occur in three-dimensional microstructures, where the heterogeneous deformation between and within grains co-exists during deformation [19]. In CPFE simulation, the deformation gradient F is defined as the combination of e p elastic (F ) and plastic (F ) parts, which is the same as conventional FEM: F = F eF p (1.5) and by applying the flow rule,  F p = Lp F p   (1.6) 18   p The plastic velocity gradient (L ) is composed of the shear contributions of all slip systems [58] in CPFE modeling, ( L p = ∑ γ α mα ⊗ nα S ) (1.7) where Sα = mα ⊗ nα is the Schmid matrix for a given deformation system. This is one of the key equations that combine the crystallographic nature of dislocation slip with conventional finite element method. The slip rate of a given deformation system ( γ α ) is also closely related to the crystallographic nature of the materials, such as critical shear stresses (CRSS) and hardening parameters. When the resolved shear stress of a given system is above its CRSS, the given system will contribute to the total slip rate. On the other hand, if the local stress becomes lower than the CRSS after work hardening, this system will be shut down in the simulation. There are different ways to express the slip rate: n τα γ = γ 0 α sgn τ α [59] s α ( ) (1.8) or γ α = τ α − xα − r α K n ( ) sgn τ α − xα [60] (1.9) where τ α is the resolved shear stress on the given deformation system. sα , xα , rα are related to CRSS and hardening parameters, the values are time- and/or straindependent to reflect the strain hardening process.   19   Since the CPFE method is usually used to simulate microstructure scale deformation processes, it is also important to experimentally identify the grain orientations and shapes within the microstructral patches of interest before simulation. EBSD and Orientation imaging microscopy (OIM) have been widely used to identify crystal orientations, helping to generate meshes for CPFE simulations. Two kinds of three-dimensional meshes can be found among these works. One is called a quasi-3D mesh that represents the three-dimensional microstructure by extending vertically the surface grain information perpendicular to the interior of the grain (forming a set of columnar grains) [61, 62]. The other type of mesh is a real 3-D mesh that includes real crystal information, shape, and orientation from beneath sample surface [63, 64]. To develop such meshes, electro-polishing has often been used to remove surface layers. However, the shortcoming is that it is hard to control the thickness of the removed layer, which may result in errors in the construction of 3-D meshes. More recent studies [65, 66] have used 3D-EBSD (a combination technique of focused ion beam (FIB) and EBSD)), during which the removed layer can be more easily controlled, leading to more accurate 3D-mesh simulation studies. To date, most CPFE microstructure modeling efforts have focused on cubic and near-cubic materials such as aluminum [59, 67, 68], copper [19, 61], stainless steel [69], and TiAl [70]. Some of the microstructural simulations results have been compared to experimental data to assess the accuracy of these models. In a study of high-purity columnar-grain aluminum under plane-strain compression [59], the quasi-3D CPFE model was able to successfully predict both the average deformation texture and the local grain rotations, but it failed to predict the morphology of the “deformation bands”.   20   A dislocation-density-based finite element method (FEM) simulation results of the heterogeneous strain field of a single-layer-grain copper sample plastically deformed by uniaxial tension [19] matches well with the experiment results measured by grids deposited on the polished prior to deformation sample surface. In a study of polycrystalline copper by Musienko et al. [61] using both quasi-3D and real 3D meshes, the simulation of local lattice rotation and local strain fields of a 100-grain patch during uniaxial tension based on the two kinds of meshes, as shown in Figure 1.7, has been both compared to the experimental data measured by EBSD. The results for the quasi-3D mesh retain some traces of the experimental behavior, but the results are not as good as those obtained with the real 3-D mesh. Nevertheless, unlike CPFE simulations of cubic metals, fewer simulation studies of heterogeneous deformation of hexagonal metals and alloys have been reported [71, 72], and have not been adequately compared to experimental data. Details on the simulation works of hcp materials will be discussed in Section 1.5. Although various approaches of microstructural-scale computational simulations of heterogeneous deformation have been conducted in recent years, it is still difficult to simulate the behavior of grain boundaries and dislocations near grain boundaries, resulting directly in difficulty in predicting damage nucleation at grain boundaries. In most of the simulations discussed in the previous paragraphs, grain boundaries are simply treated as if they are transparent, which means that dislocations can effectively transfer across boundaries. It is very difficult to simulate how deformation transfer occurs by dislocation absorption and emission at grain boundaries due to the limited understanding of the atomic details of grain boundaries and dislocation sources.   21   Figure 1.7 The various mesh types used in Ref. [61]: (a) real 3D mesh (31,780 nodes), (b) real 3D mesh (130,818 nodes), (c) quasi-3D mesh (14, 076 nodes).     22   Grain boundaries can be modeled as a location where interactions of a series of Burgers vectors take place [73, 74]. In some simulation approaches, for example, low angle boundaries were modeled as a series of edge dislocations existing in the lattice [73]; while at high angle boundaries, the dislocation structures become much more complicated, making interactions between boundaries and near boundary dislocations unpredictable. In other studies [75-77], special FEM elements (with different mechanical and crystallographic properties from elements in the interior of grains) are added at grain boundary regions to represent the different behavior of grain boundaries (see Section 1.4). 1.4 Current Understanding of the Relationship between Heterogeneous Deformation and Damage Nucleation Many of previous studies have been focused on the parameters of crack growth and final critical crack length, with the assumption that initial microcracks exist in the materials, but these parameters are useless in predicting locations of damage nucleation. Since the 1980s [75], more emphasis has been put on the initiation of microcracks. There are several mainstream approaches toward predicting damage nucleation mechanisms at grain boundaries during plastic deformation at ambient temperature based on the current understanding of heterogeneous deformation and slip transfer mechanism. The first paper on predicting and simulating the damage nucleation and the development of the crack critical length was presented by Needleman [75] based on a parameter called the cohesive interface energy. In this paper, void nucleation by atom   23   de-bonding was correlated with normal and shear traction history in the grain boundary. It was assumed that there exists a cohesive zone in front of the physical crack tip that consists of upper and lower surfaces, called cohesive surfaces, held together by the cohesive traction. The cohesive traction is correlated to the separation distance between the cohesive surfaces by the cohesive zone model. The two cohesive surfaces separate gradually under an external applied stress. The separation of these surfaces at the tail of the cohesive zone will reach a critical value under a specific stress, and the crack growth will stop. A fundamental parameter of the cohesive zone model is the work of separation per unit area of cohesive surface, called the cohesive energy density [76, 77]. This two-dimensional model was first developed for simulating the micro-crack growth process using finite element method. Later, the idea of energy equilibrium was adapted in damage nucleation models. Hao et al. [78, 79] and Clayton and McDowell [80, 81] have further developed the model to predict successfully to some extent local stress and local strain evolution history and the boundary fracture. In their papers, each grain was modeled with a large number of finite elements, permitting gradients of elastic and plastic deformation to develop, both among different grains and within the same grains. Cohesive zone elements were inserted at all grain boundaries, allowing both normal and tangential inter-granular fracture initiation, and all the grain boundaries were assumed to have the same strength. In the model, the heterogeneous deformation was simply correlated with the residual lattice energy density between two grains. The locations of damage nucleation in a material are based on how much residual energy could be released by voids at grain boundaries under a certain applied stress level. The cohesive energy density at   24   grain boundaries with different orientations varies with respect to applied stress. Boundaries with higher cohesive energy, which are weakly bonded, may crack first. This model, however, is not perfect, considering the lack of dependence on different grain boundary strength arising from grain boundary structures and local grain misorientation. Thereby, the predicted locations of void nucleation are less persuasive. Recently, Simkin et al. [82] showed that in near-gamma polycrystalline TiAl (near cubic structure), microcracks were observed where highly localized deformation twins interacted with grain boundaries. It is the twin shear that causes a concentrated stress on one side of the boundary (shown in Figure 1.8 (a)), which cannot be explained by the theories introduced earlier. To fix the deficiency, Simkin et al. introduce a fracture initiation parameter that contains more geometrical factors to interpret this phenomenon [83]. The fracture initiation parameter (fip) is defined as: F = mtw btw ⋅ t ∑ ord btw ⋅ bord (1.10) where, as shown in Figure 1.8(b), mtw is the Schmid factor for a specific twinning system in the grain, assuming a local stress state equal to the global stress state, btw and bord are the unit Burgers vectors for the single twin system and ordinary dislocation systems in each grain respectively, t the direction of the traction force acting on the boundary, btw ⋅ t is an expression of the opening force acting on the boundary, and ∑ ord btw ⋅ bord describes how well the local shear direction at the boundary can be accommodated by dislocations in adjacent grain. The results for TiAl show that larger values of F correspond to the experimentally observed locations for crack initiation, while maxima of the components of fip show weak relationship among themselves and   25   (a) (b) Figure 1.8. (a) Backscattered electron image of γ-γ TiAl grain boundary damage initiation due to twin interaction with the grain boundary [82]; (b) 2D schematic of crack nucleation at a concentrated band of slip or twinning [82, 83].     26   fracture initiation. This indicates that a combination of several factors, rather than a single factor, influences the damage behavior at grain boundaries. By combining several geometric factors, a fip can identify the “weak” grain boundary conditions in the context of a specific state of applied stress [70]. The fip may also identify when plastic shear at boundaries can cause partial slip transfer, which will leave the residual dislocations in the boundary, leading to damage nucleation at these boundaries. Nevertheless, the fip only considers slip/twinning directions with respect to the applied stress, instead of both the orientations of the slip/twinning planes and directions, implying that slip directions play more significant roles in crack nucleation. The fip works quite well in TiAl system because of the highly localized twinning behavior during deformation [82], but its credibility in other alloy systems remains unproved. 1.5 Studies on Commercial Purity Titanium and Other Hexagonal Materials In commercial purity titanium, four dislocation slip systems have been reported [85, 90], as shown in Figure 1.9. {1010} < 1210 > prismatic slip is the primary slip system during room temperature deformation, but the other three slip systems, {0001} < 1210 > basal slip, {1011} < 1210 > pyramidal slip, and {1011} < 2113 > pyramidal slip, can be activated in some orientations as well as in order to satisfy the von Mises criteria to permit an arbitrary plastic strain in polycrystalline titanium. The CRSS of slip systems in α-titanium determined in different studies are summarized in Table 1.1. Among all slip systems, prismatic slip has the lowest CRSS. There are also four deformation twinning systems in α-titanium that can contribute to deformation (detailed information is given in Table 1.2 [91]), as shown in Figure 1.10. When the   27     Figure 1.9 Slip systems for α-titanium: (a) prismatic slip, (b) basal slip, (c) pyramidal slip, and (d) pyramidal slip. The arrows only represent the slip directions, not the magnitudes of the dislocation slip. Figure 1.10 Twinning systems for α-titanium: (a) T1 twinning, (b) T2 twinning, (c) C1 twinning, and (d) C2 twinning. The arrows only represent the twinning directions, not the magnitudes of the twinning.   28   Table 1.1 CRSS of different slip modes reported for α-titanium. Prismatic slip Basal slip Pyramidal Pyramidal Refence [86] (relative values) 1 3~6 6~15 3~6 Refence [87] 30 MPa 150 MPa 120 MPa N/A Refence [92] 37 MPa 49 MPa 197 MPa N/A     29   Table 1.2 Deformation twinning modes in α-titanium. Mode Twining plane and direction Theoretical shear T1 {10 12} < 1011> {1121} < 1 126 > {1122} < 1123 > {10 11} < 10 12 > 0.171 T2 C1 C2   30   0.629 0.221 0.101 maximum principal stress direction is close to the crystal c-axis orientation, either slip, tensile (or extension) twinning modes (T1 and T2), or compressive (or contraction) twinning modes (C1 and C2) have high Schmid factors, and their operation often contributes to the crystal shape change. Among these four twinning systems, {1012} < 1011> (T1 extension) twinning is the most commonly observed at room temperature due to its relatively low magnitude of shear (0.17). In this project, “soft” crystal orientations and/or grains in α-titanium are defined as those grains with high Schmid factors for prismatic slip – those grains having their c-axis nearly perpendicular to the applied uniaxial stress. Hard crystal orientations and/or grains, however, are those grains with low Schmid factors for prismatic slip, with their c-axis parallel to applied uniaxial stress. Compared to cubic metals, fewer works have been reported on heterogeneous deformation of polycrystalline hexagonal materials (Mg, Ti, Zr, etc.) than those for cubic materials. Simulations of a polycrystalline magnesium alloy [71] indicated that a high stress concentration always occurred near grain boundaries between an untwined grain and a twinned grain, regardless of the amount of global strain. The results demonstrate that it is important to include deformation twinning and twin-induced reorientation into the CPFE model analysis of heterogeneous deformation in hexagonal materials. Dunne et al. [72] demonstrated that elastic anisotropy in titanium alloys has a significant effect on the local stress distribution and the magnitude of accumulated slip. They also showed that the inclusion of length-scale effects tended to lead to a more smooth plastic strain distribution locally near grain boundary by generating GNDs. In a study of columnar polycrystal zirconium under tension by Héripréa et al. [93], CPFE modeling   31   results were compared to the experimental strain field obtained by digital image correlation at the scale of the grains. They suggested that this comparison could be used for identification of parameters of crystallographic constitutive laws, such as CRSS and hardening parameters for CPFE simulations. To date, there is still no breakthrough on predicting damage nucleation in hexagonal materials. 1.5 Motivations for This Study Hexagonal polycrystals offer a variety of opportunities to study heterogeneous deformation and damage nucleation research. First of all, hexagonal metals have not been studied as much as cubic materials, mainly because the process of modeling hexagonal metals is more complicated. The CRSS and hardening parameters of the various deformation systems are difficult to determine experimentally for these materials. For cubic metals, uniaixal tension and/or compression tests on single crystals are often used to measure the values of CRSS for deformation systems. Work by Williams et al. [84] has used hexagonal close-packed (hcp) TiAl alloy single crystals to measure the CRSS of the material. However, for hexagonal metals, the values of CRSS for the different systems are more often measured by carrying out statistical analysis of slip traces in polycrystals [85] or curve fitting in association with CPFE simulations [86]. Consequently, the published values of the CRSS of hexagonal metals vary significantly [84-87]. Secondly, the lower crystal symmetry of hexagonal materials is expected to result in more deformation heterogeneity than cubic metals. In general, the c axis of hexagonal crystals is inextensible, except by the pyramidal slip or twinning.   In contract to cubic metals, in hexagonal materials, there are fewer slip 32   systems in the most easily activated categories of slip systems (one slip plane and three slip directions for basal slip; three slip planes and one slip direction on each plane for prismatic slip) than those in cubic materials. This makes it easier to identify specific activated slip or twin systems using trace analysis [55] under defined stress states. Thirdly, hexagonal crystal materials also show strong texture after rolling processing. Those hexagonal metals having higher values of the c/a ratio such as zinc or cadmium reveal textures with basal poles tilted about ±15° to ±25° from the normal direction towards the rolling direction [88]. Metals with c/a ratio lower than the ideal value, like titanium, show a splitting of the basal poles in the transverse direction [88, 89]. With the help of the EBSD technique, it is easy to recognize grain boundaries with high c-axis misorientation, where strong deformation heterogeneity might exist. Based on the discussions above, the long-term goals of this study is to understand the relationships between activated slip systems and grain boundaries, and to model their inflences on damage nucleation at grain boundaries in commercial purity titanium. Indeed, the ability to predict damage nucleation is also one of the major goals of computational plasticity. Since heterogeneous deformation is commonly viewed as a precursor to damage nucleation, this long-term goal is actually based on accurately measuring heterogeneous deformation in polycrystalline specimens, and developing more reliable multi-scale models that can accurately model conditions that favor damage nucleation in interfaces.   33   1.6 Overview In this study, commercial purity titanium (hcp metal) was used as a model material to study different scales of heterogeneous deformation in fatigue crack samples and four-point bending samples. A series of comprehensive experimental and modeling analysis have been conducted to characterize the heterogeneous deformation within different scales of microstructure patches. Chapter Two provides all the material parameters and experimental details and constitutive model using CPFE method including in this project. In Chapter Three, the heterogeneous deformation within micro-scale patches is presented. Based on a series of observations of the activation of deformation systems, the formation of grain boundary ledges, and grain rotations within poly-crystalline αtitanium, the heterogeneous deformation within a carefully chosen microstructure patch (including about 20 grains) was studied in detail. The local dislocation activity of this patch was fully quantified using a new combined technique using atomic force microscopy (AFM), Backscattered electron (BSE) imaging, and EBSD. EBSD data was collected before and after four-point bending to facilitate trace analysis of deformation activity in grains. Distribution maps of local shear caused by dislocation slip derived from z-displacement data measured by AFM were then directly compared to results of CPFE simulations that incorporate a phenomenological model of the deformation processes to evaluate the ability of the CPFE model to match the experimental observations. The existing CPFE model successfully predicted most types of active dislocation slip systems within the grains at correct magnitudes, but the spatial   34   distribution of strains within grains differed between the measurements and the simulation. As the difficulty of directly measuring the CRSS of different deformation systems in titanium can be a major cause of the deficiency of the match between CPFE modeling results and experimental measurements, in Chapter Four, a nano-indentation technique is introduced to examine the CRSS and flow laws of commercial purity titanium. Arrays of conical nano-indentations were made on patches of grains, and AFM and EBSD scans of on the patch indicates that the surface pile-up topography was strongly crystallographically dependent. Corresponding CPFE simulations using a similar constitutive model to that used in Chapter Three predicted the pile-up patterns in good agreement with the experimental measurements. The CRSS values for prismatic, basal, and pyramidal slip of the CPFE model were identified by optimizing the simulation results (load–displacement and residual pile-up pattern) of the indentation process in different grain orientations using a non-linear optimization method. A rerun of the CPFE simulation of the same microstructure patch studies in Chapter Three with the optimized parameters, however, showed only minor differences, which indicates more quantitative studies on strain hardening behavior and slip transfer across grain boundaries are strongly needed. In Chapter Five, a novel approach using BSE imaging and discrete Fourier transformation (DFT) processing was established to measure the meso-scale plastic strain distribution within plastic zone around a fatigue crack tip. An in-situ uniaxial tensile calibration specimen indicates that the central peak widths of DFT of BSE images have a close relationship with the residual strain with the grain patches. The   35   measurement results were demonstrated to have good agreement with the plastic zone size and shape measured using thermoelastic stress analysis (TSA). This approach should prove to be an effective way to measure local heterogeneous deformation in other kinds of deformed samples. In the future, ECCI will be used to observe the dislocation activity near grain boundaries to help further understanding the mechanism of heterogeneous deformation. Future experiments will also be focused on establishing a new experimental characterization method to facilitate the building of a real-3D mesh, and to compare the CPFE simulation results of the real-3D mesh with the existing results from the quasi-3D mesh. Future simulation work will be focused on developing a more advanced CPFE model, considering the information of slip transfer across grain boundaries, to simulate heterogeneous deformation within grain patches more reliably.   36   CHAPTER TWO MATERIAL, EXPERIMENTAL DETAILS, AND SIMULATION FRAMEWORK 2.1 Material Two different sources of commercial purity titanium were used for different aspects of this program. These will be designated as Material A and Material B in this dissertation. Detailed information about the two sources of commercial purity titanium is provided below. 2.1.1 Material A Material A was provided by Max-Planck-Institut für Eisenforschung, Düsseldorf, Germany (MPIE) as a plate approximately 15 cm × 20 cm × 5 cm in dimension. Unfortunately, the original supplier and processing history of the plate are not known. It had a moderately strong texture (about eight times random, as shown in Figure 2.1) and an average grain size of about 80 microns. X-ray diffraction studies done in MPIE prior to deformation indicated that the grains generally contained a low density of dislocations, though the internal stress fields could be significant in minority grain orientations, arising from the anisotropic coefficient of thermal expansion. Specimens with dimensions of 25 mm × 3 mm × 2.5 mm (shown in Figure 2.2) for four-point bending tests and nano-indentation tests were cut from the as-received annealed plate with their longitudinal direction oriented at 0° and 45° from the plate rolling direction, denoted as specimen series S0 and S45, respectively.   37     Figure 2.1 {0001} and {10 10} pole figures of (a) specimen S0, and (b) specimen S45.   38   Figure 2.2 Specimen orientations in Material A plate (left), and the size of specimens for four-point bending tests and nano-indentation tests (right).     39   2.1.2 Material B The other commercially purity titanium plate, denoted as Material B, had a weak texture (about four times random, as shown in Figure 2.3) with an average grain size of about 20 microns. Specimens for in-situ tensile tests and fatigue tests were cut from the as-received annealed plate. The sizes and shapes of the two types of specimens are shown in Figure 2.4 and Figure 2.5, respectively. 2.2 Sample Preparation Prior to deformation, the specimens were mechanically polished. The specimens were first ground using SiC grinding papers through 240, 400, 800, 1200, 2500, and 4000 grits (European units). Each grinding step took about 10 minutes, and specimens were rinsed using water and methanol between each step. The pre-ground specimens were then polished on a specific polishing cloth (Struers, Code: MECHE, Category No. 40400094) with 0.05 µm colloidal silica plus 30% hydrogen peroxide. The samples were then dipped into glycerin 68% + nitric acid 16% + hydrofluoric acid 16% for 4 to 5 seconds at room temperature, and rinsed with distilled water after etching. Polishing with colloidal silica and etching was repeated three to four times to remove the thin deformation layer resulting from sample cutting and pre-grinding of the specimen surfaces. The polished surfaces were considered to be acceptable when high-quality EBSD patterns (confidence index greater than 0.8 for most of the patterns) could be obtained from the specimen surface prior to deformation. To carry out differential image correlation studies, fine tungsten cross patterns were deposited using FIB on some areas of some Material A specimens prior to deformation, as shown in Figure 2.6.   40   Figure 2.3 {0001} and {10 10} pole figures of Material B.     41   Figure 2.4 Size and shape of specimens for in-situ tensile tests cut from Material B.     42   Figure 2.5 Size and shape of specimens for fatigue tests cut from Material B plate.     43   Figure 2.6 Backscattered electron image of a microstructure patch with deposited fine tungsten cross patterns.   44   2.3 Experiment Details 2.3.1 Scanning Electron Microscopy 2.3.1.1 Electron Imaging Two SEMs were used in this project to conduct imaging before and after deformation. One was a CamScan 44FE field emission scanning electron microscope (CamScan, Cambridge, UK) with a polepiece-mounted, four-quadrant, silicon diode type BSE detector, located in Department of Chemical Engineering and Materials Science, Michigan State University (MSU). Backscattered electron (BSE) and secondary electron (SE) imaging were conducted at an accelerating voltage of 25 kV, with a working distance of 15 mm. Images were captured using an external personal computer with 640 × 480 pixel digital resolution. The other SEM was a high-resolution JEOL JSM 6500 F SEM, located in Department for Microstructure Physics and Metal Forming, MPIE. SE imaging was conducted at an accelerating voltage of 15 kV, and captured with 1024 × 768 pixel digital resolution. 2.3.1.2 EBSD/OIM measurement Grain orientations were determined using identical orientation imaging microscopy (OIMTM) systems (TSL, Draper, UT, USA) installed on both the CamScan 44FE and the JSM 6500. A schematic diagram of a typical EBSD system is shown in Figure 2.7, and detailed information about the system can be found in a review paper of Humphreys [94]. EBSD was carried out at a working distance of 33 mm on specimens that were tilted 70 degrees from the horizontal surface. All of the EBSD data in this   45     Figure 2.7 A typical setup of an EBSD/OIM system in SEM [94].   46   project was collected using beam-scanning mode, during which the electron beam spot scanning processes are controlled by the EBSD acquisition software. The EBSD data was collected using OIM Collection v.OIMDC 4.51 (TSL, Draper, UT, USA), and the EBSD data processing software used in this project was OIM Analysis v.5.31 (TSL, Draper, UT, USA). 2.3.1.3 In-situ Tensile Testing In-situ uniaxial tensile tests were conducted on the tensile specimen, shown schematically in Figure 2.4(a). The commercial 100 lb in-situ tensile stage (E. Fullam, Latham, NY, USA), shown in Figure 2.8, was installed within the CamScan 44FE scanning electron microscope. The in-situ tensile testing provided a calibration standard for Material B, which will be discussed in detail later in Chapter Five. Another separate continuous tensile test (non-stop test from 0% strain to specimen fracture) was conducted to measure the stress-strain curve for material B. The strain increments at which the in-situ test was stopped and images are taken were marked out by black arrows on the stress-strain curve, as shown in Figure 5.6. BSE images were taken from six areas in the middle of the specimen at each deformation step, under the identical imaging conditions (brightness, contrast, magnification, and working distance) as those for imaging the fatigue specimen of Material B. Because the brightness and contrast knobs in the CamScan are not number-based, the positions of these knobs were carefully noted when changing the specimens to guarantee the same brightness, contrast, and working distance.   47   Figure 2.8 Deformation stage for in-situ tensile test.   48   2.3.2 Four-point Bending Testing Four-point bending tests produce maximum tensile stress states at the directly observable surface of the deformed specimens, also eliminating the triaxial stress states present in uniaxial tensile tests. Also, according to an elastic-plastic continuum FEM simulation of this geometry [95], the plastic strain at the surface is approximately uniform in the region between the two inner bending pins spaced 5 mm. Therefore, the assumption of uniaxial tensile stress can be made for the surface region within the two inner bending pins during bending tests. The bending stage shown in Figure 2.9 was designed so that a strained but still loaded specimen could be investigated in the SEMs as well as the 3-D XRD facility at Argonne National Laboratory (ANL). The polished samples examined in this project were deformed by four-point bending to achieve a surface tensile strain of about 8% and 15% for S0-1, 1.5%, 3%, and 5% for S45-1, and 1.5% and 4% for S0-2, respectively. Microstructure investigations were concentrated in a 2.5mm × 2.5mm surface area located near the center of the bend specimens, where the stress state was well approximated as uniform tension. 2.3.3 Atomic Force Microscopy AFM was performed in “tapping” mode using two Dimension 3100 AFMs (Digital Instruments, Plainview, NY, USA) in this work. One was located at MSU and the other at MPIE. The AFM probes used in this particular project were TESP sharp silicon probes (Bruker, Billerica, Massachusetts, USA). Figure 2.10 shows both a schematic drawing (a) and an SEM image (b) of an AFM probe tip. The AFM data was collected   49   Figure 2.9 Four-point bending stage with a bended specimen loaded.   50   Figure 2.10 Schematic drawing (a) and an SEM image (b) showing the size and shape of the AFM probe tips used in this project [96].   51     and processed using the software Nanoscope III v. 5.31 (Digital Instruments, Plainview, NY, USA). In “tapping” mode, the cantilever oscillates at about its resonance frequency with an amplitude typically from 100 nm to 200 nm. When the cantilever comes close to the sample surface, the amplitude of the oscillation will be decrease, mainly due to Van der Waals force and electrostatic forces. An electronic servo will adjust the height of the cantilever to maintain a constant oscillation amplitude while the cantilever scans across the sample. Consequently, a tapping-mode AFM scan is produced as a result of the height change of the cantilever. 2.3.4 Nano-indentation Nano-indentation was carried out with the help of Claudio Zambaldi at the MPIE using a TriboScope 900 (Hysitron, Minneapolis, MN, USA). The nano-indentation in this work was carried out under load-controlled conditions with a load of 6 mN. Instrumented indentation was performed with an indenter of sphero-conical tip geometry. Rectangular arrays of indentations were made on specimens S0-2 (see Figure 2.11) and S0-3 (see Figure 2.12). Each region included about 20 grains with both soft (c-axis parallel to the specimen surface) and hard grain orientations (c-axis perpendicular to the specimen surface), with a total of 223 indents for specimen S0-2 and 220 for specimen S0-3. The residual indentation topographies were analyzed using a Dimension 3100 (Digital Instruments, Plainview, NY, USA) atomic force microscope outlined in Section 2.3.3.   52   Figure 2.11 BSE image of the indented microstructure patch in specimen S0-2.   53   Figure 2.12 BSE image of the indented microstructure patch in specimen S0-3.   54   2.3.5 Fatigue Testing and Thermoelastic Stress Analysis Measurement Before fatigue testing, the polished surface of the fatigue specimens (see Figure 2.5) were sprayed with a thin coat of matte black paint using a fine airbrush, in order to provide uniform emissivity for the TSA measurement before fatigue test. The fatigue tests were performed on a 50kN MTS servohydraulic loading frame in air, with mean load = 600 N, and load amplitude = 150 N. The test frequency was 20 Hz, and the tests were stopped after 27,200 cycles. During the fatigue crack growth, TSA data were recorded at regular intervals. The thermoelasticity analysis technique is based on the principle that under adiabatic and reversible conditions, a cyclically loaded structure experiences in-phase temperature variations that are proportional to the sum of the principal stresses, which are measured using a sensitive infrared detector. Around the tip of a fatigue crack during fatigue tests, however, there is a heat generation process due to the plastic deformation/work, which is related to the dislocation motion and grain boundary sliding. Heat evolved as a result of the relative phase difference due to these non-adiabatic effects has been shown to be useful for dynamically imaging the plastic zone size [97, 98]. The TSA technique allows the plastic zone to be monitored in near real-time, in a non-contacting manner. The phase of each TSA image was shifted so that the region far away from the crack tip (far field) was zero in order to represent the phase difference relative to the far field. The non-adiabatic region is thus represented by a negative phase difference, and therefore the size and shape of the plastic zone can be quantified using a binary filter on each image [98]. The fatigue tests and TSA measurements were   55   conducted with the help of Rachel Tomlinson from Shefield University, while visiting MSU. Following fatigue testing, the matte black paint was removed by ultrasonic cleaning in methanol bath for about three minutes. Subsequent SEM imaging reveled the cleaning procedures allowed the matte black paint to be removed to reveal surfaces consistent with pre-painted polishing conditions. 2.4 Slip/twin Plane Trace Analysis When a grain orientation is measured by EBSD, described as three Euler angles (φ1, Φ, φ2), the rotation matrix (R) of this orientation is given by: ⎡ cos φ 2 ⎢ R = Rφ 2 RΦ Rφ1 = ⎢ sin φ 2 ⎢ ⎢ 0 ⎣ − sin φ2 cos φ2 0 0 ⎤⎡ 1 0 0 ⎥⎢ ⎥ ⎢ 0 cos Φ − sin Φ 0 ⎥⎢ 1 ⎥ ⎣ 0 sin Φ cos Φ ⎦ ⎤ ⎡ cos φ1 − sin φ1 0 ⎤ ⎥ ⎥⎢ ⎢ sin φ cos φ1 0 ⎥ ⎥⎢ 1 ⎥ ⎥⎢ ⎦⎣ 0 0 1 ⎥ ⎦ (2.1) For a given slip/ twin system (hkl)[uvw], the slip/twin direction (m) and plane normal (n) in the sample/lab coordinate system can be then identified as: T ⎡ u ⎤ m=R ⎢ v ⎥ ⎢ ⎥ ⎢ w ⎥ ⎣ ⎦ (2.2) ⎡ h ⎤ n= R ⎢ k ⎥ ⎢ ⎥ ⎢ l ⎥ ⎣ ⎦ (2.3) T All of the possible traces of deformation systems in a given grain orientation were calculated and drawn using a MATLAB (MathWorks, Natick, MA, U.S.A) code (see Appendix I), based on the principle discussed above. The deformation systems in grain   56   patches were then effectively identified by comparing their slip/twin plane traces in the BSE images with calculated traces, because of the limited number of slip directions on each deformation plane in α-titanium with hcp structure. In this project, all the Schmid factors were calculated based on the assumption of global uniaxial stress states: m = cos φ cos λ (2.4) where φ is the angle between the stress axis and the slip plane normal, and λ is the angle between the stress axis and the slip direction. In the trace analysis, for deformation systems having very similar plane traces, two factors were considered to determine the deformation systems. First, the system with the highest Schmid factor was chosen. This criterion has its limitations, because of the difference between the assumption of global uniaxial stress states and the actual complicated local stress state in a given grain. Therefore, if the two systems in a given grain have both similar plane trace and similar high Schimid factors (both above or around 0.4), the second factor, slip transfer criteria (see Section 1.2.1), was considered. If the deformation system(s) in its neighboring grain have been determined, the deformation system in the given grain having higher m’ with the determined system in neighboring was chosen. 2.5 Constitutive Law in Crystal Plasticity Finite Element Method CPFE method simulations were conducted to model the deformation processes of uniaxial tension and nano-indentation, and the results will be discussed in detail in Chapter Three and Chapter Four. The basic constitutive description framework of the material in these two simulations is the same, which are both based on a crystal   57   plasticity formulation using the multiplicative decomposition of the total deformation gradient and considering the anisotropic elastic constants of α-titanium, which have been addressed in detail in many prior theoretical works [58, 59]. A brief summary of the theoretical formulations is given as following: The deformation gradient, denoted as F, is decomposed into two parts in a finite e p deformation framework, the elastic gradient, F , and the plastic, F : F = F eF p (2.5) The elastic strain part can be then expressed as: Ee = ( 1 eT e F F −I 2 ) (2.6)  I is a second-order identity tensor. The evolution of plastic gradient, F p , is given by  F p = Lp F p (2.7) where the plastic velocity gradient L resulting from activity on all deformation systems p is described [59] as N L = p def .sys. ∑ α Pα γ τ 0 α sα n sgn (τ α ) (2.8) with Pα = mα ⊗ nα as the Schmid matrices with respect to the undeformed state, γ 0 = 10 −3 s −1 as reference shear rate, n the constant stress exponent, τ α the resolved shear stress, and sα the shear resistance (in other word, CRSS). The evolution of sα during deformation is written as   58   N  sα = def .sys. ∑ β a ⎛ ⎞ sα ⎟ β⎜ h 1− γ β 0 ⎜ α⎟ s ⎠ ⎝ s (2.9) β with h0 , a, and sα the three hardening parameters. s The CRSS and hardening parameters were based on prior published experimental results [85, 86, 92] and adjusted slightly to enhance numerical stability and to better reflect the real deformation process during simulations. The chosen values for these parameters for each simulation are listed in Chapter Three and Chapter Four.   59   CHAPTER THREE MEASURING AND CPFEM SIMULATING MICRO-SCALE HETEROGENEOUS DEFORMATION In this chapter, a novel technique that combines AFM, BSE imaging, and EBSD is developed to allow quantitative analysis of the active deformation systems in different grains in a selected microstructural patch of polycrystalline titanium of commercial purity. The specimens were deformed by four-point bending, and EBSD data was collected before and after bending tests to facilitate trace analysis of deformation activity in grains. After each step of deformation, the surface step height changes due to dislocation slip and twinning activity in a grain patch were analyzed using AFM. Maps of local shear at three different strain levels (1.5%, 3%, and 6%) were generated to help visualize the heterogeneous deformation in this particular microstructure patch. Furthermore, by comparing these maps to the corresponding CPFE results, the quality of the constitutive model employed and CPFE mesh effects can be critically assessed, with the ultimate goal of accurately modeling the influence of grain boundaries on damage nucleation.   60   3.1 Microstructure in a Four-point Bending Sample 3.1.1 Slip Lines and Twins Five specimens were loaded by four-point bending to various amounts of global strains. On the tensile surface of deformed specimens, which was polished prior to deformation, the surface topography change of the grains varies from one grain to another, due to the differences in the grain orientation and the number density of dislocations on activated deformation systems. Figure 3.1 shows a typical microstructure patch of a deformed specimen after 1.5% global strain, and three highmagnification images of three grains from the patch with different topography. Figure 3.1(a) shows a grain with dense and deep slip lines, while in Figure 3.1(b), the density of slip lines in the grain is much lower. Also, a thin twin is shown in Figure 3.1(c).   3.1.2 Grain Boundary Ledge Formation Since under a given stress state, the deformation systems that can be activated in a given grain orientation in α-titanium are limited, additional plastic deformation mechanisms may be necessary for polycrystalline compatability according to von Mises criteria [99]. In this case, grain boundary ledges were also found under various amounts of tensile strain. In sample S0-1 after 8% tensile deformation, SEM images gave the appearance of grain boundary cracks, but upon closer inspection, these were found to be ledges. As shown in Figure 3.2, when sample is rotated 180 degrees in the SEM, the circled grain boundaries show reverse contrast in two images, indicating ledge formation along these boundaries rather than cracks.   61     Figure 3.1 A typical microstructure patch from S45-1 at 1.5% strain. (a) shows a grain with dense slip lines, while (b) shows a grain with much less slip lines. Twins also shows in grains, as illustrated in (c).     62   Figure 3.2 Secondary electron images show ledges (circled) along grain boundaries in S0-1under 8% global strain; (b) was taken of the same microstructural patch with the whole sample rotated 180 degrees from (a). Circled boundaries showing reverse contrast in two images, indicating these are ledges and not cracks.     63   Even at small strain levels, as low as 1% to 2%, some grain boundaries in polycrystalline titanium begin to show the development of grain boundary ledges, which can be clearly observed using AFM. Figure 3.3 shows AFM scans of three characteristic grain boundary ledges with different surface topography. Figure 3.3(a) and (c) shows sheared grain boundary ledges. Local surface rotation is found in the right grain in Figure 3.3(a), in (c), however, no obvious local surface rotation is apparent in both grains. In Figure 3.3(b), the surface topography across the grain boundary is much flatter than the other two. These different features of grain boundaries may result from heterogeneous strain field near these grain boundaries. 3.1.3 A Selected Microstructure Patch for Further Studies As discussed in Section 1.5, dense slip lines usually shows up in those soft grains with their c-axis almost perpendicular to the uniaxial tensile stress. However, less deformation may occur in those hard grains with their grains with their c-axis close to the tensile stress. To better understand the heterogeneous deformation in α-titanium, detailed studies of grain patches that have a mixture of both soft and hard grains are essential. To develop this understanding, a microstructure patch was selected in sample S45-1 prior to deformation. Figure 3.4(a) shows the selected microstructure patch at about 1.5% global tensile strain. Grains 3, 6, and 9 show the most obvious slip traces, while less pronounced slip bands are found in grains 8 and 10. T1 twinning in grain 2 is confirmed by EBSD (the misorientation angle between T1 twin and matrix lattice is about 85 degrees [91]). Figures 3.4(b) and (c) show OIM maps of the dashed box in Figure   64   Figure 3.3 AFM “Section” scans shown ledges with different topography at grain boundaries in S0-2 under 1.5% global strain, indicating that heterogeneous deformation occurs in small deformation.     65   3.4(a) before and after deformation, which provides orientation and shape change information. The variation of deformation activity (Figure 3.4(a)) and crystal rotations (Figure 3.4(c)) results from heterogeneous deformation among and within these grains. Two further deformation steps were conducted on specimen S45-1 to obtain global strain level of about 3% and 6% respectively. At 3% global strain, as shown in Figure 3.5(a), the growth of the twins in grain 2 can be easily observed, while no obvious secondary slip and twinning were found at this stage. At 6% strain, in addition to the growth of twins, secondary deformation systems began to show up in several hard grains, as illustrated by the white lines in Figure 3.5(b). Also, sharp ledges formed at the boundaries of grains 8-9, 9-10, and 3-8 boundaries. The deformation systems for all slip bands/twins in each grain were effectively identified by the trace analysis discussed in Section 2.4. Three types of dislocation slip systems (prismatic, basal, and pyramidal slip) and one T1 twinning system (in grain 2) were identified in the grain patch at 1.5% global strain, as illustrated in Figure 3.6. Prismatic slip dominates most of the soft grains, such as grains 1, 3, 6, 9, and 14, while basal slip were found in grains with relatively hard orientations, such as grains 5, 7, 11, 13. Pyramidal slip was only identified in grain 10, and no obvious slip traces were observed in grain 4. Secondary deformation systems, in paticular pyramidal slip and T1 twin, were identified in hard grains at 6% strain level. Detailed deformation system information in each grain is shown in Table 3.1.   66     Figure 3.4 Microstructure patch of specimen S45-1 at ~1.5% tensile strain. The patch within the solid outlined box was quantitatively analyzed using AFM-based method, and the patch within the dashed outlined box was modeled by CPFEM. Inverse pole figures of the grain orientation changes within the solid outlined box before and after deformation are shown in (b) and (c), respectively. Directions of slip lines and twins in each grain are highlighted by white lines. These activated deformation systems were identified using trace analysis based on EBSD data.     67     Figure 3.5 The same microstructure patch in S45-1 specimen as shown in Figure 3.3(a) following further deformed to ~3% and ~6% tensile strains, shown in (a) and (b) respectively. The patch within the solid outlined box was quantitatively analyzed using the AFM-based method. The growth of five T1 twins in grain 2 with the increase of the global strain can be observed. Furthermore, secondary deformation systems, highlighted by white solid lines in (b), were observed in several hard grains, such as grain 2, at 6% strain level.   68   Figure 3.6 BSE image illustrating the area characterized using AFM and CPFEM at 1.5% global strain. Different activated deformation systems characterized using trace analysis are denoted by color lines.   69   Table 3.1 Deformation systems in grains within the solid-line box in Figure 3.3(a) and 3.4(a)-(b) at global strain level 1.5%, 3%, and 6%. Grain number Identified deformation systems using trace analysis 1.5% strain 3% strain 6% strain 1 (10 10)[12 10] (10 10)[12 10] (10 10)[12 10] 2 (10 12)[1011] (10 12)[1011] 3 (10 10)[12 10] (10 10)[12 10] (10 10)[12 10] 4 N/A N/A N/A 5 (0001)[12 10] (0001)[12 10] (0001)[12 10] 6 (10 10)[12 10] (10 10)[12 10] (10 10)[12 10] 7 (0001)[12 10] (0001)[12 10] (0001)[12 10] 8 (01 10)[2 1 10] (01 10)[2 1 10] 9 (10 10)[12 10] (10 10)[12 10] (10 10)[12 10] (0001)[12 10] (0001)[12 10] (0001)[12 10] (10 11)[1123] (10 11)[1123] (10 11)[1123] 11 (0001)[12 10] (0001)[12 10] (0001)[12 10] 12 (0001)[2 1 10] (0001)[2 1 10] (0001)[2 1 10] 13 (0001)[1 120] (0001)[1 120] (0001)[1 120] 14 (10 10)[12 10] (10 10)[12 10] (10 10)[12 10] 10     70   (10 12)[1011] (1101)[12 13] (01 10)[2 1 10] (0 111)[1213] 3.2 Quantitative Measurement of Heterogeneous Deformation within a Grain Patch AFM is an effective tool for characterizing fine-scale topography induced by different deformation processes. Since the 1990s [100], AFM has been used to study surface relief produced during fatigue damage evolution in both face-centered cubic (fcc) [101,102] and body-centered cubic (bcc) [103-105] metals, providing quantitative data about surface relief evolution and the dislocation structures of persistent slip bands. AFM has been used to investigate the deformation process of nickel-base superalloys during compression tests [106] and duplex stainless steel in in-situ tensile tests [107], and the number of dislocations responsible for a slip step has been determined based upon AFM data [107]. However, detailed systematic AFM data has not been used to quantify deformation system activity and local shear strains in a polycrystalline array in previous works. In the following sections, a newly developed technique combining AFM and EBSD-based trace analysis was introduced, to quantify the heterogeneous deformation system activity in different grains within microstructure patches.   3.2.1 AFM Measurement and Accuracy Check Using T1 Twins Two examples of AFM scans collected from grains 1 and 6 after ~1.5% strain are shown in Figure 3.7. Twins and slip traces are clearly visible in the topography maps. The surface steps in Figure 3.7(left) correspond to the (1012)[1011] T1 twin system identified by EBSD orientation analysis and slip trace analysis. The twins cause more uniform surface height change than the dislocation slip shown in Figure 3.7(right).   71     Figure 3.7 Examples of surface topography in grain 2 (left) and grain 3 (right) measured by AFM after deformation. Line section profiles reveal that much larger and more homogeneous surface steps result from twins (left) than from slip lines (right). Dashed line is interpreted as the undistorted surface inclination and serves as the basis for evaluating the overall height change along the section.   72   Prior to calculating the amount of plastic deformation from surface topography the accuracy of the AFM measurements needed to be checked. For this purpose, T1 twins (in grain 2) were chosen, since, in contrast to dislocation slip where many dislocations may traverse a given slip plane, only a single twinning dislocation with Burgers vector b passes on each adjacent twinning plane (separated by d along plane normal n ) during the formation of a twin, thus resulting in a precisely known amount of shear (0.17 for T1 twins [91]). Therefore, the resulting surface step height is fully determined by the crystallography and the thickness of a twin measured along n . Figure 3.8 illustrates the principle of the accuracy check. For a given twin system the number of displaced twin planes N follows from the measured surface step h and the twin Burgers vector b projected onto the surface normal ez: N= h b⋅e (3.1) z This can be compared to the alternative derivation based on the projected twin thickness tx, the twin plane normal n, and the inter-plane spacing d: N= t ⋅n x d  t ⋅n ≈ x d (3.2) The true projected twin thickness tx is approximately equal to the apparent, i.e.,   measured, projected twin thickness t , if h << t . Using the AFM “Section” data of the x x very left twin shown in Figure 3.7(left), the two alternatively derived values of N agree to within 2%. This shows that the AFM measurement process is robust, so that a similar analysis can be used to calculate the number of slip dislocations responsible for a   73         Figure 3.8 Schematic representation of a cross-section through a twinned portion of a grain. The twin trace on the sample surface is normal to the sketch (i.e., parallel to the Y direction). The surface displacement h due to twinning is determined from an AFM line section in the X direction. tx and ttrue are the twin width in the X direction (projected twin width) and the twin thickness along the twin plane normal direction, respectively.     74   series of parallel slip bands in a grain, such as the slip bands shown in Figure 3.7(right) that create relatively small steps on the surface. 3.2.2 Generation of Local Shear Distribution Maps Figure 3.9 illustrates the procedure for calculating shear on deformation system α from the individual surface steps along an AFM section line. By identifying the system α related to each surface step along a section line of length Xmn and cumulating the overall height change per system, the number of individual displacements occurring along Xmn can be calculated according to Equation 3.1. Provided that the overall surface height change across Xmn is small compared to Xmn, the average shear per deformation system is given by ( γ α = bα N α )( X mn ) ⋅ nα = (bα hα ) α b ⋅e z (X mn ⋅ nα ) (3.3) The area enclosed by the solid box in Figure 3.4(a) was scanned using AFM between each deformation step during four-point bending. The surface height changes caused by deformation systems (identified independently by SEM/EBSD trace analysis) in the microstructure patch within the solid-line box in Figure 3.4(a) at 1.5%, 3%, 6% global strain level were measured by an array of horizontal AFM “Section” lines passing through the center of each of 25 × 25 tiles spaced 10 µm × 10 µm. The typical measured step height is about two orders of magnitude smaller than this grid size. The measurement of local shears at different strain levels were all checked by repeating measurements using a vertical AFM “Section” line on 50 randomly chosen tiles, with   75     Figure 3.9 Schematic drawing showing the parameters associated with calculation of the local average shear caused by the activated deformation system with the slip plane normal n and Burgers’ vector direction b in sample coordinate system. h is the surface height change caused by this deformation system. reference grid increments in the X direction.     76   Xm(n-1), Xmn, and Xm(n+1) are measurements agreeing to within about 10% in the X and Y direction for each tile, as shown in Figure 3.10. The sum of all individual local shears ( ∑ γ α ) at 1.5% global strain is plotted for α each tile in the map shown in Figure 3.11(a). While the maximum shear is 0.17, the color scale of Figure 3.11(a) – (c) is capped at 0.14 (yellow) in order to reveal details and spatial heterogeneity in the local shear distribution caused by dislocation slip. The summed local shears in grains 3, 8, and 9, which deformed primarily by prism slip, is higher than strains observed due to basal or pyramidal slip in grains 5, 7, and 10. Shear maps for the individual deformation system types in the same microstructure patch at 1.5% strain are shown in the first row of Figure 3.12. Shear caused by prismatic slip in grains 3, 8, and 9 varied from 0.02 to 0.08, while in grain 5, 7, and 10, where basal and pyramidal slip were favored, the shear varied between 0.005 and 0.02. Grains with dominant prismatic slip exhibited about three times higher shear strains than grains with either basal or pyramidal slip activity. The shear in the grain with the T1 twin varied from 0 to 0.17, consistent with the range expected for the theoretical shear of the T1 twin being 0.17. Locally smaller values of shear arose due to the fact that tiles do not necessarily comprise the same (full) fraction of a twin; hence most of the tiles exhibited a shear less than 0.17. 3.2.3 Local Shear Distribution Evolution at Different Strain Level The local shear distribution evolution with the increase of global tensile strain can be observed by comparing the sum of all individual local shear distribution map shown in Figure 3.11(a) - (c). Detailed comparison the local shear caused by each kind of   77   Figure 3.10 The calculated local shear deviation at 1.5% global strain based on the measurements using both horizontal and vertical AFM “Section” lines for 50 randomly chosen tiles. The measurements agree within about 10% in both directions for all tiles.   78   Figure 3.11 The AFM-based experimentally measured sum of all individual local shear distribution map of the highly characterized microstructure patch shown in Fig. 3.4(a) at global strain level of (a) 1.5%, (b) 3%, and (3) 6%.   79     Figure 3.12 Individual local shear associated with different types of deformation systems in the microstructure patch shown in Fig. 3.4(a). The experimental measurements show that, in the early deformation stage, the local shear increases almost proportional to the overall tensile strain in all grains, while during further deformation (3% to 6%), the local shear in grains with pyramidal slip shows a much larger increase rate than that in grains with prism or basal slip.     80   deformation system at different strain levels is shown in Figure 3.12. In the early stages of deformation (1.5% to 3%), the local shear of all identified deformation systems increased almost proportional to the overall tensile strain in all the grains. As observed from Figure 3.12, the average local shear caused by prismatic slip in grains 1, 3, and 9 increased from about 0.04 to about 0.08 during this stage, and the average basal and pyramidal slip in grains 5, 7, and 10 also increased about 0.01. During further deformation (3% to 6%), the local shear in grains with pyramidal slip exhibited a much larger increase rate than that in grains with prism or basal slip. Figure 3.12 shows that the average local shear of prismatic slip only increase about 0.01, and basal almost no increase, while the local shear of pyramidal slip increased 100% in grain 10 (i.e. from an average shear of 0.005 to 0.012). This may be because the work hardening and dislocation saturation in the grains showing prismatic and basal slip in earlier deformation stages [108-110]. 3.3 Crystal Plasticity Finite Element Modeling 3.3.1 Constitutive Law The constitutive description of the material is based on a crystal plasticity formulation using the multiplicative decomposition of the total deformation gradient and considering the anisotropic elastic constants of α-titanium, as described in Section 2.5. Dislocation slip systems in the particular model used in this work were treated as bidirectional deformation systems. Twinning systems were incorporated as unidirectional dislocation slip systems, which results in diffuse straining, contrary to the discretely sheared twins observed in the microstructure. The grain boundaries were   81   transparent to slip (i.e., no specific resistance to slip at a grain boundary), which means that local shears (dislocations) can effectively transfer across boundaries. The values of CRSS and hardening parameters, as listed in Table 3.2, were chosen based on literature [85, 86, 92], and adjusted slightly to increase the stability of the calculation processes. The material model was implemented into the commercial FEM package MSC.Marc. 3.3.2 Mesh Generation and Boundary Condition Three different sizes of quasi-3-D meshes based on the microstructure in the dashed box in Figure 3.4(a) were generated and are shown in Figure 3.13(a) – (c), respectively. The meshes were developed by distributing nodes along the straight grain boundary traces, planar surface meshing of the enclosed grains, and expansion by 50µm into the third dimension, and then evenly discretizing by five elements. Hence, all grain boundaries are perpendicular to the surface in this approximation. To replicate the constraint from the surrounding bulk material, the simulated microstructure patches shown in Figure 3.13(a) and (b) were surrounded by a rectangular-frame shape rim, while the mesh shown Figure 3.13(c) was placed in a rectangular pan-like rim (to apply a the constraint from the material beneath the modeled the microstructure patch). The crystal orientation identified as dominant from macro-texture measurement (c-axis lined up 45° from the tensile axis) was assigned to the rim. A tensile face load was imposed on the right surface of the rim, with the left surface constrained in the X direction and the back surface constrained in the Z direction for all the three meshes.   82   Table 3.2 Hardening parameters used to describe evolution of the critical resolved shear stress. Deformation System {hkl} sα / MPa α sS / MPa α h0 / MPa a Prismatic {1010} < 1210 > 60 200 100 2 Basal {0001} < 1210 > 120 300 150 2 Pyramidal {1011} < 2113 > 180 500 300 2 T1 twin {10 12} < 1011> 125 200 150 2     83       Figure 3.13(a) – (c) Three different sizes of 3D meshes of the grain patch corresponding to the microstructure within the dashed box in Fig. 3.4(a).   84   3.3.3 Modeling Results Figure 3.14(a) – (c) maps the summed individual shears for the top surface resulting from simulated tensile loading to 1.5% engineering strain using the three meshes shown in Figure 3.13, respectively. The simulations using meshes (a) and (b) show a large deformation band across grains 6 and 9, while there is little deformation in other grains. This pattern of strain was similar on the front-side and back-side, but strains were slightly larger on the backside, despite the fact that the geometry of the mesh is the same, as shown in Figure 3.15. These two simulation results, as shown in Figure 3.14(a) and (b), do not reflect the experimentally observed heterogeneous deformation in this grain patch. For example, only grain 3 and 9 show about 0.1 prismatic dislocation shear at 1.5% global strain, in which local shear varied from 0.02 to 0.08 in experimental measurements. In other grains, of which the deformation is dominated by basal, pyramidal , and T1 twin dislocation shears, almost no dislocation shear was found in the simulation. The simulation results using mesh (c) shown in Figure 3.14(c) is much more realistic local shear distribution. Large plastic strain was found in grains 1, 3, 9, and 14. Grains 2, 8, 10, 11, and 13 have harder orientations, and show comparatively little shear. The highest local shear in the simulation of around 0.14 is found in grain 14. However, the local shear in grain 14 needs to be treated with caution because the large strain concentration in grain 14 might result from an edge effect from the surrounding rim. Since the artificially added rim was directly in contact with the right part of grain 14, and also grain 14 is close to the corner of the mesh, the deformation of grain 14 could have been easily influenced by the deformation process of this artificial rim, which is   85   Figure 3.14(a) – (c) The summed individual shears for the top surface resulting from simulated tensile loading to 1.5% engineering strain using the three meshes shown in Figure 3.13(a) – (c), respectively.   86     Figure 3.15 The comparison of the summed individual shears distribution on the frontside and back-side of the meshes shown in Figure 3.13(b).   87   different from neighboring grain deformation process. Figure 3.16(b) shows the distribution of simulated shears from prismatic slip, basal slip, pyramidal slip, and T1 twinning systems (note different scales) using the mesh shown in Figure 3.13(c). The shear caused by prismatic slip is mainly concentrated in grains 3, 6, and 8. In grains 3 and 9, prismatic shear is higher (with a maximum value of 0.052) in the right of the grain than in the left. Basal shear in the left of grain 10 and grain 11 varies from 0.02 to 0.03. Basal activity is also found in the corners of grains 5 and 7. The twinning activity in grain 2 was distributed similarly to the experimental observations; high activity correlates with twins in the bottom and center part of grain 2 (see Figure 3.4(a)). The CPFE simulation does not account for the (long-range) twin habit- plane effect present in the real specimen, where twins were primarily extended along their habit plane, or for the change in crystal orientation resulting from twinning, thus, a direct match of twin geometry between experiment and simulation was not expected. Pyramidal slip activity occurred mainly along the boundary between grains 3 and 9 in this microstructure patch. Detailed comparison between experimentally measured and simulated dislocation shear distributions of this microstructure patch will be discussed in Section 3.4.3. 3.4 Discussion 3.4.1 Comparison of Experimental Observations and CPFE Simulations The CPFE modeling results can be directly compared to the details of the experimentally measured shear distribution, as illustrated in Figure 3.16. The AFMderived maps of individual shears are effectively reproduced in the simulation, but the   88   Figure 3.16 Comparison of experimental measured (shown in row (a)) and FEM simulated (using mesh (c)), shown in row (b)) individual local shear associated with different types of deformation systems in the microstructure patch shown in Figure 3.4(a) at 1.5% strain. FEM simulation results illustrate that the spatial distribution of local shear caused by prismatic and basal slip agrees well with experimental measurement. Most of the local shear caused by T1 twinning is observed in grain 2. Simulation does not successfully simulate the pyramidal slip in grain 10 where AFM measurement reveals around 0.01 local shear.   89   finer details differ. The highest shear value in the simulation is about the same as that in the AFM measurement for the seven grains (2, 3, 6, 8, 9, 10, and 13) in the center of the patch. From the AFM measurements, however, the highest shear is in the left of grain 3, while the CPFE model showed this region to have less strain than the lower right of grain 14. Secondly, Figure 3.16(a) shows that near the boundary between grains 3 and 8, the shear caused by prismatic slip in grain 3 is about 0.05 – 0.07, while its value is lower in the simulation. The shear activity of basal slip varying from 0.005 to 0.02 in grains 5 and 7 is well captured. The simulation also successfully captured the basal activity in grain 10 both spatially and quantitatively. The simulation did not predict the activity of pyramidal slip in the right of grain 10, which contributed a shear of about 0.01 in the experimental measurements. The twinning activity in grain 2 was distributed similarly to the actual specimen; the twins are thicker near the boundary of grains 1 and 2, where the twin was activated by slip–twin transfer [111]. 3.4.2 Effect of Modeled Area Size in Simulation Results As shown in Figure 3.14, the FEM simulations do not successfully reflect the real local shear distribution of the microstructure patch using the mesh (a) and (b). There are several probable reasons for the differences between the experimental results and the CPFE modeling results. The most important among these is the modeled area size effect. The interested microstructure patch has a relatively “soft” grain cluster in the right side, making it highly possible that the plastic deformation tends to be concentrated in this band in the simulation if no surrounding grain orientation and shape information were provided. Two choices can be made to potentially solve this problem.   90   The easier one is to extend the modeled area and to add an artificial pan-shape rim around the modeled microstructure patch to simulate the constraint from the neighboring microstructures, which has been demonstrated in the simulation results using mesh (c). The local shear distribution in the middle part of the mesh is relatively successfully modeled, as discussed in the previous section. The edge effect resulting from the rim, however, still existed (see grain 14). This suggests that a mesh that can correctly reflect the surrounding environment may need to include at least one layer of real surrounding grain information, which leads to the second potential method to increase the accuracy of CPFE simulation, i.e. including more surrounding real grain information (both around and beneath the microstructure patch of interest) into the mesh. The grain information beneath the surface can be collected using the EBSD-FIB technique, i.e. 3D-EBSD, as described in Section 1.3.2. Also, a more efficient and stable constitutive model is needed if a real 3D mesh is used in the simulation. In cooperation with MPIE, future work has been focused on developing a new CPFE constitutive model that uses parallel calculation processes to largely increase the simulation efficiency.   3.4.3 Effect of Rim Crystal Orientation in Simulation Results Three simulation results (to 2.2% global strain) with different crystal orientations chosen for the rim are shown in Figure 3.17 to examine the effect of rim crystal orientation on the simulation results in details. The orientation in the simulation on the left is closest to the actual average grain orientation in the material measured using EBSD, and is the orientation used in Figure 3.14 and 3.16(b). The maximum shear,   91   Figure 3.17 CPFE model illustrating how the surrounding orientation (indicated with Euler angles) affects local shear distributions at 2.2% global strain.   92   shown in the top row in Figure 3.17, does not appear to be strongly influenced by the chosen crystal orientation. However, upon closer inspection, the orientation of rim orientation does affect the shear on each particular family of slip systems, as illustrated by the second, third, and forth rows, where prismatic, basal, and pyramidal slip activities differ in some locations, especially in several edge grains (for instance grains 5 and 11). When the rim has a hard grain orientation, a small increase of the local shear of prismatic, basal, and T1 twin can be observed in the simulation, especially the shear distributions in the grains near the patch edges. Less influence of the crystal orientation of the rim can be found on the simulated deformation activity in the center grains of the patch. The results suggests that accurate modeling of deformation system activity in polycrystalline patches requires at least one layer of correctly oriented and shaped grains as a buffer zone between the rim and grains of interest in the center. 3.5 Conclusions Heterogeneous deformation activity within a polycrystalline microstructure of αtitanium with a series of tensile strain of about 1.5%, 3%, and 6% was quantitatively measured using a technique combining AFM, EBSD, and BSE imaging. AFM provides high-resolution measurement of the z-components of the slip and/or twin displacements on sample surfaces, allowing fine-scale analysis of dislocation activity and derivation of corresponding local shears averaged on a 25 × 25 grid of square tiles. The experimental result at 1.5% global tensile stress was used to effectively assess the accuracy of CPFE simulations of deformation activity on different dislocation slip and twinning systems. The phenomenological model examined successfully predicted most   93   of active dislocation slip and twinning systems. In both experiment and simulation, grains typically exhibited a single dominant slip system. Grains with dominant prismatic slip exhibited about three times higher shears than grains with either basal or pyramidal slip activity. While the magnitude of shears was reasonably predicted within most grains, their spatial distribution frequently differed from the measurement. The crystal orientation of the rim was found to have a modest influence on the simulated dislocation activity in the grains in the patch center. Better agreement with the measurements may result from (1) improved three-dimensional representation of grain geometry; (2) experimentally measuring critical resolve shear stresses and hardening parameters for different deformation systems in this material; and (3) including realistic slip resistance at boundaries into the model.   94   CHAPTER FOUR CALIBRATING PARAMETERS FOR CPFE MODELS USING NANO-INDENTATION In this chapter, a combined study of the anisotropic nano-indentation response of α-titanium was conducted using nano-indentation, EBSD, AFM, and CPFE simulations to quantitatively determine the CRSS for different slip systems. This information is critical for developing more reliable values of the key parameters in CPFE simulations. Experimental nano-indentation studies, along with CPFE simulations of the nanoindentation response, provide an alternative opportunity to study the behavior of single crystals in a polycrystalline environment. Because in most cases the grain size is much larger than the size of the indentations, the nano-indention can be treated as the deformation of a constrained single crystal. Such experiments allow the separation of the influence of intrinsic crystal properties, such as grain orientations, from the influence of polycrystalline effects, such as grain boundaries and neighboring grain orientations. Analysis of the residual surface topography after indentation revealed the crystalorientation-dependent pile-up behavior of α-titanium. The CRSS values of different slip systems were then determined using a non-linear optimization procedure [120, 121]. The calculated CRSS were 150±4 MPa, 349±10 MPa, and 1107±39 MPa for prismatic, basal, and pyramidal slip, respectively. These values were then used to simulate the same microstructure patch discussed in Chapter Three, and no significant differences were found between the two simulation results, indicating that more parameters need to be considered to further improve the simulations.   95   4.1 Importance and Advantages of Measuring CRSS Using Nano-indentation As mentioned in the conclusions of Chapter Three, accurate values of the CRSS and hardening parameters of a material are one of the critical requirements to improve the CPFEM modeling results. One conventional way to measure the CRSS for many crystalline materials is to do a series of uniaxial tension tests on dog-bone shaped single-crystal specimens of the material with different orientations [112-114]. The orientations of the specimens are usually carefully chosen to only allow a single deformation system to be activated during the tensile tests so that the value of CRSS (the yield stress on the stress-strain curve) for this particular system can be measured. It is, however, difficult to apply this methodology to measure the CRSS of hcp metals, i.e. α-titanium, because slip systems with similar slip planes and/or slip directions, as well as twinning, in these metals make it hard to guarantee the activation of only one deformation system. This is the main reason for the large variation in the experimentally measured CRSS values of α-titanium in the literature, as shown in Table 1.1. Furthermore, the single crystals of titanium and titanium alloys are also expensive and difficult to produce. Indentation (hardness) testing is an alternative tool for measuring the mechanical response of crystalline materials. Some early literature [115-116] has demonstrated the anisotropic hardness responses of crystalline materials. The residual surface topography, known as the pile-up and/or sink-in, around the indent after indentation, as shown in Figure 4.1, was treated as a fingerprint of the deformation processes, which is closely related to the orientation of the indented single grains [117]. Recently, the development of the AFM and nano-indentation techniques provides a much more   96     Figure 4.1 Material pile-up (left) and sink-in (right) increases or decreases the contact height, hc, during indentation to the maximum indentation depth, hmax. [117]   97   convenient and precise way to characterize the mechanical responses and pile-up and/or sink-in profiles of different grain orientations within a polycrystalline environment [118,119]. Since the average grain size in the tested material in this work (Material A, see Section 2.2.1) is around 80 µm, which is much larger compared to the size of the nano-indenter (around 1 to 2 microns), the mechanical responses during nanoindentation in the middle of grains can be treated as a single crystal response. 4.2 Grain Orientation Dependent Behavior during Nano-indentation 4.2.1 Relationship between Hardness and Grain Orientation A rectangular array of applied nano-indentations on the microstructure patch in specimen S0-2 is shown in Figure 4.2, on which the corresponding OIM map is overlaid. Both soft (with c-axis perpendicular to the indentation axis) and hard (with c-axis parallel to the indentation axis) grain orientations are present in this patch, which provides a wide range of grain orientations to study. Figure 4.3 shows a color-scale map indicating the maximum indentation depth (with respect of the original surface height) resulting from all the indents within this region. The measured maximum depths were almost homogeneous inside a given grain, with only small scatter. The standard deviation of the maximum indentation depths was around 5% of the mean value within a given grain. However, the heterogeneous responses of the hard and soft grains can also be easily observed from Figure 4.3. Grain 2, with the hardest grain orientation in this microstructure patch, shows the lowest average maximum residual depth (274 nm), which is only about two thirds of the average maximum residual depth in the softest grain (420 nm). Also, the indentations close to grain boundaries usually show medium   98       Figure 4.2 The array of indents applied on the microstructure in specimen S0-2. Indentions were spaced by 20 µm × 16 µm. A semi-transparent OIM map with inverse pole figure coloring is superimposed on a backscattered electron image.   99   Figure 4.3 Color-scaled map of the maximum indentation depth (unit: nm) for the indentation array shown in Figure 4.2. Each tile represents one indentation in the middle of this tile.   100     values of the maximum depth of the two grains in both sides of the boundaries. Several load-displacement curves recorded from the indentations in the middle different grains are shown in Figure 4.4. These indentation curves also clearly indicate significant differences in the mechanical responses of these grains during indentation. The higher slope of the loading indentation curve for grain 2 shows that the grains with their c-axis nearly parallel to the indentation axis are more resistant to plastic deformation. Table 4.1 shows the maximum residual depths of selected indents from grains, along with different angles between their c-axis and indentation axis. These indents were chosen from the center of each grain to prevent any influence from grain boundaries. However, it cannot be guaranteed that there is no grain boundary existing close to the polished surface below the indents. The distance from the chosen indent to the grain boundary below it should be maximized if the grains are assumed to be equiaxed, which is reasonable according to the SEM images. A general trend of the increase of maximum residual depth, i.e. the decrease of hardness, was observed with the indentation axis rotating away from the c-axis (basal plane (0 0 0 1) normal). 4.2.2 The Relationship between Pile-up Topographies and Grain Orientation To further examine the grain orientation dependent behavior of α-titanium, the residual surface topography of the selected indents circled in Figure 4.2 were examined using AFM. Two examples of the AFM scans for indents are shown in Figure 4.5, which illustrates the different residual surface topography between soft and hard grains. To study this phenomenon in a systematic manner, the surface topography of these indents are presented on an inverse pole figure, and the position of each   101     Figure 4.4 Force-displacement curves from indentations conducted in the middle of soft grains 5 and 10, and hard grain 2. The maximum load is 6 mN. The lower inclination of the curve during loading stage for grain 2 indicates that grain 2 is more resistant to plastic deformation.   102   Table 4.1 The indentation axis in the crystal coordinate system, the angle between the c-axis and the indentation axis (η) and the maximum residual depth (hres) of the indents in the center of the selected indented grains in specimen S0-2. Grain Indent Euler angles, Indentation axis Number Number η, degrees hres, nm degrees 1 53, 66, 329 [12 -6 -6 5] 66 251 2 13 – i 108, 26, 278 [3 2 -5 10] 26 188 6 10 – m 77, 81, 261 [14 -5 -9 2] 81 258 7 6–a 95, 86, 239 [26 1 -27 2] 86 252 8 6–o 186, 51, 169 [4 1 -5 4] 51 211 10 6–h 76, 87,277 [9 6 -15 1] 87 259 11 6–k 268, 56, 90 [6 -3 -3 4] 56 228 14   13 – o 1–a 102, 114, 221 [14 -5 -9 6] 66 241 103     Figure 4.5 Indent surface topography scanned by AFM in soft grain G7 (left) and hard grain G2 (right), illustrating the grain orientation dependent behavior of pile-ups.   104   topography was determined by the indent axis of the corresponding grains, as described in detail by Zambaldi et al. [117]. Figure 4.6 clearly shows the general trend that all indented grains with their caxis rotated away from the indentation axis exhibit pile-up behavior that exhibits two dominant pile-up hillocks on opposite sides of the indents. It was also found that the line connecting the two pile-up hillock maxima beside each of these indents was roughly perpendicular to the c-axis of the corresponding indented grain. The most pronounced pile-up was on the (0 0 0 1) great circle, and the heights of the pile-ups decreased with decreasing angle between the indentation axis and the c-axis (η). Roughly, when the angle η was smaller than 30°, no obvious pile-up behavior was observed near the indents. 4.3 CPFEM Simulation of Indentation and CRSS Optimization 4.3.1 Constitutive law The CPFEM simulations were run by Claudio Zambaldi in MPIE. The constitutive description of the material was based on a crystal plasticity formulation using the multiplicative decomposition of the total deformation gradient and considering the anisotropic elastic constants of α-titanium, as described in Section 2.5. Dislocation slip systems, including prismatic slip, basal slip, and pyramidal slip, in this particular model, used in this work, were treated as bidirectional deformation systems. The CRSS values for dislocation slip systems were optimized following the procedures discussed in Section 4.3.3. Deformation twinning systems were not included in the simulation, since no nano-indentation induced twinning was observed   105     Figure 4.6 Indentation topography measured by AFM, with their corresponding indentation axes, positioned on an inverse pole figure (IPF). The indents in the upper half of the IPF are also displayed in the lower half by a mirroring operation with respect to the plane (1 0 -1 0) and are marked with an asterisk.   106   using both AFM and EBSD. The material model was integrated into the commercial FEM package MSC.Marc by using the subroutine hypela2. 4.3.2 Mesh Generation and Boundary Conditions Since the size of the indenter was much smaller than the grain size, grain boundary effects were neglected when the indentation is in the center of the grains, and the indented grains were treated as a single crystal. The grain orientations of the simulated grains were collected using EBSD, and the grains were discretized by 8-node hexahedral elements. Figure 4.7 shows a deformed mesh that sliced perpendicularly accross the center of the indent. A total of 4320 elements were applied. The deformation of the sphero-conical indenter was neglected by assuming it as a rigid body. The maximum loading force was 6mN, which is the same value as that used in the experiments. 4.3.3 Crystal Plasticity Parameter Optimization Procedures The values of the CRSS and saturation resistance for three slip types, i.e. prismatic, basal, and pyramidal slip, were chosen as the six free parameters. An optimization algorithm of the six variables was carried out using the downhill simplex method provided by Nelder and Mead [120,121]. The optimization of the crystal plasticity parameters using this method has also been demonstrated by Grujicic and Batchu [122]. The objective function, the function that is to be minimized by the optimization routine, was defined as the deviation of the simulated topographies from the   107   Figure 4.7 A deformed (indented) finite element mesh of hexahedral elements [117].     108   experimentally measured ones. With the help of Gwyddion software [123], the surface topography data obtained by AFM was extracted and regenerated to a square X-Y grid domain of 2.5 µm × 2.5 µm with a step size of 0.1 µm, and the simulated CPFE topographies were also exported to the same domain. The lowest point in AFM measured topography was aligned with the indentation axis in the simulation. The topographic contribution to the objective function of one indentation, Gtopo, was described as the sum of height differences between the measured and simulated topography at all points in the square grid: i, j i, j Gtopo = ∑ zexp − zsim = ∑ Δzi, j i, j (4.1) i, j For optimization of the six constitutive parameters using two indentation topographies in different grain orientations simultaneously, the objective functions for each indentation were summed. The initial values of the six parameters were randomly chosen, and 300 rounds of simulations were conducted to minimize the objective function, i.e., to converge the six constitutive parameters to the optimum values. More details can be found in Zambaldi et al.[124]. 4.3.4 Modeling Results The simultaneous optimization process was carried out for two selected indents 6-a (indentation axis close to [1 0 1 0]) and 13-o (indentation axis close to [2 1 1 1] ). Figure 4.8 shows the measured surface and simulated surface topographies using optimized parameters, as well as the Δzi, j map, of the two indents. Table 4.2 shows   109   Figure 4.8 Experimental (left), CPFE simulated (middle) surface topographies, as well as Δzi, j (right) maps, of indents 6-a and 13-o, which were selected for the identification of the constitutive parameters. The size of each map is 2.5 µm × 2.5 µm.   110   Table 4.2 Optimized constitutive parameters for prismatic, basal and pyramidal slip systems. The other constant parameters for the three systems in the simulations were chosen as: m = 20, h0 = 200 MPa, a = 2. Slip systems CRSS ratio Saturation resistance ,MPa Prismatic slip 150 ± 4 1 1502 ± 125 Basal slip 349 ± 10 2.3 568 ± 17 Pyramidal   CRSS, MPa 1107 ± 39 7.4 3420 ± 202 111   the optimized values of the CRSS and saturation resistance of prismatic, basal, and pyramidal slip after 300 iterations. Two other indents, 6-o and 13-i, were chosen to validate the identified set of parameters, as shown in Figure 4.9. In both validation simulations, the simulated topographies show good agreement with the measured ones. The total height difference ∑ Δzi, j is 10.9 µm for 6-o, and 9.4 µm for 13-i, respectively, i, j both of which are smaller than the total height differences for indents 6-a (12.2 µm) and 13-o (12.7 µm). Figure 4.10 shows the simulated pile-up behavior of the commercial purity titanium with respect to the grain orientation using the optimized values of constitutive parameters. All the simulations were performed at 6 mN load. Good agreement was found between the simulations and the experimental measurements. The maximum heights of the pile-ups decrease for those indents with their indentation axis closer to the c-axis. It was also found that the connection line of the two pile-up maxima for each indent was roughly perpendicular to the c-axis of the grain. 4.4 Discussion 4.4.1 Automated Identified Constitutive Parameter Values Initial optimization runs for parameter identification based on individual experimental measurements of a single indent were found not so satisfying. For example, the optimization can show bad convergence behavior when the Schmid factor for one slip system is too low, and consequently negligible amounts of shear were expected. This phenomenon was observed when running the optimization procedure   112   Figure 4.9 Experimental (left), CPFE simulated (middle) surface topographies, as well as Δzi, j (right) map, of indents 6-o and 13-i, which were selected for the validation of the set of optimized parameters. The size of each map is 2.5 µm × 2.5 µm.   113   Figure 4.10 Twenty-two simulated pile-up topographies positioned on the IPF according to their indentation axes. The pile-ups are shown in gray scale, and the regions below zero are shown in blue. The contour lines are starting at -0.04 µm and 0.04 µm for below and above zero regions, and drawn every 0.04 µm. Similar trends of pile-up features were found in CPFE simulations as that shown in the experimental measurements.   114   using indent 6-a (in the grain with an orientation close to [10 1 0] ). Basal slip systems in this grain orientation are difficult to activate because the basal plane is nearly parallel to the indentation axis. The convergence result shows large oscillations for basal slip. To avoid this problem, the sum of two indented surface topographies from two grains (with one having a soft orientation and the other having a relatively hard orientation), i.e. the sum of the two Gtopo functions of the two indents were chosen to run the optimization procedure, as described in Section 4.3.3. The consideration of the two surfaces in the same run guaranteed that all three slip systems have a relatively high Schmid factor in at least one indent. The identified CRSS ratios were within the range of the ratios found in the literature [85,86,125,126]. The determined CRSS value for the most active slip system, i.e. prismatic slip, was expected to have higher accuracy, since the surface topography is mainly influenced by the most active systems. The identified values of those less active systems, however, need to be treated with caution. In this study, to simplify the optimization process, only six main constitutive parameters were chosen as free parameters. Future study can include more parameters, such as cross-hardening coefficients, to increase the accuracy of the convergence process. 4.4.2 Comparison of Simulation Results using Two Sets of CRSS A rerun of the CPFE simulation of the deformation of the microstructure patch in specimen S45-1 discussed in Chapter Three was carried out. The originally chosen and identified sets of constitutive parameters were listed in Table 3.2 and 4.2. The rerun simulation was stopped at about 1% global strain due to the stability issue of the   115   existing model. This problem may be caused by the high ratio of the CRSS for prismatic and pyramidal slip in the indentified set of values, which may lead to very high local stress state in hard grains. These elements with high stress state can usually cause the calculation instability for the simulation. However, by comparing the prismatic and basal slip shear in the two simulations at 1% strain, no significant difference was found in the simulation results with identified parameters using nanoindentation. As discussed in previous section, six free variables may be enough for optimizing the deformation procedure in single crystals. However, in a more complicated situation, i.e. poly-crystalline environment, more variables need to be considered. For example, the optimization procedure for nano-indentation did not consider the T1 twin activity, which was found in poly-crystalline microstructure patches. Therefore, a further study is under way to optimize the simulation results using more free variables, i.e. the CRSS and saturation resistance of pyramidal slip. Two other possible reasons for the lack of improvement are that the loading conditions for indentation contain significant hydrostatic compression, which may affect slip resistance by non-Schmid stress components that affect slip activation [127]. Secondly, the lack of grain boundaries may also frustrate dislocation nucleation processes for non-prism slip [121]. Furthermore, the CPFE simulation does not contain any form of additional slip resistance across a grain boundary [62]. Simulation of the sub-surface grain geometry with an accurate 3D mesh may also possibly affect model accuracy. Improvement in modeling slip behavior near grain boundaries appears to be necessary to improve the agreement between experiment and simulation.   116   4.5 Conclusions A nano-indentation technique is introduced to help research the values of CRSS and saturated resistance of prismatic, basal, and pyramidal slip systems in commercial purity titanium. Arrays of conical nano-indentations were performed on patches of grains with a wide range of orientations, and AFM and EBSD scans of on the patch indicates that the two-hill-lock shape pile-up topography was strongly crystallographically dependent. Corresponding CPFE simulations predicted the pile-up patterns in good agreement with the experimental measurements. The six constitutive parameters were automatically identified by optimizing the simulation results of the indentation process in different grain orientations using a non-linear optimization method. A rerun of the CPFE simulation of the same microstructure patch studied in Chapter Three with the optimized parameters, however, showed only minor differences, which indicates that more quantitative studies on strain hardening behavior and slip transfer across grain boundaries are strongly needed.   117   CHAPTER FIVE MEASURING MESO-SCALE HETEROGENEOUS DEFORMATION IN COMMERCIAL PURITY TITANIUM In this chapter, a novel approach is established to determine plastic zone shape and meso-scale plastic strain distribution at the tip of a fatigue crack by combining BSE imaging and discrete Fourier transformation (DFT) image processing. An in-situ uniaxial tensile calibration test indicates that the full width at half maximum (FWHM) of the central peak of the DFT of BSE images has a close relationship with the induced plastic strain within selected grain patches. The plastic zone size and residual strain distribution around a crack tip were determined by applying this technique to an array of BSE images around the tip of a fatigue crack in a unloaded commercial purity titanium specimen. The measurement results were compared with the plastic zone size and shape measured using thermoelastic stress analysis, which can measure plastic zone size based on the heat generated during the plastic deformation processes. The experimental results were further compared to theoretical estimates of plastic zone size proposed by Dugdale and Irwin theories. Because of the good agreements between the two experimental results, as well as between the BSE-DFT method and the theoretical estimates, the BSE-DFT approach has been shown to be an effective way to measure meso-scale heterogeneous plastic deformation in deformed polycrystalline samples.   118   5.1 Observation of Microstructure around a Fatigue Crack Tip using BSE Images The fatigue specimen was cut out from Material B, and its configuration was illustrated by Figure 2.5. After 27,200 fatigue cycles, the fatigue specimen was unloaded from the fatigue test instrument, and an array of 15 × 15 (225 in total) BSE 2 images, as shown in Figure 5.1, were taken to cover a square area (about 7.84 mm ) around the crack tip of the sample at an instrument magnification of ×500. Each image was 115 μm × 115 μm with a center-to-center distance of 187 μm, and each image contains about 150 grains. A gradual change in the sharpness of the BSE images with distance from the crack tip can be observed, as shown in Figure 5.2. At large distances from the crack tip, e.g. shown in Figure 5.2(m14), the images display sharp, welldefined features allowing easy identification of the individual grains, which tend to exhibit uniform brightness levels that are distinctly different to their neighbors. Strong contrast variations are evident between the grains in those regions far away from the crack tip because the different grain orientations result in large differences in channeling behavior. Images acquired close to the crack tip, however, show significant variation in brightness within individual grains, making the individual grains less distinct, and consequently the images display less sharp contrast, as shown in Figure 5.2(h4). This is a result of significant local crystal rotations and residual plastic strain within these grain patches, as GNDs and statistically stored dislocation (SSDs) accumulate within grains. Grain boundary ledges, slip lines, and secondary inter- and/or trans- granular cracks were also observed in the grains close to the fatigue crack. The apparent loss of distinct individual grain definition, the decrease in image sharpness, and increase in the   119   Figure 5.1 The around-crack-tip region was divided by a 15 × 15 (225 in total) grid as illustrated, and each BSE image was taken from the middle part of each tile.   120   Figure 5.2 Upper array (shown in Figure 5.1) of BSE images around the crack tip at a center-to center spacing of 187um. The crack tip is shown in image h4. The images in near-crack-tip region show more lattice rotation and contrast variation, indicating more residual plastic deformation.   121   intra-granular features in the BSE images is consistent with increasing residual plastic strain with proximity to the crack tip. 5.2 Image Analysis Technique Using DFT 5.2.1 Basic Information of DFT The discrete Fourier transformation (DFT) provides a measurement of the frequency content and distribution of a signal. In two-dimension situation s, i.e. BSE images, the DFT of an m by n matrix X is given by another m by n matrix Y: Yp+1,q+1 = m−1 n−1 jp kq ∑ ∑ ω m ω n X j+1,k +1 (5.1) j=0 k =0 where ω m = e−2π i/ m and ω n = e−2π i/ n are the complex roots of unity. A Fourier transformation of a two-dimensional image gives a complex-number image, i.e. two images are generated, identified here, for the sake of ease of understanding, as the magnitude map and the phase map. The magnitude of a DFT provides the distribution of the frequency components in the original image, while the phase of the DFT gives the location of this frequency component in the original image. In the proposed technique, the magnitude of the DFT is considered and will be shown to reflect the contrast change within microstructure patches resulting from plastic deformation. 5.2.2 DFT Image Data Processing While the phenomenon of the contrast change in BSE images during plastic deformation, as discussed in Section 5.1, has been widely observed [48, 128], it has never been used to quantify the strain levels associated with localized plastic   122   deformation. Nevertheless, these contrast changes in BSE images can eventually lead to a change in the signal frequency distribution of features within the corresponding images. Therefore, if the frequency distribution of the BSE images can be quantified, this may be related directly to the extent of plastic deformation. DFT was conducted on all BSE images from the array (Figure 5.1) using MATLAB (MathWorks, Natick, MA, U.S.A) built-in functions. First, the MATLAB function ‘fft2’ was used to compute the complex-number DFT images of the BSE images using the fast Fourier transformation (FFT) algorithm. Functions ‘fftshift’ and ‘log2’ functions were then applied to improve the output image quality, shifting the zero frequency to the DFT image center, and enlarging the detailed frequency information, respectively. Figure 5.3(b) shows FFT magnitude images of four BSE images (Figure 5.3(a)) collected at 561 µm intervals directly ahead of the crack. Figure 3(b) shows the magnitude of the DFTs of the corresponding images, and a distinct trend in the extent of the central peak is apparent. The full width at half maximum of the central peak in the horizontal line profiles, shown in Figure 5.3(c), decreases with the increasing distances from the crack tip, indicating a potential relationship between the central peak width and residual strain level. To more accurately quantify the change of DFT with respect of levels of residual plastic strain, symmetry of the DFT plot was assumed and the left half of the line profiles crossing the center of DFTs were smoothed using the adjacent-averaging method with a five-point window. The smoothed profiles were then fitted to exponential curves in order to measure the (FWHM), as illustrated in Figure 5.4. Figure 5.5(a) shows the FWHM distribution directly in front of the crack tip, and Figure 5.5(b) illustrates the   123     Figure 5.3 (a) BSE images h5, h8, h11, and h14 from Figure 5.2; (b) their corresponding DFTs and (c) a horizontal line profile across the center of each DFT showing a decrease in the center peak width with increasing distance from the crack tip to the image location.   124   Figure 5.4 Schematic plots showing the process of fitting line profiles and measuring the full width at half the maximum of the profile. Line profiles were first smoothed using an adjacent-averaging five-point window. The left half of a smoothed profile was then fitted to an exponential curve.           125   (a) (b) Figure 5.5 The full width at half maximum of the FFT center peak versus the distance from the crack tip at (a) θ = 0° (b) θ = 45°   126   FWHM distribution 45° away from the crack line. The FWHM decreases dramatically with increasing distance from the crack tip, while the rate of change decreases with increasing distance, which further indicates the potential relationship between the FWHM in DFTs and the local residual strain levels where the corresponding BSE images were taken. 5.3 Quantifying the Relation between FWHM and Strain Using In-situ Tensile Test The results of the in-situ uniaxial tensile test were used to quantify the relationship between the residual plastic strain level in a given microstructure patch and the FWHM of the DFT magnitude. In order to obtain a detailed stress-strain relationship, a separate tensile test was carried out on the large dog-bone specimen using a conventional servo-hydraulic test machine. The engineering stress-strain curve generated (shown in Figure 5.6) shows that the material yields at about 0.4% global strain. In figure 5.6 the arrows below the stress-strain curve indicate the five strain levels at which the in-situ tensile test was interrupted to acquire the BSE images in six microstructure patches. Figure 5.7 shows one set of BSE images of a microstructure patch at different strain levels. These images were subsequently processed using the procedure described in the previous paragraph. The mean FWHM of the DFT from six microstructure patches at each strain increment were plotted, as shown in Figure 5.8, and a polynomial fitted to create a calibration curve (shown as dotted line in Figure 5.8) that could be used to determine the residual plastic strain field of the fatigued specimen. The calibration curve in Figure 5.8 was used to estimate the level of plastic strain in each of the 225 BSE images shown in Figure 5.2 after applying the DFT and calculating   127   Figure 5.6 Engineering stress-strain curve obtained for the commercially pure titanium; the arrows indicate the strain increments at which the in-situ uniaxial tensile test was interrupted to collect BSE images.   128   Figure 5.7 BSE images of a grain patch at different global strain levels during in-situ tensile test indicating the evolution of microstructure features.   129   Figure 5.8 The full width of half the maximum of the DFT as a function of the engineering strain based on images recorded at increments of plastic strain during an in-situ tensile test.     130   their FWHM. The elastic strain contribution was ignored because its contribution to the overall strain is very small compare to the measured plastic strains. 5.4 Residual Strain Map around a Crack Tip in the Fatigue Specimen Figure 5.9 shows a map of residual plastic strain field in the fatigued specimen based on the calibrated DFTs. Thus, each tile in the map represents an average residual plastic strain for a 187 μm × 187 μm area. This strain map clearly reveals the spatial heterogeneity of the local plastic strain distribution around the crack tip in the fatigue specimen. The shape of the plastic zone around the crack tip is an oval-shaped area with high strain extending about 1.5 mm ahead of the crack and about 0.5 mm on both sides of the crack. Also, the plastic wake strain is exhibited along the crack. The wake strain at the crack tip has a maximum value of 18%, which decreases dramatically with distance from the crack tip. 5.5 Discussion 5.5.1 The Influence of Brightness, Contrast, Magnification, and Calibration Procedures on the Strain Map The effects of the instrument parameters associated with the electron imaging conditions need to be assessed before further discussion. Theoretically, changes in brightness will not change the shape of the peak in the DFT, because the frequency distribution of the image remains the same. Increases in brightness will still increase the counts within the DFT peak, which means that the peak will be extended along the y-axis, but the FWHM will not be changed, as illustrated in Figure 5.10. An increase in   131   Figure 5.9 Map of plastic strain around the crack tip (the area shown in Figure 5.1) based on the FWHM of the DFTs of BSE images.   132   Figure 5.10 BSE images from the same microstructure patch and their corresponding DFT line profiles showing that the change of brightness does not affect the FWHM of the line profiles.   133   contrast will increase the difference in brightness level between bright and dark pixels, but will not change the image frequency distribution as measured by the FWHMs, as shown in Figure 5.11. However, the magnification of a BSE image can strongly influence the DFT peak because a different field of view will be present for the same size of image. Thus, a low magnification BSE image will contain more grains and the channeling bands within the grains will be finer in scale than for a higher magnification image of the same sample, resulting in a wider DFT distribution for the lower magnification image. Of course, if the sizes of features within an area of interest are finer than the pixel size in the image, then information will be lost, because the frequency of those features is too high to be differentiated using the DFT, i.e. Nyquist sampling errors will occur [129]. One the other hand, when the magnification is too high, there will be insufficient grains sampled to provide statistically significant data, and the DFT may become asymmetric about the DFT image center or too rough. In this situation, the assumption of symmetric curves in DFT images and the curve smoothing procedure using in Section 5.2 will be inappropriate. Therefore, it is important that images are acquired at an appropriate magnification for the grain size of the material being studied. Working distance will also influence the DFT distribution. A longer working distance may result in smaller rocking angle of the electron beam to image the same area, which leads to narrower SACP bands for a given grain. As a result, for the same microstructure patch at a given plastic strain, longer working distance will decrease the FWHM of DFT. Thus, it is also essential that the optical arrangements, i.e. magnification and working distance in the SEM, remain constant for both the calibration and fatigue samples in order to avoid   134   Figure 5.11 BSE images from the same microstructure patch and their corresponding DFT line profiles showing that the change of contrast does not affect the FWHM of the line profiles.   135   introducing errors. It is also important to note that the strain determined is an average for the field of view of the image so a smaller field of view would allow higher spatial resolution for the evaluation of strain (with associated errors outlined above). The strain calibration procedure is another aspect that needs to be considered briefly. The assumption in the DFT calibration presented in Section 5.3 is that an equivalent breakdown in channeling conditions, associated with a given distribution of GNDs, occurs for the same equivalent strain, regardless of the strain path experienced. Indeed, the two strain paths in the present work are considerably different. A uniaxial strain distribution with measurements was taken during monotonic loading for in-situ tensile test in SEM. In fatigue testing, a tri-axial strain distribution at the crack tip was induced by high-cycle tension-tension fatigue with measurements taken after loading in the SEM. In a study of aluminum alloy systems, Davidson and Lankford [130] proposed that the applied strain ( ε t ) – dislocation density ( ρt ) relationship in tensile test can be described as ⎛ σ ρ s/ 2 ⎞ εt = ⎜ 1 t ⎟ ⎜ σ 0 gs ⎟ ⎝ ⎠ m (5.2) where m has a value of 10 for aluminum alloys, g is a proportionality constant, and the parameters s and σ1 σ0 are dependent on the alloy systems and loading conditions. On the other hand, the relationship between cyclic strain ( Δε ) applied N times and dislocation density ( ρ f ) relationship in fatigue test was described as   136   ⎛ σ ρ p/ 2 ⎞ f f ⎟ Δε = 2 ⎜ ⎜ 2K gp ⎟ ⎝ ⎠ n where parameters p and (5.3) σf 2K are dependent on the alloy systems and loading conditions [130]. From Equation 5.2 and 5.3, with the assumption of the same dislocation density in the two kinds of deformation processes, the relationship between ε t and Δε can be derived. Along with another previous study [131], the results indicate that fatigue specimens can be expected to have a higher dislocation density for an equivalent strain level than uniaxial tension specimens. Thus, the strain deduced for the fatigue sample based on the calibration using the uniaxial sample, is probably an over-estimate of the actual strain to some small extent. However, the strain distribution observed directly ahead of the crack tip shown in Figure 5.9 is very similar to the observations and conclusions of Crompton and Martin [46]. 5.5.2 Comparison between the Plastic Zone Size Measured by BSE-DFT Method and Those Proposed by Other Works Data obtained using thermoelastic stress analysis along a line through the crack tip and along the crack path is shown in Figure 5.12. The black area indicates the measured plastic zone shape ahead the crack tip. The TSA signal and its phase difference with respect to the distance from the crack tip are plotted. The phase signal is noisy ahead of the crack tip but the far-field (about 2 mm away from the crack tip) has a mean value that deviates by less than five degrees from zero, indicating essentially   137   Figure 5.12 Phase difference and magnitude of the TSA signal as a function of distance along a line through the tip of the fatigue crack in the direction of crack growth. The extent of the plastic zone found from the phase difference map is superimposed as a black area.   138   adiabatic conditions, as would be expected for elastically deforming material. Close to and ahead of the crack tip, the phase difference exhibits a large negative value with a maximum difference of about 22 degrees, indicating large plastic deformation within this region. Immediately behind the crack tip, the phase difference is positive with a peak of about ten degrees. These two regions of phase difference with opposite sign have been attributed [98] to heat generation occurring ahead of the crack tip in the loading part of the cycle as a result of dislocation generation and movement and to heat generation behind the crack tip in the unloading part of the cycle due to contact of the recently-formed flanks. Therefore, as suggested by Diaz et al. [97], the transition between positive and negative phase difference provides an indication of the location of the crack tip along the direction of expected crack growth. Patki and Patterson [98] applied a binary filter to the phase difference and identified the shape and size of the plastic zone from the region of significant phase difference ahead of the crack tip. This is the approach adopted here and the black area superimposed on Figure 5.12 shows the cyclic plastic zone shape. According to the TSA data, the plastic zone shape is approximately circular with a radius about 0.7 mm. Patki and Patterson [98] compared the plastic zone size rp estimated using TSA with those determined using the theory proposed by Irwin [132]: ⎡ ⎤ 1 K 2r = ⎢ I ⎥ y π ⎢σ ⎥ ⎣ y⎦ 2 (5.4) where 2ry is the Irwin plastic zone diameter, KI is the applied stress intensity factor, and σy is the yield stress of the material. However, instead of using KI, ΔKI, the stress   139   intensity factor range, can usually give a better estimation of the plastic zone size [98]. The agreement between the plastic zone size measured using TSA and the estimated plastic zone size using Irwin’s expression (using ΔKI instead of KI) is usually within the 0.2mm, as found by Patki and Patterson [98]. This value of ΔKI was found using the Multiple Point Over-Deterministic Method (MPODM) to fit a Muskhelishvili-type description of the crack tip stress field to the measured TSA data following the method described by Diaz et al. [97], which was implemented in the software algorithm, FATCAT. Thus, in Figure 5.13, a comparison has been made of the crack tip plastic zones determined from the BSE-DFT and TSA data obtained in this study and based on the expressions proposed by Irwin [132] (see Equation 5.4) and Dugdale [133] (as described by Equation 5.5). 2 π ⎡ KI ⎤ ⎥ [133] 2ry = ⎢ 8 ⎢σ y ⎥ ⎣ ⎦ (5.5) The plastic zone found from the TSA data is smaller than that determined from the BSE images; but is very similar to the estimation using Irwin’s expression. The boundary of the plastic zone measured by TSA is of the same order of magnitude with its boundary in the direction perpendicular to crack growth, corresponding to about 4% plastic strain measured from the BSE images. Ahead of the crack, the boundary of the plastic zone found from the TSA data corresponds to about 8% plastic strain evaluated from the BSE images.   140   Figure 5.13 Map of plastic strain around the crack tip based on the FWHM of the DFTs of images from the region shown in Figure 5.1. The white line shows the measurement of plastic zone using TSA data. The green and blue solid lines indicate the theoretical estimations of the plastic zone size based on Dugdale’s and Irwin’s approaches, respectively.   141   The two experimental approaches are based on different physical principles, so differences in results would be expected. The data from the BSE-DFT method is based on detecting the local rotations of the grains by the geometrically necessary dislocations required for plastic deformation. In theory, this disruption in grains necessary for the observations in the SEM may not disappear when the load is removed. The removal of the loading, however, could have caused some reverse plasticity as a consequence of compatibility requirements between the crack tip plastic zone and the surrounding elastic region. This reverse plasticity can influence the measurement of the plastic strain distribution in the region closest to the crack tip but will have less effect on the plastic zone size. The evaluation of the plastic zone from the TSA data, on the other hand, is based on identifying the region in which the generation of heat by dislocation generation and movement is sufficiently large to cause adiabatic conditions to be lost, i.e. a heat flux or transfer occurs. A higher density of dislocations will generate a larger heat flux, which will be easier to detect as a phase difference, and this will tend to occur closer to the crack tip. At lower loading frequencies, though still high enough to maintain adiabatic conditions in the elastic field, there is more time for heat transfer to occur and so the TSA measurements, based on phase difference, will show greater sensitivity to the heat generated by plastic deformation; consequently, a lower density of dislocations is likely to generate sufficient heat to produce a measurable phase difference and so lower levels of plasticity will be detectable. This agrees with the observations by Tomlinson et al. [134] that the plastic zone size measured using TSA was smaller at higher frequencies. So a smaller plastic zone size measured by TSA   142   than that measured by BSE-DFT method could be expected for a high-frequency fatigue specimen. On the other hand, there is a close relation between the size of the plastic zone found using BSE-DFT measurements and the size estimated using Dugdale’s approach (Figure 5.13). Ahead of the crack, the boundary of the plastic zone boundary estimated using Dugdale’s approach corresponds to about 5% plastic strain evaluated from the BSE images. In the direction perpendicular to crack growth, the plastic zone boundary is corresponding to about 2% plastic strain measured from the BSE images. 5.5.3 The Application of BSE-DFT Method BSE imaging is a common imaging mode in the SEM for studies of metals, so this approach might be easily applied to the measurement of meso-scale heterogeneous plasticity in other kinds of deformation, such as bending, compression, and in creep tests. Also, the image collection process has the potential to be easily automated using a computer-controlled stage and automated image capture increase the collection speed. However, there are some restrictions in terms of the types of materials suitable for this method. First of all, in order to acquire statistically significant information, there needs to be sufficient grains in the field of view. Therefore, materials with large grain sizes will decrease the spatial resolution of this method. Another consideration is that the materials need to be single-phase, or multi-phase with large primary-phase grains and very small secondary-phase particles along grain boundaries and/or in grains, so that the DFT is effectively applied only to the primary-phase grains.   143   Otherwise, in materials with approximately equal amounts of two different phases, the contrast of BSE images will tend to be dominated by the difference between the phases instead of the differences in grain structure arising from GNDs, so that the DFT will be a function of the shape and quantity of these phases, which will mask the contrast associated with lattice curvature. 5.6 Conclusions The plastic strain around a fatigue crack in commercial purity titanium has been successfully evaluated using a novel technique based on using the discrete Fourier transform to analyze the structure of BSE images. The full-width at half maximum of the magnitude of the DFT of the BSE images was used as a measure of the level of plasticity based on the distortion of grains caused by dislocation accumulation during plastic deformation. The technique was calibrated using images obtained from a uniaxial tension test performed in-situ in the SEM so that the magnitude of the average plastic strain in the field of view of an image could be related to the FWHM for the image. After a fatigue crack was propagated to a length of 5.88 mm, 225 BSE images 115 µm × 115 µm were collected in an array occupying a 2.8 mm by 2.8 mm square around the crack tip and a map of plastic strain evaluated for this area. The measured plastic zone size and shape were assessed during the fatigue loading using TSA. The results from the two techniques are in general agreement and correlate with estimates of plastic zone size based on the theoretical expressions due to Irwin and Dugdale evaluated using the measured range of the stress intensity factor during the fatigue cycle of the stress intensity factor. This new approach may prove to be useful for   144   measuring local heterogeneous plastic deformation in other applications when there is a dominant primary phase present. Also, it has great potential in assisting identification of the role of the crack tip plastic zone in crack closure and plasticity-induced shielding.   145   CHAPTER SIX BROAD IMPACTS OF THE STUDY Multi-scale characterizations and simulations of deformation processes of advanced materials, such as titanium, are strongly needed to support all other relevant technological innovations. Success in understanding damage nucleation in design environments will lead to reduced waste, accelerated time to market for highly valueadded manufactured goods, improved safety and economy in all aspects of technology and engineering. The studies discussed in this dissertation combine several experimental and computational efforts to study how microstructural details affect the plastic deformation in polycrystalline specimens. These studies provide a solid foundation leading to the further understanding of the relationships between the heterogeneous deformation, often considered as a precursor of damage, crack nucleation and propagation. This research project has impacts in a number of important areas of both materials science and mechanical engineering research, including grainboundary engineering, dislocation dynamics, multi-scale finite element modeling, and fracture mechanics. In past decades, there has been considerable interest in developing advanced experimental methods and computational heterogeneous deformation in metallurgy. models to investigate micro-scale In previous studies, heterogeneous deformation was often assessed using tools such as orientation imaging microscopy and strain mapping. However, such studies have rarely analyzed the operating deformation mechanisms in the context of dislocation activity. In this dissertation, new   146   experimental methods (see Chapter Three) have been established to investigate the strain distribution and surface topography in deformed specimens in terms of activated dislocation systems, which provides an improved approach to study heterogeneous deformation. Experimental characterization indicates that, at the micro-scale, grains with different orientations respond to the same global stress states in rather different ways, including activated dislocation systems and lattice rotation gradients. Also, the experimental study on the values of CRSS and saturated resistance stresses for three deformation systems (see Chapter Four), i.e. prismatic slip, basal slip, and pyramidal slip, is an important part of further understanding dislocation activity in hcp metals. Although many previous simulation works have been conducted to study metal deformation processes, very few have been critically assessed by testing the simulation against actual microstructural deformation. The experimental characterizations in this dissertation have proved to be efficient tools to help assessment and development of the existing CPFE models, which may lead to a deeper understanding of material deformation and damage processes. With the help of these tools, the importance of different efforts, i.e. building real 3-D meshes and adding grain boundary effects, can be evaluated. Once good agreement between experimental and computational microstructural deformation is reached, the modeling can simulate details in the deformation behavior that cannot be directly measured, such as the local stress tensor. Thus, the details of deformation processes that occur prior to observed damage can be identified, and hence, the cause of damage nucleation can be quantitatively identified.   147   From a mechanical engineering point of view, the mechanical behavior of solids on the macroscopic scale is usually concerned. It sometimes ignores the discrete nature of materials, which results from microstructural variations in real materials. Some classical theories, such as classic identification of the plastic zone size in front of a crack tip, may be found satisfactory when the dimensions of the bodies of interest can be compared to the characteristic lengths, i.e. grain sizes. However, in poly-crystalline environments, grain sizes are much smaller than the specimen size, so microstructural variation must be considered to further understand mechanical behaviors of materials. In this dissertation, a new experimental method (see Chapter Five) successfully related micro-scale feature evolution, i.e. lattice rotation and dislocation activity, to the mesoscale plastic strain distribution. This method provides an opportunity to further develop classical theories based on crystal nature of the metals. Also, this method has a potential opportunity to be automated to scan large areas to quick locate microstructure patches of interest, i.e. strain concentrated area. These patches can be then further studied using other micro-scale experimental characterization methods.   148   CHAPTER SEVEN CONCLUSIONS AND FUTURE WORK 7.1 Conclusions Different scales of heterogeneous deformation in commercial purity titanium (αtitanium) have been studied in both four-point bending and fatigue specimens. A series of comprehensive experimental and CPFE simulation methods have been conducted to characterize and model the deformation behavior of different microstructure patches. Two new experimental techniques were established in this study to assist detailed observation and quantification of heterogeneous deformation in poly-crystalline specimens: • A technique combining AFM, EBSD, and BSE imaging was successfully applied to a polycrystalline microstructure to provide the fine-scale analysis of dislocation activity and derivation of the evolution of the corresponding local shears at different levels of global strains, based on the high-resolution measurements of the z-component of slip and/or twin displacements on specimen surface. Publication: Yang, Y., Wang, L., Bieler, T.R., Eisenlohr, P., Crimp, M.A., 2011, Quantitative AFM characterization and crystal plasticity finite element modeling of heterogeneous deformation in commercial purity titanium, Metall. Mater. Trans. A, 42A, pp. 636-644. • A technique based on the analyze of the frequency magnitude distribution in BSE images, by applying the DFT to the images of polycrystalline patches,   149   was introduced to map the meso-scale plastic strain distribution resulting from heterogeneous plastic deformation in complex loading and component geometries. The measured plastic zone size and shape from this technique has good correlation with estimates of plastic zone size based on the theoretical expressions of Irwin and Dugdale. Publication: Yang, Y., Crimp, M.A., Tomlinson, R., Patterson, E.A., 2011, Quantitative measurement of plastic strain field at a fatigue crack tip, Proc. Roy. Soc. A, in revision. Two sets of CPFE simulations in both single- and poly- crystal environments were conducted in this study. These simulation results were compared to the experimental measurements to assess the accuracy of the existing constitutive model and to reveal the possibility of further developments of the model: • The existing CPFE constitutive model examined, using a quasi – 3D mesh, successfully predicted most of active dislocation slip and twinning systems in a poly-crystalline microstructure patch. Most of the magnitude of dislocation shears were reasonably well predicted on a grain-to-grain basis, according to the comparison of experimentally measured and simulated local shear distribution. Simulations indicated that grains with dominant prismatic slip show about three times higher local shears than grains with either basal or pyramidal slip activity at 1.5% global strain. The dislocation shear distribution within individual grain, however, sometimes differed from the measurements. Also, the crystal orientation of the simulation mesh rim   150   was found to have a small influence on the simulated dislocation activity in the grains in the patch center. Publication: Yang, Y., Wang, L., Bieler, T.R., Eisenlohr, P., Crimp, M.A., 2011, Quantitative AFM characterization and crystal plasticity finite element modeling of heterogeneous deformation in commercial purity titanium, Metall. Mater. Trans. A, 42A, pp. 636-644. • CPFE simulations predicted the pile-up behavior in single-crystals during nano-indentation in single-crystals in good agreement with the experimental measurements within a wide range of grain orientations. The values of CRSS and saturation resistance for prismatic, basal, and pyramidal slip of α-titanium were automatically identified by optimizing the simulation results of the indent surface topographies in different grain orientations using a non-linear optimization method. A rerun of the CPFE simulation for the same poly-crystalline microstructure patch mentioned in previous paragraph with the optimized parameters showed only minor differences in results, which indicates further quantitative studies on strain hardening behavior and slip transfer across grain boundaries are needed. Publication: Zambaldi, C., Yang, Y., Bieler, T.R., Raabe, D., 2011, Orientation informed nanoindentation of alpha-titanium: indentation pile-up in hexagonal metals deforming by prismatic slip, J. Mater. Res., in print. 7.2 Future work To gain a deeper understanding of heterogeneous deformation in poly-crystalline   151   hcp materials, as well as to apply the obtained knowledge in this study to the prediction of crack initiation sites, series of further experimental and simulation studies are suggested: In the experimental aspect, future work can be focused on the observation of dislocation activity near grain boundaries using ECCI in the early stages of plastic deformation, a technique providing a smaller scale observation of dislocation activities near grain boundaries or within individual grains, compared to the techniques introduced in this dissertation. For example, ECCI images can be taken at grain boundaries where grain boundary ledges form at small strain level (1-2% strain) to observe the dislocation distribution near those ledges. Also, a statistical study on the relationship of the grain boundaries cracked in the first place and the deformation processes of the grains on the both sides of those boundaries is needed to build the connection between heterogeneous deformation and crack initiation. In the CPFE simulation aspect, a more comprehensive CPFE constitutive model that considers the information of slip-slip and slip-twin transfer across grain boundaries, as well as slip resistance at grain boundaries, is strongly needed to simulate heterogeneous deformation within grain patches more reliably. Also, a more detailed set of experiment is recommended to measure the critical resolve shear stresses and other hardening parameters for different deformation systems in titanium. Thirdly, an improved three-dimensional representation of grain geometry is also recommended to further increase the accuracy of simulation results. Three-dimensional information of grain patches, including both grain shape and orientation, can be obtained by 3D-EBSD and 3D-EBSD techniques.   152   APPENDICES   153   APPENDIX A MATLAB Code for Slip System Trace Analysis in Commercial Purity Ti clc; clear; % basal ssa(:,:,1) = [0 0 0 1; 2 -1 -1 0]; ssa(:,:,2) = [0 0 0 1; -1 2 -1 0]; ssa(:,:,3) = [0 0 0 1; -1 -1 2 0]; ibas = 1; fbas = 3; % prism ssa(:,:,4) = [0 1 -1 0; 2 -1 -1 0]; ssa(:,:,5) = [1 0 -1 0; -1 2 -1 0]; ssa(:,:,6) = [-1 1 0 0; -1 -1 2 0]; iprs = 4; fprs = 6; % pyramidal ssa(:,:,7) = [0 1 -1 1; 2 -1 -1 0]; ssa(:,:,8) = [1 0 -1 1; -1 2 -1 0]; ssa(:,:,9) = [-1 1 0 1; -1 -1 2 0]; ssa(:,:,10) = [0 -1 1 1; 2 -1 -1 0]; ssa(:,:,11) = [-1 0 1 1; -1 2 -1 0]; ssa(:,:,12) = [1 -1 0 1; -1 -1 2 0]; ipyra = 7; fpyra = 12; % pyramidal ssa(:,:,13) = [1 0 -1 1; 2 -1 -1 -3]; ssa(:,:,14) = [1 0 -1 1; 1 1 -2 -3]; ssa(:,:,15) = [0 1 -1 1; 1 1 -2 -3]; ssa(:,:,16) = [0 1 -1 1; -1 2 -1 -3]; ssa(:,:,17) = [-1 1 0 1; -1 2 -1 -3]; ssa(:,:,18) = [-1 1 0 1; -2 1 1 -3]; ssa(:,:,19) = [-1 0 1 1; -2 1 1 -3]; ssa(:,:,20) = [-1 0 1 1; -1 -1 2 -3]; ssa(:,:,21) = [0 -1 1 1; -1 -1 2 -3]; ssa(:,:,22) = [0 -1 1 1; 1 -2 1 -3]; ssa(:,:,23) = [1 -1 0 1; 1 -2 1 -3]; ssa(:,:,24) = [1 -1 0 1; 2 -1 -1 -3]; euler = [0 0 0]; sliptrace = [0,0,0]; phi1 = euler(1,1);   % Euler = euler-[90,0,0] input crystal orientation 154   PHI = euler(1,2); phi2 = euler(1,3); g11 = cosd(phi1)*cosd(phi2)-sind(phi1)*sind(phi2)*cosd(PHI); g12 = sind(phi1)*cosd(phi2)+cosd(phi1)*sind(phi2)*cosd(PHI); g13 = sind(phi2)*sind(PHI); g21 = -cosd(phi1)*sind(phi2)-sind(phi1)*cosd(phi2)*cosd(PHI); g22 = -sind(phi1)*sind(phi2)+cosd(phi1)*cosd(phi2)*cosd(PHI); g23 = cosd(phi2)*sind(PHI); g31 = sind(phi1)*sind(PHI); g32 = -cosd(phi1)*sind(PHI); g33 = cosd(PHI); g = [g11,g12,g13;g21,g22,g23;g31,g32,g33]; sigma = [0,0,0;0,1,0;0,0,0]; c_a=1.59; % rotation matrix % input global stress state for i = 1:1:24 % convert n & m to unit vector n = [ssa(1,1,i) (ssa(1,2,i)*2+ssa(1,1,i))/3^.5 ssa(1,4,i)/c_a]; m = [ssa(2,1,i)*1.5 3^.5/2*(ssa(2,2,i)*2+ssa(2,1,i)) ssa(2,4,i)*c_a]; ss(1,:,i) = n/norm(n); % alpha plane ss(2,:,i) = m/norm(m); % alpha direction end for k = 1:1:24 burger(1,:,k) = ss(2,:,k); rot_burger(:,1,k) = g'*burger(1,:,k)'; z = [0 0 1]; burger_z(k) = rot_burger(:,1,k)'*z'; end % rotate slip direction for j = 1:1:24; plane = ss(1,:,j); plane = plane'; rot_plane = g'*plane; z = [0,0,1]; sliptrace(:,:,j) = cross(rot_plane',z); x = sliptrace(1,1,j); y = sliptrace(1,2,j); X = [0,x]; Y = [0,y]; % rotate slip plane normal axis([-1 1 -1 1]); if j < 4 k = j; elseif j <= 6   155   k = j-3; elseif j <= 9; k = j-6; elseif j <= 12; k = j-9; elseif j <= 15; k = j-12; elseif j <= 18; k = j-15; elseif j<21; k = j-18; else k = j-21; end axis square axis([-1 1 -1 1]) plot(X,Y); text(x-0.05*k,y,num2str(j)); hold on end; % plot traces axis square axis([-1 1 -1 1]); SF = zeros(24,1); sigma = g*sigma*g'; for i = 1:1:24 a = ss(1,:,i); b = ss(2,:,i); N = g'*a'; M = g'*b'; SF(i) = N(2)*M(2); schmidfactor(i) = a*sigma*b'; end;   % Shimid factor calculation % alpha plane % alpha direction 156   APPENDIX B Procedure File for Simulating 3D Meshes for CPFE Modeling Shown in Figure 3.11(c) *new_model yes *select_reset *plot_reset *expand_reset *move_reset 365 237 0 371 220 0 430 225 0 40 130 0 138 179 0 272 143 0 296 138 0 354 168 0 370 144 0 430 155 0 215 104 0 267 126 0 296 110 0 308 87 0 359 108 0 401 74 0 430 110 0 40 40 0 220 40 0 305 40 0 410 40 0 430 40 0 0 530 0 40 530 0 430 530 0 470 530 0 0 0 0 40 0 0 430 0 0 470 0 0 *stop_procedure *fill_view *set_curve_type line *add_curves 1 2 1 6 6 7 2 7 2 3 3 8 |++++++++++++++++++++++++++++++ | points used to build the sample geometry |++++++++++++++++++++++++++++++ *add_points 40 490 0 120 490 0 230 490 0 380 490 0 430 490 0 40 460 0 108 470 0 219 464 0 127 459 0 174 441 0 225 423 0 378 407 0 40 405 0 103 389 0 192 407 0 301 383 0 361 389 0 430 395 0 132 348 0 168 353 0 178 324 0 278 330 0 284 283 0 389 296 0 430 300 0 40 290 0 121 301 0 139 296 0 157 296 0 231 238 0   157   8 9 7 8 11 16 12 4 3 4 5 12 6 13 9 14 19 15 10 11 20 21 16 17 24 23 22 24 18 13 26 19 27 28 21 29 23 31 32 37 36 30 32 25 26 34   10 10 9 11 16 17 17 12 4 5 18 18 13 14 14 19 20 20 15 15 21 22 22 24 31 31 23 25 25 26 27 27 28 29 29 30 30 32 38 38 37 36 33 33 34 35 28 35 35 41 41 42 36 42 42 43 37 43 38 39 39 40 33 40 34 48 48 49 41 49 49 50 44 50 43 44 44 45 39 45 45 46 40 47 46 47 46 51 50 51 47 52 51 52 53 54 54 55 55 56 1 54 5 55 53 57 56 60 48 58 52 59 57 58 58 59 59 60 *stop_procedure *set_curve_tolerance_absolute *redraw *remove_points all_existing *sweep_all *remove_unused_nodes *remove_unused_points *renumber_all |++++++++++++++++++++++++++++++ | set divisions for each curve (line) 158   |++++++++++++++++++++++++++++++ *set_curve_div_type_fix_ndiv *set_curve_div_num 8 *apply_curve_divisions 1 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 6 *apply_curve_divisions 12 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 3 *apply_curve_divisions 13 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 3 *apply_curve_divisions 2 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 8 *apply_curve_divisions 14 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 7 *apply_curve_divisions 3 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 15 *apply_curve_divisions 15 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 3 *apply_curve_divisions 4 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 5 *apply_curve_divisions 16 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 11 *apply_curve_divisions 5 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 10 *apply_curve_divisions 17 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 3 *apply_curve_divisions 6 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 5 *apply_curve_divisions 18 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 7 *apply_curve_divisions 7 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 5 *apply_curve_divisions 19 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 5 *apply_curve_divisions 8 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 7 *apply_curve_divisions 20 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 2 *apply_curve_divisions 9 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 8 *apply_curve_divisions 21 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 10 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 5 *apply_curve_divisions 22 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 8 *apply_curve_divisions 11 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4   159   *apply_curve_divisions 23 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 10 *apply_curve_divisions 35 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 6 *apply_curve_divisions 24 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 11 *apply_curve_divisions 36 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 25 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 8 *apply_curve_divisions 37 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 26 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 5 *apply_curve_divisions 38 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 3 *apply_curve_divisions 27 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 2 *apply_curve_divisions 39 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 10 *apply_curve_divisions 28 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 2 *apply_curve_divisions 40 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 6 *apply_curve_divisions 29 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 3 *apply_curve_divisions 41 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 10 *apply_curve_divisions 30 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 10 *apply_curve_divisions 42 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 6 *apply_curve_divisions 31 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 7 *apply_curve_divisions 43 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 8 *apply_curve_divisions 32 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 2 *apply_curve_divisions 44 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 5 *apply_curve_divisions 33 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 6 *apply_curve_divisions 45 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 34 # | End of List   *set_curve_div_type_fix_ndiv *set_curve_div_num 7 160   *apply_curve_divisions 46 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 3 *apply_curve_divisions 58 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 2 *apply_curve_divisions 47 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 3 *apply_curve_divisions 59 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 10 *apply_curve_divisions 48 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 6 *apply_curve_divisions 60 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 6 *apply_curve_divisions 49 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 7 *apply_curve_divisions 61 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 8 *apply_curve_divisions 50 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 9 *apply_curve_divisions 62 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 16 *apply_curve_divisions 51# | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 18 *apply_curve_divisions 63 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 11 *apply_curve_divisions 52 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 6 *apply_curve_divisions 64 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 12 *apply_curve_divisions 53 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 9 *apply_curve_divisions 65 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 10 *apply_curve_divisions 54 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 66 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 6 *apply_curve_divisions 55 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 2 *apply_curve_divisions 67 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 2 *apply_curve_divisions 56 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 6 *apply_curve_divisions 68 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 3 *apply_curve_divisions 57 # | End of List   *set_curve_div_type_fix_ndiv *set_curve_div_num 4 161   *apply_curve_divisions 69 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 81 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 6 *apply_curve_divisions 70 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 53 *apply_curve_divisions 82 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 71 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 53 *apply_curve_divisions 83 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 72 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 84 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 3 *apply_curve_divisions 73 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 85 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 11 *apply_curve_divisions 74 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 86 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 7 *apply_curve_divisions 75 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 39 *apply_curve_divisions 87 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 2 *apply_curve_divisions 76 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 88 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 77 # | End of List |++++++++++++++++++++++++++++++ | partition elements |++++++++++++++++++++++++++++++ *af_planar_quadmesh 1 2 3 4 # | End of List *select_elements all_existing *store_elements G1 all_selected *identify_sets *regen *select_elements all_existing *set_curve_div_type_fix_ndiv *set_curve_div_num 39 *apply_curve_divisions 78 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 79 # | End of List *set_curve_div_type_fix_ndiv *set_curve_div_num 4 *apply_curve_divisions 80 # | End of List   *af_planar_quadmesh 4 5 6 7 8 9 # | End of List *select_elements all_existing *store_elements G2 all_selected 162   *select_clear *select_elements all_existing *af_planar_quadmesh 12 30 31 32 33 29 # | End of List *select_elements all_existing *store_elements G9 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 6 15 14 13 12 11 10 # | End of List *select_elements all_existing *store_elements G3 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 13 18 35 34 30 # | End of List *select_elements all_existing *store_elements G10 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 14 16 17 18 # | End of List *select_elements all_existing *store_elements G4 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 36 20 22 38 37 # | End of List *select_elements all_existing *store_elements G11 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 3 9 21 20 19 # | End of List *select_elements all_existing *store_elements G5 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 23 27 41 40 39 38 # | End of List *select_elements all_existing *store_elements G12 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 8 25 24 23 22 21 # | End of List *select_elements all_existing *store_elements G6 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 28 33 43 42 41 # | End of List *select_elements all_existing *store_elements G13 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 7 10 26 25 # | End of List *select_elements all_existing *store_elements G7 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 43 32 44 45 46 47 48 # | End of List *select_elements all_existing *store_elements G14all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 11 29 28 27 24 26 # | End of List *select_elements all_existing *store_elements G8 all_selected *select_clear *select_elements all_existing   *af_planar_quadmesh 31 34 50 49 44 31 # | End of List *select_elements all_existing *store_elements G15 all_selected 163   *select_clear *select_elements all_existing *af_planar_quadmesh 58 46 59 69 68 67 # | End of List *select_elements all_existing *store_elements G22 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 51 37 39 53 52 # | End of List *select_elements all_existing *store_elements G16 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 60 71 72 70 69 # | End of List *select_elements all_existing *store_elements G23 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 53 40 42 48 56 55 54 # | End of List *select_elements all_existing *store_elements G17 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 66 68 70 73 74 # | End of List *select_elements all_existing *store_elements G24 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 45 49 61 60 59 # | End of List *select_elements all_existing *store_elements G18 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 73 72 75 76 # | End of List *select_elements all_existing *store_elements G25 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 62 52 54 64 63 # | End of List *select_elements all_existing *store_elements G19 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 1 5 15 16 81 78 80 # | End of List *af_planar_quadmesh 80 2 19 36 51 62 84 86 82 77 # | End of List *af_planar_quadmesh 81 79 83 88 85 75 71 61 50 35 17 # | End of List *af_planar_quadmesh 84 63 65 74 76 85 87 # | End of List *select_mode_invert *select_elements all_existing *store_elements rim all_selected *select_clear *identify_sets *regen *af_planar_quadmesh 64 55 57 67 66 65 # | End of List *select_elements all_existing *store_elements G20 all_selected *select_clear *select_elements all_existing *af_planar_quadmesh 56 47 58 57 # | End of List *select_elements all_existing *store_elements G21 all_selected *select_clear *select_elements all_existing   *sweep_all *remove_unused_nodes *remove_unused_points *renumber_all *check_distorted *expand_reset 164   *set_expand_translation z 10 *set_expand_repetitions 10 *expand_elements all_existing *select_reset *select_method_box *select_elements *store_nodes left_nodes all_selected *select_reset *select_clear *select_method_box *select_nodes -1 550 -1 550 -1 1 *store_nodes back_nodes all_selected *select_clear *select_method_box *select_faces 469.5 470.5 -1 530 -1 110 *store_faces right_faces all_selected *select_clear *select_method_box *select_nodes -1 1 -1 11 -1 21 *store_nodes fixed_nodes all_selected 0 470 |x region 0 530 |y region 50 100 |z region *store_elements backside all_selected *stop_procedure *set_nodes off *elements_solid *redraw *stop_procedure |++++++++++++++++++++++++++++++ | parameters definition for loading |++++++++++++++++++++++++++++++ *define max_disp .25 *define max_time 150 *define n_steps 150 *define El_time 20 *define Pl_time 80 *define P2_time 150 *define El_load -250 *define Pl_load 300 *define P2_load -330 |++++++++++++++++++++++++++++++ | boundary conditions |++++++++++++++++++++++++++++++ *apply_type face_load *apply_dof p *apply_dof_value p *apply_dof_table p load_time *select_clear *select_faces *select_sets right_faces *add_apply_faces all_selected *stop_procedure *new_md_table 1 1 *set_md_table_type 1 time *table_name load_time *table_add 0 0 El_time El_load |elastic loading, a point (x,y) in time-load figure Pl_time Pl_load |plastic loading P2_time P2_load *new_apply *apply_type fixed_displacement *apply_dof x *apply_dof_value x *add_apply_nodes left_nodes # |end list *new_apply *apply_type fixed_displacement *apply_dof z *apply_dof_value z *add_apply_nodes back_nodes # |end list *new_apply *apply_dof x *apply_dof_value x *apply_dof y *apply_dof_value y *apply_dof z *apply_dof_value z *add_apply_nodes fixed_nodes # |end list *table_fit *stop_procedure *select_reset | define lower nodes, upper faces, back nodes and fixed nodes *select_clear *select_method_box *select_nodes -.5 .5 -1 550 |1