You are on page 1of 375

EXPERIMENTAL INVESTIGATION OF HIGH VELOCITY IMPACTS ON BRITTLE MATERIALS

by DAVID ISAAC NATHENSON

Submitted in partial fulfillment of the requirements For the degree of Doctor of Philosophy

Thesis Advisor: Dr. Vikas Prakash

Department of Mechanical & Aerospace Engineering CASE WESTERN RESERVE UNIVERSITY

May, 2006

CASE WESTERN RESERVE UNIVERSITY SCHOOL OF GRADUATE STUDIES

We hereby approve the dissertation of

David Isaac Nathenson ______________________________________________________


candidate for the Ph.D. degree *.

(signed)_______________________________________________ (chair of the committee)

Vikas Prakash

________________________________________________

John Lewandowski

________________________________________________

Clare Rimnac

________________________________________________

Joseph M.Prahl

________________________________________________

________________________________________________

(date) _______________________

*We also certify that written approval has been obtained for any proprietary material contained therein.

Copyright 2006 by David Isaac Nathenson All rights reserved

Dedicated to my wife, Randi Gross Nathenson, For her constant Love and support.

1
Table of Contents Title Page Signature Sheet Copyright page Dedication Page Table of Contents List of Tables List of Figures Acknowledgements Abstract Chapter I Introduction A. Properties of Brittle Materials A.1. Soda Lime Glass A.2. Silicon Nitride B. Intent of This Dissertation References Chapter II Configuration and Procedures for Particle Impact Experiments A. Components A.1. Projectile Firing Mechanism A.2. Gun Barrel and Sabot Stripper A.3. Impact Chamber and Target Holder B. Measuring Systems B.1. High Speed Camera B.2. Strain Gages B.3. Laser Velocity System C. Experimental Procedures C.1. Series Preparation C.1.1. Soda Lime Glass Specimen Preparation C.1.2. Sabot/Projectile Preparation C.2. Setup Sequence C.2.1. Projectile Accelerator Alignment C.2.2. Measurement System Preparation C.3. Pre-Firing Sequence D. Summary References Figures Chapter III Particle Impact Experiments on Soda Lime Glass A. Background on Particle Impact of Brittle Materials A.1. Theoretical Calculations A.1.1. Hertzian Theory A.1.2. Computed Longitudinal, Shear and Rayleigh Wave Profiles A.1.2.1. Lambs Solution A.1.2.2. Mitras Solution A.1.3. Estimation of the Energy Dissipated by Elastic Waves A.2. Previous Research on the Cracking Patterns of Soda Lime Glass A.2.1. Configurations and Computations for Previous Experiments A.2.2. Results of Previous Experiments B. Design of Soda Lime Glass Particle Experiments C. Results and Analysis of the Soda Lime Glass Experiments C.1. Projectile Velocity vs. Firing Chamber Pressure Calibration C.2. Time of Contact Between the Sphere and the Plate C.3. Examination of the Impact Surface Strain Gage Records C.3.1. Strain Wave Components Explained Using Theory and Numerical Methods i ii iii iv 1 4 5 7 8 9 10 11 18 22 25 28 29 29 31 32 34 34 35 36 37 37 37 38 39 39 39 41 42 43 43 48 48 49 50 52 53 55 58 61 61 64 69 71 71 73 74 74

2
C.3.2. Description of Experimental Strain Records C.3.3. Correlation Between Rear Surface Vibration and Crack Oscillation C.4. Impact Energies and Coefficients of Restitution C.4.1. Determination of Experimental Impact Energy and Coefficient of Restitution 4.2. Elastic Stress Wave and Vibrations as Portions of the Impact Energy C.5. Cracking Pattern Behavior C.5.1. Details of Cracking Following Experiments C.5.2. Dynamic Evolution of Cracking D. Significance of Results References Tables Figures Chapter IV Planar Impact Experiments: Configuration and Procedures A. The Single Stage Gas Gun and Observation Systems A.1. Description of the Single Stage Gas Gun A.2. Laser Velocity, VISAR and Tilt Pin Measurement Systems A.2.1. Laser Velocity System A.2.2. Tilt Measurement and Triggering System A.2.3. VISAR Interferometer B. Experimental Procedures B.1. Specimen and Flyer Preparation B.1.1. Materials for Specimens and Flyers B.1.2. Assembly of the Projectiles and Specimens B.2. Gas Gun Setup Sequence B.3. Firing Sequence C. Summary References Figures 77 79 81 81 84 85 86 91 95 99 101 106 124 124 125 127 127 128 129 132 132 132 133 135 137 140 141 142

Chapter V The Shock Response of AS800 Grade Silicon Nitride 147 A. Background on Shock Compression Experimentation 148 A.1. The Theory of Planar Shock Compression 148 A.1.1. Elastic Impact Theory 148 A.1.2. Elastic-Plastic Impact Theory and Experiments 151 A.2. The Failure in Dynamic Tension due to the Spallation of Silicon Nitride 155 A.2.1. The Theory of Material Spall 155 A.2.2. Experimental Determination of Spall Strength 159 B. Experimental Design 162 B.1. Description of Materials Under Investigation 162 B.2. Experimental Matrix for the Shock Compression Tests 163 B.3. Experimental Matrix for the Pressure-shear Impact Tests 164 C. Experimental Analysis and Results 168 C.1. Elastic and Hydrodynamic Relationships in the Hugoniot State 168 C.1.1. Shock Velocity vs. Particle Velocity in the Hugoniot State 168 C.1.2.Difference in Hugoniot State Variables due to Pressure-Shear Loading 170 C.1.3. Elastic Hugoniot Stress and Strain Relationship 170 C.1.4. Calculation of the HEL 171 C.1.5. Determination of the Hydrodynamic Hugoniot Stress and Strain States 173 C.2. Measurement of the Spall Strength 174 C.2.1. Derivation of the Spall Strength Formulae 176 C.2.2. Spall Strength and Material Failure Mode during Spall in Pure Shock Compression 179 C.2.3. Spall Strength and Failure Modes Under Combined Pressure and Shear Impact Loading 182 C.3. Material Failure Modes during Dynamic Spall and Fractography of the Spall Surface 183

3
D. Summary of Planar Results References Tables Figures 186 190 191 193

Chapter VI The Effect of Shock Re-Shock, Shock Unloading, and Stress Reverberations on the Behavior of Silicon Nitride 206 A. Theory and Computations for Multiple Shock Experiments 208 A.1. Shock Re-shock and Shock Unload Experiments 208 A.2. Reverberation Experiments 213 B. Design of Shock Re-shock, Shock Unload, and Reverberation Experiments 221 C. Results and Analyses of Multiple Shock Experiments 223 C.1. Shock-Re-shock and Shock-Unload Experiments 224 C.1.1. Computations for Shock Re-Shock and Shock Unload Experiments 223 C.1.2. Observations and Analysis of Shock Re-shock and Shock Unload Experiments 228 C.2. Reverberation Experiments 230 C.2.1. Computations for Shock Reverberation Experiments 230 C.2.2. Observations and Analysis of the Shock Reverberation Experiments 234 D. Conclusions and Discussion 236 References 238 Tables 239 Figures 240 Chapter VII Investigation of Shock Induced Failure Wave Propagation in Soda Lime Glass 248 A. Description of the Nature of the Failure Wave 249 A.1. Experimental Determination of the Properties and Effects of the Failure Front 249 A.2. Proposed Mechanisms for the Propagation of the Failure Front 254 B. Design of Experiments to Investigate the Failure Wave in Soda Lime Glass 261 B.1. Description of the Experimental Configurations 261 B.2. Modifications to the VISAR System to Enabled the Simultaneous Measurement of the Normal and Transverse Components of the Particle Velocity Histories 263 C. Experimental Results and Analysis of Failure Waves in Soda Lime Glass 265 C.1. Spall Strength of Glass under Shock Compression and Pressure-Shear Loading 265 C.1.1. Normal Particle Velocity Calculations 265 C.1.2. Transverse Particle Velocity Calculations 268 C.1.3. Spall Strength and Particle Velocity Observations for Aluminum Impacting Glass 270 C.2. The Effects of the Failure Wave on the Properties of Soda Lime Glass 272 C.2.1. Computations for the Elastic Predictions 273 C.2.2. Experimental Observations of Spall Strength and Transverse Particle Velocity 274 C.3. Observation of the Impedance in the Comminuted Material 278 C.3.1. Elastic Computations for the Internal Stresses and the Acoustic Impedance in Experiments G/WC1 and G/WC2 279 C.3.2. Measurements of the Normal and Transverse Particle Velocities and the Impedance Change Due to the Failure Wave 280 D. Conclusions and Discussion 283 References 287 Tables 288 Figures 289 Chapter VIII Summary of Experimental Analyses 303 A. Particle Impact Experiments on Soda Lime Glass 304 B. Planar Shock Compression and Pressure-Shear Experiments on AS800 Grade Silicon Nitride 308 C. The Effect of Dual Shock Loading and Shock Reverberation on the Strength of AS800 Grade Silicon Nitride 311

4
D. Shock Induced Failure Waves During Dynamic Compression of Soda Lime Glass E. Summary References 312 316 318

Appendices 321 Appendix 1: Equations of Wave Propagation Resulting from a Point Load on a Elastic Half Space 321 1.1. Using the Equations of Motion to Generate the Radial and Vertical Surface Displacements 321 1.2. Calculating the Radial and Vertical Stresses on the Surface 325 1.3. Specific Solution for a Concentrated Vertical Pressure 329 1.4. Generalizing the Equations of Displacement for an Arbitrary Loading 338 Appendix 2: Calculations for the Surface Wave Produced by the Loading of a Circular Area of a Half Space by an Impulse Load. 340 2.1. Solutions to the Equations of Motion Transformed using Laplace Transforms 340 2.2. Cagniards method for obtaining the inverse Laplace transforms of, I1, I2 346 2.3. Surface Displacements 351 Appendix 3: Calculation of Energy Absorbed by Elastic Waves During Impact References Figures 358 368 369

5 List of Tables
Table 3.1: Experiment List Including Specimen Impact Velocity and Thickness. Table 3.2: Experimental, Theoretical, and Numeric Peak Contact Times. Table 3.3: Experimental Strain Measurements. Table 3.4: Impact Kinetics and Coefficients of Restitution. Table 3.5: Post Impact Static Cracking Pattern Details Table 5.1: Experimental Results for Hugoniot variables Table 5.2: Mohr's Circle Stress Calculations Table 5.3: Experimental Results for Spall Stress and Recovery Time Table 6.1: Dynamic Properties for Experiments SR-1 and SU-1 Table 6.2: Dynamic Properties for Experiments RB-1 and RB-2 Table 7.1: Parameters for spall strength and tungsten carbide on glass experiments Table 7.2: Parameters for glass on tungsten carbide experiments 101 102 103 104 105 192 192 193 239 239 288 288

List of Figures
Figure 2.1: Complete view of projectile accelerator Figure 2.2: View of firing chamber with piston and sabot loading procedure Figure 2.3: Sabot and projectile assembly diagram Figure 2.4: Sabot stripper diagram Figure 2.5: Impact chamber diagram and Measurement Setup Figure 2.6: Target holder diagram Figure 2.7: CEA-06-032WT-120 strain gage image Figure 3.1: Hertzian approximation and Numerical Simulations for Impact Force Curve Figure 3.2: Surface strain profile using the equations from Lamb Figure 3.3: Surface strain profile using the equations from Mitra Figure 3.4: Calibration curve for projectile impact including all experiments Figure 3.5: Plot of experimental, theoretical and numerical contact time vs. velocity Figure 3.6: LS-DYNA surface strains profiles on a half space Figure 3.7: LS-DYNA surface strain profiles on plates with infinite lateral boundaries Figure 3.8: LS-DYNA surface strain profiles on plates with finite rectangular geometries Figure 3.9: Comparison plot of strain profiles in 350 m/s range for all four thicknesses. Figure 3.10: Variation of strain with specimen thickness and velocity Figure 3.11: Rear surface strains from a bending series impact of a 5 mm specimen Figure 3.12: Rear surface strains from a bending series impact of a 15 mm specimen Figure 3.13 Example of oscilloscope output from laser velocity triggering system Figure 3.14: Variation in rebound kinetic energy to impact kinetic energy with velocity Figure 3.15: Variation in coefficient of restitution with velocity Figure 3.16: Impact of 5 mm thick specimen at 323 m/s (BQ-13) showing conical cracks Figure 3.17: Impact of 3 mm thick specimen at 345 m/s (BQ-26) showing radial cracks Figure 3.18: Impact of 15 mm thick specimen at 371 m/s (BQ-25) showing lateral cracks Figure 4.1 Overview of the single stage gas gun Figure 4.2: Firing chamber and air supply diagram Figure 4.3: Valyn VISAR interferometer design Figure 4.4: Fringe patterns and velocity curve for representative experiment Figure 4.5: Representative velocity measurement system beam intensity plot Figure 4.6: Specimen and flyer ring configurations for pressure-shear experiments. Figure 5.1: Calculation plots for elastic compression impact experiments. Figure 5.2: Calculation plots for elastic-plastic compression impact experiments Figure 5.3: Normal experimental free surface velocity history plots 43 43 44 44 45 46 47 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 142 143 144 145 145 146 193 193 194

6
Figure 5.4: Shear experiments free surface velocity history plots Figure 5.5: Calculation plots for pressure-shear impact experiments. Figure 5.6 Shock velocity plotted versus Hugoniot state particle velocity Figure 5.7: Hugoniot elastic and hydrodynamic stresses and strains Figure 5.8: Rear surface velocity history plot showing Hugoniot Elastic Limit. Figure 5.9: Spall strength plotted versus free surface velocity for normal impact Figure 5.10: Spall strength plotted versus free surface velocity for pressure-shear Figure 5.11: Post impact fragment images Figure 5.12: SEM images of SC-8, 31.8 m/s Hugoniot particle velocity Figure 5.13: SEM images of SC-2, 100 m/s Hugoniot particle velocity Figure 5.14: SEM images of SC-4, 178 m/s Hugoniot particle velocity Figure 5.15: SEM images of SC-12, 152 m/s Figure 5.16: SEM images of Polished surface of specimen, pre-impact Figure 6.1: Theoretical stress vs. strain diagram Figure 6.2: Reverberation experiment design Figure 6.3: Time vs. distance and stress vs. velocity diagrams experiment RB-2 Figure 6.4: Design for shock re-shock and shock release experiments Figure 6.5: Time vs. distance and stress vs. velocity diagrams experiment RB-1 Figure 6.6: Time vs. distance and stress vs. velocity diagrams experiment SR-1 Figure 6.7: Time vs. distance and stress vs. velocity diagrams experiment SU-1 Figure 6.8: Velocity vs. time profiles for experiments SR-1 and SU-1. Figure 6.9: The stress vs. strain plot for reloading and unloading experiments Figure 6.10: Velocity vs. time profiles experiments RB-1 and RB-2. Figure 6.11: Reverberation experiments stress and strain levels Figure 7.1: Specimen configuration for normal spall strength experiment. Figure 7.2: Specimen configuration for pressure-shear spall strength experiment. Figure 7.3: Time vs. distance and stress vs. velocity diagrams for spall experiments Figure 7.4: Specimen configuration for WC/Glass experiments Figure 7.5: Time vs. distance and stress vs. velocity diagrams exp. WC/G1, WC/G3 Figure 7.6: Time vs. distance and stress vs. velocity diagrams exp. WC/G2, WC/G4 Figure 7.7: Specimen Configuration for Glass/WC Experiments Figure 7.8: Time vs. distance and stress vs. velocity diagrams exp. G/WC1, G/WC Figure 7.9: Experimental measurement technique Figure 7.10: Velocity Profile for Spall experiment Al/G1 Figure 7.11: Velocity Profile for Spall experiment Al/G2 at 18 degrees Figure 7.12: Velocity Profile for Experiment WC/G1 Figure 7.13: Velocity Profile for Experiment WC/G2 Figure 7.14: Velocity Profile for Experiment WC/G3 Figure 7.15: Velocity Profile for Experiment WC/G4 Figure 7.16: Velocity Profile for Experiment G/WC1 Figure 7.17: Velocity Profile for Experiment G/WC2 Figure 7.18: Transverse velocity measurements Figure A.1: Surface displacement as described by Lambs equations Figure A.2: Schematic solutions of entire wave based on Lambs equations Figure A.3: Mitras surface displacement in the radial direction 195 195 196 197 198 199 200 201 202 203 203 204 205 240 241 241 242 242 243 243 244 246 246 247 289 289 290 290 291 291 292 292 293 294 295 296 297 298 399 300 301 302 369 370 371

7
Acknowledgments

I would like to acknowledge the assistance of Processor Vikas Prakash and J Michael Pereira from NASA without whom this work would not have been possible. I would also like to thank both the NASA Glenn Research Center and Case Western Reserve University for financial support. Additionally, I wish to thank the members of the D. K . Wright lab for their assistance and support: Mostafa Shazly, Liren Tsai, Guodong Chen, George Sunny, Fu Ping Yuan, Naoto Utsumi, and Tang Xin.

Experimental Investigation of High Velocity Impacts on Brittle Materials

Abstract by

DAVID ISAAC NATHENSON

Experiments were conducted on soda lime glass and AS800 grade silicon nitride. Soda lime glass is often used in windows of military vehicles and aircraft where integrity in the event of shrapnel impacts is of vital concern. AS800 grade silicon nitride is considered one of the leading material candidates for the next generation of aircraft engine turbine blades because of its superior high temperature properties when compared with nickel based super-alloys. The suitability of these materials for their applications depends upon their response to point and planar dynamic impact loading. An experimental apparatus was constructed to fire one-sixteenth inch diameter hardened chrome steel ball bearings at 50 mm square soda lime glass blocks of thicknesses between 3 mm and 25.4 mm. Inelasticity due to the crushed zone effects the coefficients of restitution and the surface strains. The change in severity of cracking with velocity and specimen thickness is observed. Shock compression and pressure-shear experiments were conducted by means of a single stage gas gun capable of attaining impact velocities of 600 m/s. High velocity planar shock compression experiments on soda lime glass reveal a lack of spall strength, and a decrease of shear impedance and shear strength in the presence of a failure wave. The longitudinal impedance remains nearly constant. The spall strength of glass is 3.49 GPa and is sensitive to the presence of shear. Shock compression studies on silicon nitride using normal shock compression show that the material has a Hugoniot Elastic Limit of 12 GPa and that the spall strength decreases with increasing impact velocity due to damage below the HEL. The presence of inelastic deformation stops this trend, while the presence of shear increases the rate of spall strength drop by five times because of more severe microscopic damage. Experiments involving multiple shocks on silicon nitride show that material loading and unloading follows the shock Hugoniot closely. The HEL of the shocked material is decreased by 6%, while the residual strength remains high. This indicates that the longitudinal properties of silicon nitride including its shock impedance are not dramatically affected during the initial shock compression.

Chapter I: Introduction
Brittle materials have many uses in aerospace and military applications. Glass is one brittle material that is utilized in vision elements of both aircraft and combat vehicles. In these situations glass is subjected to impact conditions from objects such as projectiles or runway debris. Due to the lack of plastic energy dissipation, glass fails primarily by brittle cracking. Ceramics such as AS800 grade silicon nitride are candidates for the next generation of aircraft engine turbine blades and casings because they have higher strength properties at elevated temperatures than current nickel based super-alloys. However, ceramics are more brittle than these alloys. In order to determine the suitability of these brittle materials for applications where resistance to fracture is important, their response to dynamic impact loading must be studied. Both small particle impact and planar shock compression loading are among current testing methods. By impacting soda lime glass with spherical particles, the three dimensional wave and cracking behavior can be examined. Shock compression involves a state of plane strain in which the behavior of glass and silicon nitride can be quantitatively examined. Through these methods the dynamic response of the material can be discerned. This assists in the determination of their suitability for aerospace applications. In this chapter, the specific properties of the two materials are discussed. The specifics of each materials brittle behavior are explored. The cracking behavior in soda lime glass under particle impact and the failure wave propagation under shock compression are summarized. The material properties and point impact classification of this specific silicon nitride are reviewed. From this, the reasons for this study emerge. The specifics of the study and its goals are then explained. Finally, the point impacts on soda lime glass, and the various shock compression experiments that are performed in this study are categorized.

10

A. Properties of Brittle Materials Brittle materials have many qualities that recommend them for aerospace and military applications. Ceramics for example when compared with ductile materials retain more strength at elevated temperatures. Typically, their melting temperatures are higher [ASM, 1999]. They also have less creep than ductile materials. Glasses on the other hand are commonly used in windows because of their clarity. They also have high strength and high shock compressive yield limits [Rosenberg et al., 1985; Bourne et al., 1996]. These properties recommend ceramics and glasses for their respective applications. However, brittle materials have a distinct downside in their lack of ductility. In broad terms, brittle cracking and fracture occurs when irreversible damage occurs before the yield stress is reached, while in a ductile failure, the material will yield before failure. The brittle fracture leaves a glittering, crystalline appearance, or one covered with river lines caused by tearing between slip planes. High strength materials, such as ceramics, like silicon carbide and high strength steels such as 18 Ni maraging steel are examples of brittle materials [Dowling, 1999]. One method of quantifying brittle versus ductile materials is by examining the plastic strain to failure. Ceramics and glasses have almost no plasticity before failure. As a result, brittle fracture occurs at all normal temperatures and strain rates in these materials. For some of these materials such as ferritic steels raising the temperature above a critical ductile to brittle transition temperature (DBTT) causes a substantial increase in ductility [Anderson, 1995]. Ductile failure can show both global and local necking. Micro-voids typically expand and coalesce causing the failure. This results in a dimpled appearance either locally or globally. Examples of ductile materials include lower strength steels such as AISI 1020 and aluminum 7075-T6 [Dowling, 1999]. Brittle fractures are generally less desirable in applications subjected to large static or dynamic stresses occur due to the lack of warning before failure and the catastrophic nature of the failure. However, ductile materials usually have lower strength and higher creep rates, especially at elevated temperatures. Thus, brittle materials are preferred in applications where strength is the overriding concern. For this reason, any

11

attempt to characterize the acceptability of brittle materials for aerospace or military applications must include an assessment of the dynamic fracture/spall strength. In the following sections, the relevant behavior and properties are reviewed for both soda lime glass and silicon nitride. A.1. Soda Lime Glass Soda lime glass has many uses and similar properties to many glass ceramics. It is employed as the vision elements of military and aviation vehicles. Glass is also employed as a modeling material for glass ceramics that have similar brittle behavior. Because soda lime glass is brittle, it has practically no plastic deformation. Rather, stress waves and failure waves are responsible for most of the energy dissipation upon impact. These effects can be studied using particle impact and shock compression experimentation. Experimentally and theoretically, cracking and stress behavior in glasses have been explored. Under lower velocity dynamic loading a combination of stress and cracking systems dominates the material response. Two types of stress fields are applicable to small particle impacts. The first is Hertzian which occurs for a blunt tipped projectile and produces a cone crack that follows the path of the smallest principal stress [Evans and Wilshaw, 1976]. The other field is Boussinesq, which involves point impacts and generates stresses which cause half penny radial cracks on loading and upon unloading saucer shaped lateral cracks [Lawn and Fuller, 1975]. At higher impact velocities the stress waves have a more pronounced effect on the cracking patterns [Field, 1988; Field et al., 1989]. Using high speed photography and strain gages, these stress and cracking patterns are observed in particle indentation experiments. As described by Hertz and others, the crack formation in the case of a blunt spherical indenter follows a distinct path. The crack path is defined by principal stress trajectories. This path is, to a first approximation, orthogonal to the highest principal stress in the stress field prior to the cracking. The crack thus

12

propagates along the planes defined smaller principal stresses. This was determined by comparing experimental crack paths with contours of the principle stresses [Frank and Lawn, 1967]. Ideally, the crack tries to propagate in a straight line. However, oblique stresses are generated on the crack when the direction of the principal stresses changes. These stresses force the crack to correspond to the principal stress directions. [Frank and Lawn, 1967]. The cracking is divided into four parts: nucleation, formation, propagation, and unloading [Lawn et al., 1974]. The initial crack forms just outside the contact circle radius where the stresses are tensile. There may be more than one crack initiated. These cracks form at existing flaws at the surface, which extend as the stress level increases. After the cracks initiate, the dominant crack extends in a circular manner around the region of contact. Thus, a shallow surface crack is created. For any stable extension of the surface crack a critical level of energy must be overcome by increasing the contact force. Overcoming this energy results in stable downward motion and a ring-like crack [Lawn et al., 1974]. The crack continues downward to a critical depth of about 0.1 times the contact radius. Then, the crack begins to propagate in an unstable manner. This propagation occurs in the form of a Hertzian cone with a depth about equal to the contact radius. Sharp indenters, such as a Vikers pyramid or a cone have distinctly different stress fields and hence cracking patterns when compared with blunt indenters. Theoretical elastic stress intensity varies inversely with the square of the distance from the tip. This is commonly known as the Boussinesq field [Lawn and Fuller, 1975]. These sharp indenters are more often employed because the indentation size has no effect on the contact pressure [Lawn and Wilshaw, 1975]. The largest tensile stress is now immediately beneath the tip. Micro-cracking initiates there [Lawn and Fuller, 1975]. These micro-cracks are then extended to form macroscopic cracks in near-penny shapes. The cracks will grow during loading in the direction of the trajectories of the smallest principal stress as in

13

the blunt sphere case. The crack will propagate both downward, in a median pattern both downwards and outwards orthogonal to the hoop stress [Lawn and Swain, 1975]. These cracks, called radial cracks, because of their extension on radial (constant ) planes will grow in a stable fashion and full-penny geometry until they reach a depth of approximately one contact radius [Lawn and Fuller, 1975]. Following this, the cracks will break through compressive stress lobes in the hoop direction that were restricting the propagation. The restrictive compressive stress lobes occur at an angle greater than or equal to 51.8 in the second principle stress field [Lawn and Swain, 1975]. The median cracks will then reach the free surface and continue to propagate in a stable fashion. They then have half penny geometry [Lawn and Fuller, 1975]. Additionally, during unloading, the residual stress field causes the creation of cracks in the direction parallel to the free surface. These lateral cracks nucleate at the edge of the deformation zone close to the surface and propagate in a saucer shape. This can cause chipping of the surface [Lawn and Fuller, 1975]. This chipping is the removal of material from the surface because of the lateral crack extension [Evans and Wilshaw, 1976]. Thus two types of crack systems, median/radial and lateral, exist for brittle materials under the Boussinesq stress field. Under short duration and high amplitude loading inertia effects start to dominate the stress field patterns. For 3 mm and 5 mm diameter steel projectiles on soda lime glass this shift in stress patterns is observed to occur at around 250 m/s [Field, 1988; Field et al., 1989]. The internal fractures become very chaotic and the surface cracks are caused by Rayleigh surface waves. Back surface damage is caused by the reflection of the compressive loading wave as a tensile wave from the back surface [Field et al., 1989]. It is apparent that this is a different category of cracking than that obtained at lower impact speeds, where cracking patterns similar to those that exist in quasi-static indentation testing are observed [Field, 1988].

14

In glasses under planar shock compression, a failure wave has been observed by several investigators. This failure wave has been deduced from the gradual propagation of failed material resulting in a comminuted state [Kanel et al., 1992]. It has also been observed directly using high speed photography [Brar et al. 1992; Bourne et al., 1995]. Many researchers like Clifton et al. [1997] have investigated glass using shock compression experiments. Espinosa et al. [1997] and Feng [1999] have suggested models for failure wave. Failure waves have been identified in several brittle materials. Brar et al. [1992] noted a failure wave propagating in pyrex bars and soda lime glass plates under impact loading. Photographic records were recorded during the pyrex bar experiments showing the propagation of this wave. It is noted in these experiments that impacting with steel plates instead of pyrex bars resulted in a higher failure wave velocity; this is understood to be due to the larger impact stress that is generated with the steel bar at the same impact velocity [Brar et al., 1992]. Brar et al. [1992] and Kanel et al. [1992] suggest that the generation of a compression wave at the intersection of the release wave in shock compression experiments on soda lime glass means that the failed material has a slightly lower impedance than the intact material. Brar et al. [1992] suggest that this is due to the comminution of the glass, a fact which is confirmed by the lack of spall strength. The spall strength drops from 30 kbar before the failure wave to almost zero behind the failure wave. Also, transverse strain gages indicate a jump in the transverse stress level behind the failure wave at experiments with impact stresses greater than 38 kbar. This increased stress reflects a drop in the shear strength of the material behind the failure wave. The failure wave has been observed to travel at a speed of about 1.5 to 2.5 km/s [Ginzberg and Rosenberg, 1998] which is lower than the elastic wave speed of 5.7 km/s. The resulting material alteration causes the maximum shear stress the material can withstand and the spall strength to drop. However, the longitudinal

15

wave speed and the elastic impedance remain relatively unchanged [Ginzberg and Rosenberg, 1998]. This failure wave has been observed to initiate during shock compression at impact stresses between half of the Hugoniot Elastic Limit (HEL) of the material and the HEL [Ginzberg and Rosenberg, 1998]. The HEL of the material represents the uniaxial stress component in the impact direction during shock compression at which the inelastic deformation begins. This failure wave is reported in experiments throughout the literature, but the cause of its formation remains uncertain. The systematic study of this failure wave is necessary in order to determine its nature and causes. Additional images of the propagating failure wave were taken by Bourne et al. [1995]. These plate impact experiments employ copper flyer plates and soda lime and pyrex glass specimens. The wave front for soda lime glass consist of many bifurcating cracks. In pyrex, a greater number of small cracks are observed. In both glasses, crack nucleation is observed at a small number of sites just before the front. Crack propagation first links up the initiation sites and following coalescence the entire front propagates [Bourne et al., 1995]. Also in Bourne et al. [1995], the failure wave propagation velocities were observed to increase with impact velocity. For soda lime glass, the failure wave was observed to travel at 1.8 km/s for a 250 m/s (2.2 GPa) impact and to increase to 3.6 km/s for a 760 m/s (6.8 GPa) impact. The smaller failure wave velocity is from about half of the shear wave speed while the higher wave velocity is about the shear wave speed. This variation is not seen over the same range of impact velocities in the pyrex glass where for the same set of impact velocities the failure wave only increases from 3.4 km/s to 3.8 km/s, which is close to the shear wave velocity in the material. From the study it was concluded by Borne et al. [1995] that the failure wave is associated with the shock front shear stresses.

16

Several models exist describing the failure wave including that of Espinosa et al. [1997]. There the suggestion that this failure front in plates is due to inhomogeneous shear flow is discussed. The compressive shock wave causes this flow on fault planes. The observed micro-cracks then nucleate along shear plane intersections. This model suggests that the damage is a function of the stress level [Espinosa et al., 1997]. A comprehensive model for the failure wave is discussed by Feng [2000]. This model classifies the failure propagation as a diffusive phenomenon. The shock compression of the material causes heterogeneous microfissuring, shear dilatancy and void collapsing under a confining stress all at once. Thus the micro-scale damage is controlled by the heterogeneous deformation. This process causes the mean stress to increase and the devatoric stress to decrease. The failure wave results in the conversion of the devatoric strain energy into volumetric potential energy [Feng, 2000]. In this model the propagation of the failure wave occurs due to the progressive percolation of microfissures [Feng, 2000]. The failure process initiates at the impact surface. Flaw nucleation occurs rapidly followed by material comminution without substantial crack growth. The presence of damage at the surface causes stress concentrations within the material, which provide local initiation sites and the microfissures spread. The failure propagates laterally much faster than longitudinally. This spreading of cracks accounts for the diffusive nature of the model [Feng, 2000]. A comparison of Fengs [2000] model to experimental results indicate that the model predicts the longitudinal stress and particle velocity measurements. Additionally using the time between loading and the initiation of damage as an alternative to inferring the variations in the failure wave velocity by measuring the rise time of the waves across the strain gages is suggested. This suggestion provides better agreement with experimental results. Some questions are left

17

unanswered by the model, however such as inconsistencies in experimental longitudinal strains and stresses [Feng, 2000]. One set of experiments that have been carried out on soda lime glass use Hampden steel flyer plates to elevate the stress level to 5 GPa [Clifton et al., 1997]. Other shock compression tests were carried out by Dandekar [1998] and Sundram and Clifton [1998]. In Cliftons Hampden steel experiments, the glass was impacted at 302 m/s and exhibited a clear spall signal. No failure wave was observed in this experiment. At higher velocities, near 400 m/s, a failure wave was exhibited by means of a recompression of the material. No spall strength behind the failure wave front was observed in these experiments. The transverse compressive stress was also observed to increase behind the failure front, representing a decrease in the shear strength of the material [Clifton et al., 1997]. However, the steel exceeds its HEL at this stress level. By using tungsten carbide flyers which will remain elastic, the soda lime glass response can be observed with greater accuracy. By using shock compression-shear experimentation at an angle of 18 and varying the thickness of the specimen and flyer the spall strength and the shear strength of the glass can be observed. In examining soda lime glass both point and planar impacts will be studied. The point impacts enable the material cracking patterns and stresses of three dimensional impact to be characterized at varying velocities and thicknesses. The energy dissipation is also determinable from these experiments. This allows for the characterization of the impact process on a brittle material that has practical aerospace and military applications. In examining the failure wave at high velocities, the characterization of this phenomena is extended. Again, knowing the effects of this failure wave on material properties such as the spall strength, and the shear stress provides designers with information on the limitations of the material for its applications.

18

A.2. Silicon Nitride AS800 grade silicon nitride is a high strength ceramic that is under consideration to replace current nickel based superalloys in aircraft engine turbine blades due to its superior properties at high temperatures. For example, the melting temperature of silicon nitride is around 1800-1900C [Brandes and Brook, 1999] whereas the melting temperature for INCONEL MA 956, a nickel superalloy designed for use in aircraft engines by Special Metals is only between 1311C and 1400C [Matweb, 2005]. In creep testing, at 1350C, the bulk material samples of AS800 silicon nitride have a lifetime of 100 hours at a tensile stress of about 220 MPa and a lifetime of 1000 hours at about 170 MPa [Lin et al., 2001]. Whereas, INCONEL MA 754, which is also used in aircraft engine components, has a lifetime of 100 hours at about 110 MPa rupture strength and a lifetime of 1000 hours at about 100 MPa at 1093C [Brandes and Brook, 1999]. Silicon nitrides thus have superior properties of melting temperature and creep resistance to nickel based superalloys. However, the ductility of silicon nitrides is less than that of the superalloys. Ceramics in general are also not as thermally conductive and more likely to exhibit thermal shock than metallic based high temperature materials. However, silicon nitride in particular has a thermal shock resistance and toughness that recommends it for aerospace components [ASM, 1999]. The brittle nature of failure in the ceramic is still of concern. For this reason, the specific dynamic strengths of the proposed material must be explored through high velocity shock experiments. The AS800 grade silicon nitride was developed by Honeywell Engines. It was designed to have elongated grain structures and has been toughened [Choi et al., 2002]. It was also annealed prior to testing at 1200 C for 2 hours to reduce any residual machining stresses. [Choi et al., 2002]. Most of the grains were fine,

19

having a size of about 0.5 m. About 15 to 20 % of the grains were larger and elongated, with sizes of 1.5-2.0 m. The aspect ratio of these grains was between 5 and 12. Also, a second phase in the material was observed to be about 10 % of the total material. This second phase consists of Y10Si3N4O23. No differences were observed in the structures of the elongated grains and the second phase grains [Lin et al., 2001]. This grade of silicon nitride has been demonstrated to have favorable properties under high temperatures such as creep and high strength retention. Lin et al. [2001] studied these properties for AS800. The result is a material that has favorable properties for aerospace applications [Choi et al., 2002]. Silicon nitride has been examined under particle impact and its superiority to various other brittle materials has been established [Choi et al., 2002]. In this and other studies, silicon nitride was subjected to impact by 1.59 mm diameter steel spheres. The AS800 grade silicon nitride was compared with two other grades: SN282 and NC132. Flexure specimens were impacted at velocities ranging up to about 440 m/s. The residual strength of the material was then observed by means of a bending test [Choi et al., 2002]. A critical velocity for impact failure was observed for these three silicon nitride grades. The experiments indicate that AS800 has a critical velocity of 400 m/s, as opposed to 300 m/s for SN282, and 230 m/s for NC132. Also, for the specimens that did not fail under impact, the post-impact strength of AS800 is larger than that of SN282 between 220 m/s and 440 m/s. The key parameter in the increased foreign impact damage resistance is the fracture toughness. In fact, the fracture toughness is linearly related to the critical impact velocity. The AS800 has a fracture toughness of 8.1 0.3 MPa-m1/2. Additionally, an increase of fracture resistance is recorded when a copper layer is employed to reduce Hertzian contact stresses [Choi et al., 2002]. This and other studies have characterized the particle impact behavior of silicon nitrides.

20

Shock compression studies on silicon nitride have been performed. However they are not extensive and leave unanswered questions. Studies, such as Nahme et al. [1994] have been performed on silicon nitrides of varying densities. Similar experimental techniques as those utilized in the present work, a gas gun and VISARTM interferometer, was employed. One of Nahmes materials had similar density, 3.15 g/cm3, and longitudinal wave speed, 10.7 km/s, to the current studys AS800. These grades had small particles with 1 m size. Voids with diameters between 10 m and 100 m are responsible for the density difference. The resultant spall strength was measured and determined to be between 0.5 GPa and 0.8 GPa [Nahme et al., 1994]. However, the relationship of this spall strength to velocity was not discussed. Also, these experiments were normal shock compression with no shear effects. The Hugoniot Elastic Limit of the material has also been estimated using shock compression experimentation. In Nahmes paper, an HEL of 12.1 GPa is reported for silicon nitride with a density of 3.15 g/cm3 [Nahme et al., 1994]. Other papers have noted a variation of the HEL with grain size. For example, Mashimo [1998] quoted three HEL ranges for silicon nitrides that have grain sizes varying from 0.15 m to 1.2 m. The HEL is observed to decrease from 17-20 GPa in the smaller grain sizes (0.15 m to 0.3 m) to 10-12.5 GPa in the larger grain sizes (0.5 m and 1.2 m) [Mashimo, 1998]. This variation due to grain size means that establishment of the HEL for each grade of silicon nitride is necessary. Silicon carbide (SiC) is another ceramic material with properties that recommend it for aerospace uses. Silicon carbide exhibits a high HEL. This varies with microstructure, but is 11.5 GPa for -SiC (6H) [Feng et al., 1998] and between 13.2 and 15.7 GPa for three other types whose densities vary from 3.16 g/cm3 to 3.23 g/cm3 [Bourne and Millet, 1997]. In shock compression experiments performed by Bourne and Millet [1997], no spall strength was observed in specimens impacted at stress levels exceeding the HEL. The spall strength decreased as the stress approached the elastic limit [Bourne and Millet, 1997].

21

The shear strength of the SiC was shown to increase from 4.5 GPa at the HEL, when the stress level was 11.5 GPa, by a factor of two at 23 GPa [Yuan et al., 2001]. However, in Bourne and Millets experiments, a failure wave, such as that seen in the glass plate impact experiments was also seen [Bourne and Millet, 1997]. This failure wave caused reductions of the shear stress in two of three specimen microstructures of SiC that they studied. One microstructure shows a near constant shear level and another shows a decrease of 20% from the HEL to 1.4 times the HEL. During the same interval the shear stress in the material increases by 30 to 40% for all three microstructures before the failure wave and in the third specimen type after the failure wave [Bourne and Millet, 1997]. This failure wave degradation of the shear strength and the drop in the spall strength with increasing impact speed are of concern. Therefore, experiments examining the degradation in spall strength and observing the shear behavior of silicon nitride are performed in the current study. These and other papers leave open several issues of the dynamic strength which are examined in this study. First, the spall strength variation with both increasing velocity, is important to investigate. Also the inclusion of a skew angle and its effects on the spall strength are studied. The material HEL is examined and from it evaluate the dynamic yield strength is found. Also the strength of the shocked state of the material is determined. These results provide a significant increase to the available knowledge of the dynamic behavior of aerospace grade AS800 silicon nitride.

22

B. Intent of This Dissertation The study of brittle materials under point impact loading and planar shock compression is vital to the understanding of dynamic material behavior. The velocities necessary to perform experiments which test these conditions can be generated by means of compressed air gas guns, which fire projectiles at hundreds of meters per second. The result is the capability to examine the material properties under dynamic loading conditions that approximate real world impacts. In this dissertation, the effect of dynamic point loading on soda lime glass and that of planar shock compression on soda lime glass and AS800 grade silicon nitride are examined. Small particle impacts at speeds approaching sonic velocity conducted in this study are similar to real world impacts and can be used to predict damage patterns. Examples of these real world impacts include foreign impact damage in aircraft engines and projectile or runway debris impact on windows. The resulting damage and stress patterns in the materials are critical to determining the degradation in strength of the material. In order to examine the properties of soda lime glass under particle impact conditions, a projectile accelerator system was designed and constructed. This system, employing a compressed nitrogen firing system and a specially designed impact chamber allowed for measurements to be taken using several methods. The records included high speed photographs and surface strain readings. This system is discussed in Chapter II. Experimental results are covered along with theoretical and numeric studies in Chapter III. Previous research related to the impact of the materials is summarized and its relation to the current work is expressed. The various data collected provides a comprehensive picture of the variations with velocity and thickness in the cracking patterns, the energy dissipation, and the surface strain wave patterns.

23

Planar shock compression impacts are used to study the material properties in a state of uniaxial strain. This condition is achieved by the impact of two precisely parallel plates under normal impact loads. The condition enables the study of several basic properties such as the materials HEL and spall strength. The shock experiments were carried out on the existing single-stage gas gun. The experimental design and procedures are described in Chapter IV. Using the gas gun facility, three different sets of experiments were carried out. The shock response of silicon nitride is examined in Chapter V by employing shock compression experiments. The material begins to exhibit inelastic deformation in compression at the Hugoniot Elastic Limit, which is proportional to the dynamic yield stress. The Hugoniot Elastic Limit, or HEL, is an extrinsic property that represents the maximum stress to which the material can be loaded prior to plastic deformation. This represents the component of stress generated in the target plates by the impact in the direction of compression. The dynamic yield, on the other hand accounts for the HEL stress and the stress components orthogonal to the direction of impact. In other words, the dynamic yield stress represents the combination of the longitudinal and transverse stresses necessary to cause yielding. Also, the limit in resistance to tensile rarefaction wave induced spallation otherwise known as the spall strength is examined. The current study evaluates the change in spall strength with impact velocity under shock compression and under combined shock compression-shear loading. Additionally, from the data generated by the shock compression impacts, equations are written relating the materials elastic and hydrodynamic stress and strain response in the elastic and the beginning of the plastic regimes. Also the shock velocity versus particle velocity relationship in the elastic regime is examined. These two curve types are known as the equation of state and the Hugoniot curve respectively. In Chapter VI, further study of the dynamic yield stress of silicon nitride was conducted by employing shock re-shock and shock release experiments. These experiments make use of multiple loading states to examine the response of the deformed material to shock loading. In the shock re-shock and shock unload

24

experiments a two material flyer is used with the impedance of the second material being greater and less than the silicon nitride. By this means, two velocity profiles are generated which correspond to the reloading and unloading states. From this response the yield strength of the deformed material can be calculated. Another method involves a thin specimen of silicon nitride impacted by a relatively thick flyer plate. This causes reverberations within the flyer, which successively unload the specimen. This method also enables the examination of the deformed material. Reinhart and Chhabildas [2002] and Dandekar et al. [2003] used these methods, respectively to examine aluminum oxide and glass-fiber-reinforced polyester. Experiments using soda lime glass specimens and tungsten carbide flyers under shock compression-shear at 18 are conducted to observe the effect that the failure wave has on the spall strength and the shear stress. The failure wave data is discussed in Chapter VII. By employing angled projectiles, a shock compression-shear loading is produced. Varying the thickness of the specimens and flyer enable the examination of the spall strength and shear stress both before and after the passage of the failure wave. Additionally, the material impedance behind the failure wave is investigated by impacting tungsten carbide specimens with soda lime glass flyers. These experiments detail the behavior behind the failure front. The data collected provides a critical step in several areas. First, the soda lime glass particle impact study provides a critical look at the damage patterns at varying velocities and thickness, which can be used both to calibrate numerical models and to compare with other brittle materials. Second, the shock experiments provide a profile of the two materials. From these experiments, the effects of shear loading and velocity on various strength measures are observed. This information will help predict the dynamic response of AS800 grade silicon nitride and soda lime glass and help determine their suitability for aerospace and military applications.

25

References
Anderson, T. L. 1995. Fracture Mechanics: Fundamentals and Applications. CRC Press. ASM. 1999. Elevated Temperature Characteristics if Engineering Materials. Heat Resistant Materials. ASM Specialty Handbook. Bourne, N. K.; Z. Rosenberg; and J. E. Field. 1995. High-Speed Photography of Compressive Failure Waves in Glasses. Journal of Applied Physics. 78 [6] 3736. Bourne, N. K. et al. 1996. The Shock Wave Response of a Filled Glass. Proceedings: Mathematical, Physical and Engineering Sciences. The Royal Society. 452 [1951] 1945-1951. Bourne, N. and J. Millet 1997. Delayed Failure in Shocked Silicon Nitride. Journal of Applied Physics. 81 [9] 6019. Brandes E. A. and G. B. Brook ed. 1999. Smithells Metals Reference Book. Butterworth Heinemann. Brar, N. S.; S. J. Bless; and Z. Rosenberg. 1991. Impact-Induced Failure Waves in Glass Bars and Plates. Applied Physics Letters. 59 [26] 3396. Clifton R. J., M. Mello, and N. S. Brar. 1998. Experiments on soda lime glass: Shock Compression of Condensed Matter. Eds. Schmidt, Dandekar, and Forbes. 521-524. Choi et al. 2002. Foreign Object Damage of Two Gas-Turbine Grade Silicon Nitrides by Steel Ball Projectiles at Ambient Temperature. NASA/TM-2002-211821. Dandekar, D. 1998. Shock, Release, and Tension Response of Soda Lime Glass. Shock Compression of Condensed Matter - 1997. The American Institute of Physics. 1-56396-738-3/98. Dandekar, D. P., C. A. Hall, L. C. Chhabildas, W. D. Reinhart. 2003. Shock Response of a Glass Fiber-Reinforced Polymer Composite. Composite Structures. 61 [1-2] 51-59. Dowling, N. E. 1999. Mechanical Behavior of Materials. Prentice Hall. Espinosa, Y., et al. 1997. Micromechanics of Failure Waves in Glass: II. Modeling. Journal of the American Ceramic Society. 80 [8], 2074-2085.

26

Evans, A. G. and T. R. Wilshaw. 1976. Quasi-Static Solid Particle Damage in Brittle Solids I. Observations, Analysis and Implications. Acta Metallurgica. 24: 939-956 Feng, R.; G. F. Frasier; and Y. M. Gupta. 1998. Material Strength and Inelastic Deformation of Silicon Carbide Under Shock Wave Compression. Journal of Applied Physics. 83 [1] 79. Field, J. E. 1988. Investigation of the Impact Performance of Various Glass and Ceramic Systems. U.S. Army: Final Technical Report Contract Number DAJA45-85-C-0021. Field, J. E., Q. Sun, and D. Townsend. 1989 . Ballistic Impact of Materials. Inst. Phys Conf. Ser. No 102: Session 7. Oxford. Frank, F. C. and B. R. Lawn. 1967. On the Theory of Hertzian Fracture. Proceedings of the Royal Society. A229. 291. Ginzberg A. and Z. Rosenberg. 1998. Using Reverberation Techniques to Study the Properties of Shock Loaded Soda-Lime Glass. Shock Compression of Condensed Matter 1997. The American Institute of Physics. Kanel, G. I. et al. 1992. The Failure Waves and Spallation s in Homogeneous Brittle Materials. In: Shock Compression of Condensed Matter 1991. Elsevier Science Publishers. Lawn B. R. and E. R. Fuller. 1975. Equilibrium Penny-Like Cracks in Indentation Fracture. Journal of Materials Science, 10(12): 2016-2024. Lawn B. R. and M. V. Swain. 1975. Microfracture Beneath Point Indentations in Brittle Solids. Journal of Materials Science, 10: 113-122. Lawn B. R. and T. R. Wilshaw. 1975. Review Indentation Fracture: Principles and Applications. Journal of Materials Science. 10 1049-1081. Lawn B. R., T. R. Wilshaw, and N.E.W. Hartley. 1974. A Computer Simulation Study of Hertzian Cone Crack Growth. International Journal of Fracture, 10(1): 1-16. Lin H. T. et al. 2001. Evaluation of Creep Property of AS800 Silicon Nitride from As-Processed Surface Regions. Ceramic Engineering and Science Proceedings. 22 [3] 175-182. Mashimo, T. 1998. In: High-pressure shock compression of solids III. L. Davison, M. Shahinpoor, editors New York : Springer 101-146.

27

Matweb. 2005. www.matweb.com Nahme, V. Hohler, A. Stilp. 1994. Determination of the Dynamic Material Properties of Shock Loaded Silicon-Nitride. American Institute of Physics. 765-768. Reinhart, W. D.; L. C. Chhabildas. 2002. Strength Properties of Coors AD995 Alumina in the Shocked State. International Journal of Impact Engineering. 29 [1-10] 601-619. Rosenberg, Z. et al. 1985. Spall Strength of Shock-Loaded Glass. Journal of Applied Physics. 58 [8], 3249-3251. Sundaram, S. and R. J. Clifton. 1998. Flow Behavior of Soda-Lime Glass at High Pressures and High Shear Rates. Shock Compression of Condensed Matter 1997. The American Institute of Physics. 1-56396-738-3/98. Yuan, G.; R. Feng; and Y. M. Gupta. 2001. Compression and shear wave measurements to characterize the shocked state in silicon nitride. Journal of Applied Physics. 89 [10] 5372.

28 Chapter II Configuration and Procedures for Spherical Particle Impact Experiments Aerospace and military vehicles are subjected to impacts under normal operating conditions. For example, windows on aircraft and ground vehicles can be impacted by shrapnel or runway debris. The response of brittle materials like soda lime glass to dynamic impacts is therefore critical to their use as windows. In the current study, spherical ball bearings are employed to impact soda lime glass. Spheres cause stress waves and cracks that are easier to predict using closed form equations than those from a more complex geometry of impacting particle. The experimental behavior of the simple geometry may be used to perfect a numerical model for more complex projectiles. Correlation between the equations, numerical modeling, and the experimental results can be used to explore the material behavior. Studying brittle materials under dynamic impact conditions required the creation of a mechanism to accelerate projectiles consistently to given velocities and to a precise location on the target. The specimens that were chosen for this phase of the study were soda lime glass plates 50 mm square with thickness of 3 mm, 5 mm, 15 mm and 25.4 mm. Soda lime glass is a brittle, transparent material. Velocities of between 300 and 400 meters per second were desired to simulate real world collisions with debris. An appropriate particle delivery system needed to be determined to accelerate steel ball bearings of 1/16th inch diameter. It is the focus of this chapter to describe this system. By observing the impact directly, the time of contact of the projectile, its impact velocity, and rebound velocity can be determined. Additionally, the cracking patterns can be tracked as they propagate. To accomplish this, a high speed camera is employed. This camera takes pictures with microseconds between frames in order to capture the motion with sufficient accuracy. Surface strain profiles indicate the propagation of stress waves. To observe these strains, gages are placed on the front or rear surfaces. On the front surface, the gages measured strains both along the expected direction of wave propagation and perpendicular

29 to the direction of propagation. The rear surface gages are located opposite the impact site. These two data collection methods provided the means to examine the energies, cracking patterns and surface strains involved in the impact. From the impact and rebound velocities, the coefficient of restitution and total impact energy were determined. Combined with theoretical equations and numerical modeling, these values are used to examine the partitioning of energy during the impact process. The velocity of impact and the thickness of the specimen categorized the damage done by cracking patterns. Experimental strain records were compared with those from the closed form solutions and the numerical model. The effects of velocity and finite thickness on the strains are determined. These results are discussed in Chapter III. A. Components To this end, a system was designed in several phases (Figure 2.1). First, a firing chamber and gun barrel were designed to accelerate the desired projectiles to the specified velocities. Secondly, the target was surrounded with a chamber to contain the impact while allowing for visual observation of the impact and access for other measuring elements. An impact system was designed that consists of both the afore mentioned impact chamber, as well as mechanisms for orienting the specimen and safely stopping the projectiles components. A.1. Projectile Firing Mechanism In order to carry out these requirements, it was decided to utilize an existing design for a Hopkinson compression bar firing chamber. The design consists of a flanged cylinder capped on both ends by 12 inch square steel plates which are one inch thick. These end pieces are connected with 12 5/8 inch diameter bolts to the flanged cylinder. Polytetrafluroethylene gaskets seal the pieces. These plates contain various connections to allow the inflow and outflow of high pressure air. One end piece contains two holes machined with a three-quarter

30 inch and a one inch NPT thread for air intake and exhaust respectively. The other plate has a one inch through hole and four 5/8 inch connection bolt holes for the barrel. It also has a inch NPT threaded hole with a 1/8 inch through hole inside of it for a pressure gage and trim valve. Contained within the chamber is a six inch diameter piston, which is constructed of aluminum to reduce mass and has two rubber o-rings to provide a seal along its outer rim. This piston is connected to a shaft with a plug on the other end, which when the chamber is filled seals the connection to the gun barrel. This design arrangement allows for a reservoir of high pressure air to build up in a chamber on both sides of the piston (Figure 2.2). Filling the chamber is accomplished by means of standard plumbing connections. A compressed nitrogen tank is utilized to provide an air source. Typical loading pressures are between 100 and 500 psi. Two 1/16 inch diameter holes bored through the piston allow the equalization of pressure between the two sides of the piston during the pressure buildup stage. A valve on the back plate is then opened to atmosphere, causing the piston to pull back and open the breach. The remaining air in front of the piston then pushes the projectile down the barrel. This firing chamber is mounted upon a wheel and track setup which allows the whole firing chamber and barrel arrangement to be adjusted in distance from the impact chamber. The track is made of L section steel beams and is 24 inches long. The four wheels are four inches in diameter and are anchored to both sides of the end plates. This track system allows for easier loading of the projectile and unloading of the components for maintenance. The several components of the firing chamber are the cylinder, which holds the piston, the end plates, and the wheel and track system. The firing chamber design provides a simple and repeatable method of providing the force for accelerating the projectile. The entire firing chamber and its track are supported on an aluminum plate above a set of I-beams. The gun barrel and sabot stripper are also attached to this surface. These I-beams are attached to the table upon

31 which the impact chamber is located. This arrangement accounts for the differences in height between the axis of the gun barrel and the target location. A.2. Gun Barrel and Sabot Stripper The gun barrel was designed to use a three-quarter inch diameter cylindrical sabot and carry it for four and a half feet (1.37 meters). The length of 4.5 feet was chosen by calculation of the best length for achieving a final velocity in the 300 to 500 m/s range. The barrel was honed at Commercial Honing and machined with a 1 inch NPT thread at one end to screw into a flange connected to the front of the firing chamber. This connection, the breach, is unscrewed to load the projectile (Figure 2.2). Velocity is confirmed by means of a laser velocity system that measured the sabot emerging from the barrel end and by camera measurement of the projectile in flight. This method enables a calibration curve that refines the calculated firing pressure versus velocity measurement. The calibration curve is discussed further in Chapter III. The system of sabot and projectile was chosen because it is more versatile than using a straight projectile arrangement. With a cylindrical sabot, the projectile can be any shape and size (smaller than the sabot). A hole that conforms to the projectile shape is machined into the center of the sabot in order to contain the projectile. A three-quarter inch diameter sabot was chosen (Figure 2.3). When combined with the actual projectiles; in this study, spheres of diameter one sixteenth of an inch, the combination masses about 15 grams. removed along the way. In order to separate the sabot from the projectile, a sabot stripper was devised that uses a steel piece called an upright with a 7/16 inch diameter hole in it to block the sabot while allowing the projectile to continue unimpeded. The stripper consists of this upright and a supporting piece (Figure 2.4). The combination of both provides both the mass to stop the sabot and the area to prevent the fragments of the sabot from hitting the specimen. Both of these This arrangement requires that only the projectile itself continues to the specimen and the sabot be

32 components are made from 1040 steel. The sabot stripper prevents the sabot from entering the impact chamber, while passing the actual projectile through. A.3. Impact Chamber & Target Holder The impact chamber is constructed of 1040 steel and clear polycarbonate. The

side walls of the chamber are constructed of half inch steel to prevent shrapnel from escaping. The top, front wall, rear wall, and the portholes in the sides are made of clear shatterproof Lexan, a polycarbonate material. The chamber is 30 inches long, 17.5 inches wide and 19.75 inches high. This design contains room for the sabot stripper, target holder, and its support apparatus (Figure 2.5). The front of the chamber has a 2.25 inch diameter hole for the sabot and projectile to pass through. Portholes are located in each side wall to allow for the viewing of the impact. These portholes are covered by Lexan shields to prevent shrapnel from damaging cameras and other important instruments. The holes are covered in such a way that they can be opened to allow the entry of strain gage wires and other equipment. Holes are also machined in the front lexan plate to allow for the entry of the laser beams of the velocity system. A steel shelf is included for the placement of optical instruments such as mirrors and beam splitters. Lexan shield protects optics on the shelf from stray shrapnel. The impact chamber contains the aforementioned sabot stripper as well as a target holding apparatus that consists of multiple stages allowing for both angular and linear alignment (Figure 2.6a). The target holder can articulate angularly in two directions using pivots. The connections between the first three stages employ such pivot points that enable coarse rotation of the specimen around the vertical axis and around the horizontal axis perpendicular to the projectile motion. Fine angular alignment occurs by means of three adjustment screws that connect the third stage to the linear adjustment stages. Linear alignment is carried out on the fourth, fifth and sixth stages by means of screw locked slides that enable the target to be placed precisely in the path of the projectile. The entire target holder can also be adjusted in distance along the A

33 projectile path by means of a slide at its base. The target itself is held in a two screw clamping arrangement that does not impede measurements from the side or from the rear (Figure 2.6b). This device is shaped as a square C-section of metal. The two prongs of the C are 1.5 inches long by inch thick. From these ends, smaller C-sections protrude. Two screws threaded through the back surface of the C hold the specimen in place against these smaller C-sections. The base of the C has a two inch diameter circle machined in it. This circle is used to observe the rear of the specimen. Side observations can also be made because the sides of the C are open. This arrangement allows the specimen to be securely clamped while allowing it to be viewed from three directions (both sides and the back). In this way, the specimen is secured and properly aligned with the path of the projectile.

34 B. Measuring Systems Observations of specimens are made by two real time methods: crack pattern investigation by high speed camera and point-wise deformation examination by strain gauges. DRS Hadland provided high speed cameras that are used to capture the internal cracking patterns of the soda lime glass during the impact. This allows for comparison to static cracking patterns and their effect on stress fields. Strain gages give surface measurements of strains on both the front and rear surfaces. With these methods, the state of the deforming material is examined. B.1. High Speed Camera The DRS Hadland ULTRA 17 and IMACON 200 high speed cameras allow for a small set of pictures to be taken within a very short span of time. The ULTRA 17 camera takes pictures at a rate of up to 150,000 frames per second. That is, one frame every 6.67 microseconds. The IMACON 200 has a maximum framing rate of 200,000,000 frames per second and takes 16 pictures. The frame rate is adjustable to allow for the best capturing of the material behavior. These cameras provide a series of images of the changes in the material from which the impact and rebound velocities as well as the contact times can be observed. In order to best capture the various images, the high speed camera can be configured for different types of inputs. For simple visual imagery, the camera uses a Photogenic Power Light 2500 DR flash which delivers 1 kJ within the span of about 200 microseconds. This flash illuminates the target either from the side (Figure 2.5). magnification. The camera and flash are connected to a triggering system that is setup to allow for the maximum amount of light to illuminate the target while the pictures are being taken at the earliest possible time while allowing for accurate timing. The laser velocity system, discussed in section B.5., is used to send a trigger signal for The camera is then placed at an optimum distance for the required magnification and configured with appropriate Nikkon lenses for the best

35 the camera and through a delay generator to the strain gages. The optimal delay time varies with velocity. For example, it is about 310s for a 150 m/s shot. In order to account for some degree of inaccuracy arising from fluctuations in velocity in each individual experiment, the camera is set to a small frame speed, such as 200,000 frames per second or 150,000 frames per second to ensure capture of the entire impact. Once the imagery is obtained, it is processed through a proprietary program of DRS Hadland. This program outputs either single images for each of the seventeen frames, or a movie file. The images can be specifically analyzed in the program to obtain the impact and rebound velocities of the ball bearing. The images and their results are discussed further in chapter III. B.2. Strain Gages Stacked rosette strain gages are utilized to gain a picture of the radial and circumferential (hoop) strains on the surface of the glass during the impact process. These gages, CEA-06-032WT-120, from Vishay Measurements Group, use two perpendicular gages placed one on top of the other in order to get the measurement of the strains in two perpendicular directions at the same point. (Figure 2.7). These gages are made of constantan alloy with self temperature compensation close to the soda lime glass compensation, 120 ohm resistance, and copper tabs for soldering. Placed upon all of the specimens at a radial position 5-20 mm away from the projected point of impact, these gages provide readings of strain very close to the impact location. This provides a time history of the strains moving parallel to the surface of the specimen upon impact. By positioning the gage so that one component of the rosette is parallel to the radius, the radial and circumferential (hoop) surface stresses are measured.

36 B.3. Laser Velocity System It is extremely important to ensure the velocity of each projectile due to individual variations in sabot firing conditions. Therefore, in addition to the approximate velocity determined by the calibration curve between velocity and firing pressure, actual velocity measurements are taken during each shot. A laser system was chosen to accurately measure the velocity. The laser trigger for the camera provided by DRS Hadland was converted into a velocity measurement system by means of a pair of 5 mm long cube beam splitters that create two parallel beams which are projected across the path of the firing sabot. The beams are 5.9 mm apart. This is done despite the slight drop in velocity caused when the impacting particle separates from the sabot at the sabot stripper because the one-sixteenth inch diameter particle is much more difficult to measure with a laser beam than the three quarter inch diameter sabot. The sabot passes through the two beams in sequence and the drop in intensity each time is recorded by means of a high speed photodiode (Figure 2.5). These drops in light intensity both trigger the camera system and also provide accurate velocity measurement.

37 C. Experimental Procedures An experimental procedure was developed in order to perform the experiments in the most effective fashion. This procedure contains instructions for the preparation of the specimens with the application of the strain gages. It also provides a method for standardized sabot production. Both of these steps are done in groups so as to minimize the time of preparation. Also, the method of aligning the specimen and setting up the projectile accelerator and the various measuring systems is discussed. Finally, the procedure for firing the projectile is elucidated. This total procedure allowed for efficient experimental performance. C.1. Series Preparation Before each experiment, both the specimen and the projectile were configured. The specimen was prepared with a strain gage to measure the surface strains. The small projectile was glued to the larger sabot. These processes are described in detail below. C.1.1. Soda Lime Glass Specimen Preparation The specimens were provided by American Precision Glass. These specimens were procured as a group of 60 and are made from soda lime glass. The specimens are rectangular prisms with two inch by two inch (50.8 mm) square sides as the large faces. The thickness of the specimens comes in four varieties: 3mm, 5mm, 15mm and 25.4mm. The specimens were required to have dimensional tolerances of 0.1mm, flatness of 5 to 25 wavelengths of light and a parallelism of 3 to 5 arc minutes. In order to prepare the glass specimens, several steps are taken. First, both the rear and front surfaces of the specimens are cleaned. Next, strain gages are attached to the front face of each specimen. The gages are glued to the front surface in precise orientation to the center of the specimen. After a 24 hour curing time, wires are soldered to the gages. Care is taken during this process

38 not to contaminate the other surfaces. C.1.2. Sabot/Projectile Preparation The projectiles that were used to impact the specimens are made of hardened chrome steel obtained from the Bearing Service Company in 1/16 inch diameter ball bearing form. These projectiles are encased in sabots that are stripped off prior to impact. The sabots allow a wide range of projectile sizes (anything below 7/16 inch diameter) and shapes to be fired from the same gun. The sabots are constructed of nylon 6/6 so as to be easily breakable. They are made of a cylindrical piece that is one and a half inches long with a three quarter inch outer diameter. from McMaster Carr. These parts are machined from nylon bar stock obtained The individual pieces are then cut from the bar and

machined so that the front end is perpendicular to the sides and to remove rough edges left by the cutting process. The design is such that the sabot hits the sabot stripper and splits apart. After the sabot pieces are machined, a hole of diameter 40/1000 inches is drilled in the exact center of the forward face. This hole has a depth of 1/16 inches so as to allow the projectile to be centered on the sabots front face. Following this, the hole is filled with glue and the projectile and sabot are pressed together. Once the adhesive is dry, the sabot is ready for use. The sabot and projectile are then covered with vacuum grease to reduce friction in the barrel. The gun breach (the connecting piece between the firing chamber and the barrel) is then unscrewed and the sabot is inserted into the barrel, projectile end facing towards the impact chamber. The breach is then sealed.

39 C.2. Setup Sequence Following the preparation of the specimen and the sabot/projectile combination, and the loading of each of these into their proper places, each experiment proceeds along a similar process. This process includes the alignment of the various components of the projectile accelerator. Also included is the setup of the measurement systems. C.2.1. Projectile Accelerator Alignment First, the alignment between barrel and sabot stripper is made placing a sabot with its glued on projectile in the front end of the barrel and pushing the barrel and firing chamber forward until there is contact. The best alignment is when the projectile is centered inside of the sabot strippers 7/16 diameter hole. In order to adjust the alignment, the steady rests (which hold the barrel in a specific alignment by means of three connecting rods) are unlocked. The barrel is then adjusted so that the projectile is centered on the sabot strippers hole, and the steady rests are re-tightened. Next, the alignment between the target and the sabot stripper occurs. The

specimen is placed in the C-section and the screws are tightened. The C-section is then attached to the rest of the target holder. The target holder is manually adjusted angularly first until it is parallel with the sabot stripper. Then, the linear slides are adjusted until the specimen center is along the axis of the barrel in both the vertical and horizontal directions. Finally, the screws for the linear slides are locked in place. C.2.2. Measurement System Preparation For velocity measurement and triggering of the flash prior to impact, the laser trigger is used. As mentioned above, the velocity measurement system is a pair of beam splitters that are employed to split a single laser beam into two useful parallel beams (and one perpendicular beam that is not used). This system is set as close to the impact chamber as possible in order to minimize flight time for

40 the projectile. For this reason, the beams pass in front of the sabot stripper. The system is setup according to the diagram in Figure 2.5. The elements must be properly positioned so that both laser beams meet at the receiving diode. Following the laser setup, the exact distance between the laser beams and the specimen is measured. From this and the anticipated speed of the projectile the time from triggering to impact can be determined. When the camera is going to be used, the velocity measuring oscilloscope must be hooked up to the delay generator which triggers the camera and the oscilloscope attached to the strain gage at the appropriate time. The delay generator has multiple channels. One channel, for the strain gage oscilloscope, is set to a delay time ten to 15 microseconds shorter than the estimated impact time. This allows for the proper triggering of the strain gages. Another channel is set to 10 nanoseconds of delay time, and is used to trigger the camera. This system ensures that the camera triggering and the strain gage triggering are always separated by a known time factor. The DRS Hadland cameras are used for impact and rebound velocity measurement, as well as qualitative cracking imagery. The camera is setup by first selecting the appropriate lens and extension tube combination. The lens and however many doublers and extension tubes needed to focus at the proper distance and magnification are attached to the camera. The camera is then connected to its trigger, the flash, and the computer. The camera computer recording program is then setup. The delay time to firing is input here and the frame rate and exposure time are also programmed. Finally, the camera is tested to insure the aim and focus are correct. Another system that is always in use are the strain gages. The strain gages are already attached to the specimen. There are three wires for each gage; one red, black and white wire. The black and white wires have previously been soldered to the negative terminal on the gage and the red wire is connected to the positive terminal. All that is required is for the six wires to be threaded out of the

41 chamber in such a manner that they do not pass in front of any other recording device. The gage wires then connected to the correct terminals on the circuit box. This box contains Whetstone quarter-bridge circuits that accurately determine the strain. This circuit box is connected to the amplifier which magnifies the output voltage one thousand times. Once this is done, the amplifier is turned on. An input voltage of 3 volts is used for the experiment. Using the knobs on the amplifier with calibration switches off, Whetstone bridge circuits are balanced. The shunt voltage is checked (the calibration coefficient is about 1000 - per 1.5 volts). C.3. Pre-Firing Sequence Following the attachment of the strain gages and the setup of the laser velocity system and the camera, the two recording oscilloscopes are set for the experiment. The oscilloscope settings are checked, confirming the voltage and time ranges are appropriate for the experiment. The trigger levels are checked to ensure triggering at the appropriate time. Once this is complete, the oscilloscopes are set for single acquisition and the ready lights are is ensured to be on. At this time, the high speed camera is also armed. Next, the fire, input and trim valves on the firing chamber are all closed. The compressed air tank, which has two valves, the main valve and the pressure control value, is opened. The high pressure valves, both the main valve and the pressure control valve, and the firing chamber intake valve are opened and the load chamber is filled to the desired pressure. The load pressure must be greater than the desired firing pressure by about 50 to 100 psi to ensure rapid loading. The tank pressure and the load pressure are recorded. Once the desired firing chamber pressure has been reached, the main tank valve, the tank pressure valve and the intake valve are closed in sequence. All recording devices are checked to confirm that they are armed and ready. The pressure is verified to be at the

42 desired firing level. The firing pressure is recorded. The firing valve is opened, releasing the sabot and performing the experiment. Following the experiment all data is recorded from the oscilloscopes and the camera computer for later analysis. D. Summary This design of equipment, experiments and setup procedure were established in order to examine the impacts of spherical steel projectiles on brittle materials. The gas gun system utilizes compressed nitrogen to propel a nylon sabot down a 4.5 foot barrel at speeds up to 400 m/s. Following impact with a sabot stripper, which removes the nylon sabot, the steel projectile impacts the specimen. This specimen, held and aligned by a target holder, is located inside of a protective impact chamber. Laser velocity measurements are taken of every experiment. Instruments such as a high speed camera and strain gages record the strains of the specimens and velocities of the projectiles as well as the cracking patterns of the specimens. The laser velocity measurement system and the high speed camera provide a comprehensive picture of the relationship between the firing pressure of the chamber and the resulting velocity of the sabot and projectile. With this calibration curve, target velocities can be reached consistently. Also, in observing the velocities of sabot versus projectile, it can be seen that little to no energy is lost during the sabot stripping process. This verifies that almost all the kinetic energy is retained by the projectile following the impact with the sabot stripper. Armed with this data, the primary experiments of the soda-lime glass phase were undertaken. Experiments utilizing the camera and the strain gages in the discussed configurations were performed. This data provides a fundamental understanding of the process of dynamic impact useful as a comparison to numeric simulations of dynamic impacts of this and other brittle materials.

43 References
Vishay Measurements Group online handbook, 2002. www.vishay.com/brands/measurements_group/guide/500/gages/032wt.htm

Figures

Figure 2.1: Complete view of projectile accelerator

Figure 2.2: View of firing chamber with piston and sabot loading procedure

44

Figure 2.3: Sabot and projectile assembly diagram

Figure 2.4: Sabot stripper diagram

45

Figure 2.5: Impact chamber diagram and Measurement Setup

46
a)

b)

Figure 2.6: Target holder diagram: (a) Target holder overview. (b) Close up of specimen holder

47

Figure 2.7: CEA-06-032WT-120 strain gage image (Vishay Measurements Group, 2002).

48

Chapter III Spherical Particle Impact Experiments on Soda Lime Glass


In this chapter, the results of a series of experiments are discussed that were carried out in order to better understand the impact of spherical ball bearings on soda lime glass plates. For the study four different velocities and plates of different thickness of soda lime glass were chosen. These results are compared with theoretical analysis, numerical simulations [Nathenson et al., 2005], and previous experimental work. Three specific types of results are analyzed: surface strain profiles, impact kinetics, and cracking patterns. These results of the experimental study are consistent with the previous experimental work in the literature, but provide some unique insights when coupled with the numerical study and theoretical equations. A. Background on Particle Impact of Brittle Materials Stress waves are created in any impact. At low impact velocities, these waves have small amplitudes and their effects are negligible. However, when the impact velocity is sufficiently large the inertia of the material plays a role in controlling its response [Spath, 1961]. There are two ways to approach the study of dynamic impacts. In the long time solution, stress waves have been given sufficient time to reverberate and bring the entire specimen to a state of equilibrium. This method uses quasi-static methods to study the impact while accounting for the impact velocity. The partitioning of impact energy and the coefficient of restitution can be studied with this method. The second method involves the study of stress waves. This analysis considers the transient effects of the stress waves generated by the impact. The impact of the sphere on a half space generates waves that travel outwards from a central point either as circular waves on the surface, or as spherical waves within the body. There are three main types of waves, the Rayleigh surface wave, the longitudinal (pressure), and the transverse (shear) waves [Lamb, 1904]. The combined effects of these waves are used to develop theoretical models to predict the material behavior under impact. These models will be compared with numerical simulations and experimental measurements of surface strains generated by the waves.

49

For the long time solution, the results from high speed photography of the experiments are compared to predictions of the Hertzian theory, along with a LSDYNA 3D computer model. In this way, the long time solution is probed and the coefficient of restitution and impact energy partitioning are examined. The high speed photography and post impact examinations are used to study the cracking patterns. These findings are compared with those described in the literature. By correlating the results from the surface strain gages with the results from three theoretical models for the surface strains and the LS-DYNA numerical simulation, the stress wave response is categorized. Before the results are discussed, the previous literature on theoretical formulations of particle impact and their predictions are reviewed. A.1. Theoretical Calculations The first step in understanding high velocity particle impacts is to discuss the varied existing theoretical models and their predictions for understanding transient and long time solutions for various quantities of interest. In particular, these models will provide the framework for interpreting the data collected in the experiments. Hertzian theory for quasi-static impacts provides a method of examining the impact and rebound energy, coefficient of restitution, peak force, and contact time. The experimental data is compared to the results of these equations and to numerical simulations. In Appendices 1 and 2 the equations for the surface displacements generated by the dynamic stress waves are derived. By using the derivatives of these equations two estimates of the surface strain can be obtained. These formulae from Lamb [1904] and Mitra [1964] provide estimates for different geometrical conditions. Plots of these equations for strain versus time are compared to the experimental strain records and to numerical simulations.

50 A.1.1. Hertzian Theory The Hertzian theory for a spherical particle impacting on a half space is subject to the restrictions of the Love Criterion [Love, 1944]. This criterion defines the limiting maximum impact velocity for which the quasi-static assumptions of Hertzian theory are accurate.

( )
vC

1.

{III.A.1}

In equation {III.A.1}, v, represents the impact velocity and, C, is the longitudinal wave speed. Quantities that can be estimated by means of this theory include the peak force and total contact time. Also, the impact energy and coefficient of restitution can be determined [Hunter, 1957]. These are determined by the equations shown in their respective sections. The predictions of these equations can then be compared to experimental measurements determined by means of the strain gage and the high speed camera, which are discussed in Section C below. The Hertzian force versus time profile for a spherical projectile on a half space specimen is approximated as a sine curve [Goldsmith, 2001]. This is for elastic conditions where neither the sphere nor the half space deform plastically. The equations still approximate the true behavior even if a small amount of plasticity in the projectile occurs, as long as the specimen remains elastic. The following equations describing this force versus time profile are taken from Goldsmith [2001]. The subscript 1 refers to the sphere and subscript 2 refers to the half space in the equations. 1.140v 2 1.068vt sin , k1m m F (t ) = 0, m 1.068v m t 1.068v

0t

{III.A.2}

Here, Goldsmith approximates the force between two impacting bodies as a symmetric sine curve. In equation {III.A.2}, the impact velocity is, v, and the

51 time is, t. Here and always in this thesis, , is the numeric quantity Pi. Also, k1 and, m , are defined by equations {III.A.3} and {III.A.4}, respectively, and are given below: 1 3 = , m1 41R13

k1 =

{III.A.3}

and, 15v 2m1 (1 + 2 ) 5 . m = 16 R1


2

{III.A.4}

In the above equations, m1, is the mass of the sphere, 1, is the density, and, R1, is the radius of the sphere [Goldsmith, 2001]. The quantities, 1, and, 2, for the sphere and half space respectively are defined as follows:
1 1 12 , E1 1 1 22 , E2

1 =

{III.A.5}

2 =

{III.A.6}

In equations {III.A.5} and {III.A.6}, E, is modulus of elasticity, , is Poissons ratio, and, , is the mass density [Goldsmith, 2001]. For the experiments, the mass density, Youngs moduli and Poissons ratio values are 2,500 kg/m3, 79.2 GPa and 0.19 respectively for the soda lime glass plates, and 7,800 kg/m3, 201 GPa and 0.3 for hardened chrome steel spheres [Matweb, 2005]. The Hertzian force versus time equation {III.A.2} results in a function that is plotted in Figure 3.1. As can be seen, the impact force rises to a maximum force at half of the total time of contact and then falls to zero symmetrically. This results in a maximum force when the time is half of the total contact time, tMAX:

52 tMAX m = . 2 2.136 v The corresponding maximum force is: {III.A.7}

FMAX = F tMAX 2 = 1.140

m1v 2 . m

{III.A.8}

The total contact time of the sphere and half space can also be found using the densities, Youngs moduli, and Poissons ratios of the projectile and specimen: (1 + 2 ) m1 5 . = 4.53 R1vI
2

TCONTACT = TMAX

{III.A.9}

The results from these theoretical equations for the forcing versus time curve, contact time, and peak force are compared with the force vs. time function determined in a numerical study using LS-DYNA performed by Guodong Chen [Nathenson et al., 2005]. These curves are also plotted in Figure 3.1. Close agreement between the theoretical and numerical results is an indication of the accuracy of the numerical model. The forcing function could not be determined experimentally, but the contact time could be estimated with the high speed camera. The numerical, theoretical, and experimental contact times are discussed further in the Section C. A.1.2. Computed Longitudinal, Shear and Rayleigh Wave Profiles The use of strain gages to measure strain waves on the impact surface of the material provides a method for examining the stress wave propagation within the specimen. These experimental results can be directly related to both theoretical and numerical calculations. In measuring the strain profiles, certain waves are expected. The longitudinal and shear waves in soda lime glass, which are spherical in shape and disperse with increasing distance travel approximately at speeds of 5300 m/s and 3500 m/s respectively [Field, 1988]. The Rayleigh wave

53 does not disperse with increasing propagation distance and is cylindrical in nature. It travels at the slower velocity of 3150 m/s [Goodier, 1959]. On the surface, an effect of the three stress waves is strain waves propagating in the radial direction outwards from the impact point. The strain waves, which are measured by the surface gages in the experiments, are discussed in this chapter. In Appendices 1 and 2 two different theoretical estimates of the surface displacements associated with these stress waves are calculated. Each estimate has its own impact geometry. By taking the time derivative of the equations for the surface displacements, the surface strains are estimated. These strains, estimated at the strain gage location on the surface of the specimen, are compared with the strains measured in the experiments and simulated numerically. A.1.2.1. Lambs Solution Lambs examination of a point loading on the surface of a half-space is the first geometry [1904] discussed in this section. This geometry is the simpler of the two discussed. A result of Lambs [1904] analysis using approximations (Appendix 1) is an integral equation for the radial displacement at a the strain gage location along the surface due to the Rayleigh wave:
Rc 4 2
3 2 sin ( ) cos 2 4 2 c 3 sin ( 32 ) cos 2 d 4

qo H

2 U ( ) 2 b 0

{III.A.10}

where,

( 2 a 2 ) (b 2 2 ) U ( ) = , 4 (2 2 b 2 ) + 16 4 ( 2 a 2 )(b 2 2 )
2 = x2 + y 2 ,

b 3 (2 2 b 2 )

{III.A.11}

{III.A.12}

54

{III.A.13}

and,
t = tan1 .

{III.A.14}

For a specific loading function at a point:


R (t ) = R
2

t + 2

{III.A.15}

The terms in these equations are defined as follows: qo, represents the Rayleigh wave solution for radial displacement, x, and, y, are coordinates on the plane of the half space, t, is time; and, R , and, , are variables of the loading function {III.A.15}. Also, H, is a function of the two Lame constants, , and, ; c, is the inverse of the Rayleigh wave speed; a, is the inverse of the longitudinal wave speed, and, b, is the inverse of the shear wave speed. Solving these equations numerically produces an estimate for the radial surface displacement. In order to obtain the strain, the equation must be differentiated by time. The onedimensional nature of the equation means that differentiating by radial distance,

, is the same as differentiating by time, t.


q t = q .

{III.A.16}

The equation is one-dimensional because there is no estimated wave propagation in the hoop or vertical directions. Because of the complexity of this equation, numerical differentiation was employed. The numerical differentiation of equation {III.A.10} yields the dynamic surface radial strain [Goodier, 1959].

55 Lambs [1904] solution for strain at various radii from the impact point results in an estimated wave pattern. Plotting this wave shows one peak (Figure 3.2) that decreases in amplitude and spreads out with increasing radius. Also, the initial depression attenuates, as the distance from the impact point radius increasing. Lamb [1904] also states that the calculation in equation {III.A.10} represents only the Rayleigh wave. Because no information is contained in the solution about the longitudinal or shear waves, no statements can be made excepting what is in Lambs paper as to their intensity relative to the main pulse. Lamb [1904] states that the amplitudes of the shear and longitudinal waves are much smaller than that of the Rayleigh wave. This case represents the general solution for the Rayleigh wave generated by a loading at a point on a half space. A.1.2.2. Mitras Solution The closest case to the experimental geometry used in the current study that has been computed in closed form was by Mitra [1964]. This is the case of an impulse loading on a circle on the surface of a half space. The solutions are presented in terms of elliptical integrals. The coordinate system for this solution is cylindrical. The solution, u (r , z = 0, t ) , is on the surface, so the z coordinate is zero in the following equations. It is axi-symmetric so there is no variation in the

direction. The entire solution, u (r , 0, t ) , on the surface is given in terms of


both the Rayleigh wave component, uo (r , 0, t ) , and the shear and longitudinal components, which involve, U():

P u (r , 0, t ) = u 0 (r , 0, t ) + 2 The terms are defined:

U ( ) cos d . R

{III.A.17}

56
0 2Pa V u 0 (r , 0, t ) = [2E (k ) K (k )] ar V r 2 + a 2 V 2t 2 1 1 V 2t 2 (r a )2 E ( k ) K ( k ) V 2t 2 (r a )2 ar 0 < t < (r - a ) V (r - a ) V < t < (r + a ) V , when, {III.A.18} (r + a ) V < t

and,
0 2 2 6K (l ) 18 (8l , l ) + (6 4 3 ) (20 12 3 )l , l 3 U ( ) = 1 2 8 + (6 + 4 3 ) (20 + 12 3 )l 2 , l 3K ( 1 ) 9 (8, 1 ) (2 3 3) (20 12 3 ), 1 l l l 1 4 2 1 3 (20 + 12 3 ), 1 l + (2 3 + 3) 0< <1 3 1 3 < < 1 . for, 1 <

{III.A.19}

The Rayleigh wave speed, V, is:

V =

2 3+ 3

{III.A.20}

Also:

57

E ( ) =

0 2

1 2 sin2 xdx ,

{III.A.21}

K ( ) =

0
2

dx 1 2 sin2 x

{III.A.22}

(n, k ) =

(1 + n sin )
2 0

d 1 k 2 sin2

{III.A.23}

t , R

{III.A.24}

R = r 2 + a 2 2ar cos ,
3 2 1 l= , 2 V 2t 2 (r a )2 , k= 4ar and,
V 2 (1 V 2 ) 1 V 2 2 = . 2 2 4 2 4 4 3V2 3V2 2V 2 2 2 V 2 (1 V 22 ) 1 V 22 2
2 2 2

{III.A.25}

{III.A.26}

{III.A.27}

( )

{III.A.28}

The coordinate system is: r, is the radial direction, z, is the vertical direction, and, t, is time. The other variables in equations {III.A.17} through {III.A.28} are: P, the impact force, , and, , the compression and shear wave velocities for the half-space, , a lame constant, , a variable of integration, and the area of loading, a. As in Lambs solution, Mitras solution is obtained by solving

58
Equation {III.A.17}, then differentiating numerically with respect to time. Equation {III.A.16} still holds equating the radial dimension derivative and the time derivative. This yields the strain in the radial direction [Goodier, 1959]. In the Mitra [1964] solution, as the wave travels further from the impact point, the relative amplitudes of the three stress waves shift. There is a distinct increase in the amplitudes of the Rayleigh and shear pulses relative to the longitudinal pulses with increasing radial distance as well as an alteration in the way the shear and Rayleigh waves combine (Figure 3.3). This is in addition to the decreasing absolute amplitude and increasing duration, of all three of the waves with time. This solution contains geometry and loading conditions that match closely with the experiments and should therefore provide an accurate theoretical approximation of the impacts.
A.1.3. Estimation of the Energy Dissipated by Elastic Waves

Another quantity that is observed directly from the experiments is the amount of energy absorbed in the impact. This is found by subtracting the remaining kinetic energy during the rebound of the projectile from the initial kinetic energy of the projectile. Many processes of energy dissipation occur within the brittle material during the impact. Elastically, stress waves and vibrations dissipate most of the energy and most of the remaining inelastic energy goes into cracking [Nathenson et al., 2005]. Hunter [1957] provides a method for determining the portion of the impact energy that is dissipated by the stress waves. This provides an estimate of the elastic coefficient of restitution of the impact assuming only stress waves dissipate energy. Hunter concludes that for the Hertzian case of the particle impacting a half-space only a negligible portion of the original kinetic energy of the small body is transferred to the specimen by the collision. That this conclusion is accurate is one reason that the Hertz theory is valid for the quasi-static cases [Hunter, 1957]. The estimate for energy dissipated by the stress waves, W, is found starting from a differential equation for the rate of work done on the specimen:

59

dW du = a 2P (t ) . dt dt

{III.A.29}

Here, a, is the contact radius, P(t), is the pressure on the surface and, u , is the Fourier synthesis of the mean surface displacement. By the derivation in Appendix 3 the energy is found to be: (1 + ) 1 2 2 M. W = o o 3 C 0 1 2
1

{III.A.30}

Hunter specializes this equation using a transient pulse force vs. time function of:
2 M o = m1Zo o ,

{III.A.31}

The energy due to vibrations normalized by the pre-impact kinetic energy is estimated as follows. 2 (4 / 3) 5 1 5g 5v = 1 C 23 m1v 2 2 W
1 2 = (1.068) (1 + 2 ) 1 2
5
2 1 22 1 1 . g = + E2 E1

{III.A.32}

2 16 6 5 15 , 2

{III.A.33}

{III.A.34}

Here, m, is mass, , is density, C, is the longitudinal wave speed, v, is the initial velocity of the projectile, E, is the modulus of elasticity, , is Poissons ratio and,

, and, , are constants related by the listed formula. A subscript 1 indicates the
projectile, while subscript 2 indicates the specimen. For a glass where the

60
Youngs modulus is 70 GPa Hunter estimated, 0.5374 , and, = 1.38 , for, = 1 4 . The coefficient of restitution is found by:
e = 1

W 1 V2 2 m1 o

).
1 2

{III.A.35}

Using equation {III.A.35} the coefficient of restitution, e, was evaluated to be, e = 0.996, at an impact speed of, v = 0.09 m/s [Hunter, 1957]. This value of, e, is slightly higher when compared to an experimental value of 0.985 estimated at the same impact speed [Tillet, 1954]. Increasing the impact velocity from 0.01 m/s to 0.3 m/s decreased the coefficient of restitution by about 0.5 % [Hunter, 1957]. Tillet [1954] observed the same variation over the same range of impact velocities. This equation will be employed to estimate the coefficient of restitution considering only stress waves as an energy dissipation mode.

61
A.2. Previous Research on the Cracking Patterns of Soda Lime Glass

Projects conducted at the Cavendish Laboratory of the University of Cambridge produced several papers concerning experimental impact studies of soda lime glass. In these papers, two of which are from the late 1980s and two of which are from 1977, tests were carried out on a large variety of specimen types. In fact, the paper from 1988 is the final report on an extensive U.S. Army research project [Field, 1988]. The additional studies were carried out by Fields group, by Knight, Swain and Chaudhri [1977], and by Chaudhri and Walley [1978]. These studies all employed spherical projectiles and included soda lime glass as a specimen material.
A.2.1. Configurations and Computations for Previous Experiments

Two main types of specimens were discussed in these studies: glass ceramics and glasses are one type and alumina and boron carbide are the other type. These two categories are distinguished mainly by their youngs moduli, E, and their longitudinal wave speeds, C. The glasses and glass ceramics have, E < 100 GPa and, C < 6500 m/s. The alumina and boron carbide specimens have, E > 250 GPa and, C > 8500 m/s. Soda lime glass, falls into the first class with, E of 79.2 3.0 GPa and C of 5600 100 m/s [Field, 1988]. These studies were conducted on a gas gun capable of using either helium or nitrogen gas to fire a projectile down its 2 meter barrel length with a 13.3 mm bore diameter. The projectile is attached to sabot in configuration similar to that discussed in chapter II. Their sabot is of 25 mm length. A muzzle block stops the sabot allowing the projectile to go through. The gun can either be triggered at low velocities by a solenoid valve, or at higher velocities with a double diaphragm technique. The second technique involves pressurizing the reservoir to the firing pressure and causing rupture by venting the area between the two diaphragms which was pre-pressurized to half the firing pressure. This produces velocities up to 1 km/s [Field, 1988].

62 The specimens were rectangular blocks, and their thickness dimensions were between 3 mm and 35 mm. Impacts were made on the large square face for most of the specimens, but on the largest thickness, some impacts were made on the smaller sides. The papers discuss many salient properties of the material including: density, 2500 kg/m3, hardness, 4.5 0.4 GPa, and fracture toughness 0.75 MN/m3/2 [Field, 1988]. In another study on soda lime glass and Pyrex the impacted soda lime glass specimens were 40 mm x 40 mm x 10 mm and 50 mm x 50 mm x 25 mm [Knight, Swain and Chaudhri, 1977], respectively. In addition to varying the specimen dimensions, multiple projectile geometries were employed. Chaudhri and Walley [1978] used both glass beads and hardened steel spheres of 1 mm diameter. Field [1988] and Field, Sun, and Townsend [1989] used larger projectiles of hardened steel. These were spheres with diameter of 3 mm and 5 mm. This apparatus provided inspiration for parts of the projectile accelerator used in the current study. The three papers focus mainly upon the changes in crack propagation with respect to velocity and projectile size. The range of experimental velocities goes from 100 m/s to over 400 m/s. Key highlights are presented in the following paragraphs. There is little discussion of change in crack pattern with respect to thickness of the specimen in the papers, except when examining the effects of placing another layer of material either in front or behind the glass plates. By the use of an IMACON high speed camera, with an image recorded every 0.95 microseconds, the cracking patterns were well resolved. Also, with the measurement feature of the IMACON, they were able to measure the crack velocity of the fastest internal cracks directly. This crack velocity was, 1550 m/s [Chaudhri and Walley, 1978] and alternatively, 1500 m/s [Field, 1989] and, 1480 m/s 50m/s [Field, 1988]. In one set of experiments [Chaudhri and Walley, 1978] the contact time, that is the time that the spheres were in contact with the glass specimens, was measured

63 from the high speed camera images. A computation similar to that described in equation {III.A.9} was used to calculate estimates, which were compared to the observed contact times. This calculation is given in equations {III.A.41} below. Chaudhri and Walleys [1978] paper also contains theoretical calculations for the maximum loads. These are found by the following equations when the impact is elastic: 5 = 1 3
3/5

Fmax,elastic

4 k 3 E

2 / 5

R 2v 6 / 5 ,

{III.A.36}

telastic

2 2 5 1 1 2 + 1 1 = 2.94 E E1 2 4

2/5

R . v 1/ 5

{III.A.37}

Where,
k=
9 2 2 E (1 2 ) + (1 1 ) 2 . 16 E1

{III.A.38}

In these three equations, 1 is the density; 1 and E1, are the Poissons ratio and elastic modulus for the projectile respectively; R, is the projectile radius; 2 and
E2, are the specimen Poissons ratio and elastic modulus respectively; and v is the

impact velocity. In the case when the projectile deforms, plasticity causes alterations in the contact time and peak force. The total time of contact for the elastic-plastic case can be approximated as the sum of the plastic loading time {III.A.41} and the elastic unloading time. Examination of the specimens impacted by 1 mm diameter steel balls showed no evidence of plastic flow in the specimen. The steel projectile did deform, however, and that is why the governing equations {III.A.36} to {III.A.38} must be modified to incorporate the effects of plasticity of the projectile [Chaudhri and Walley, 1978]:

64

2 Fmax,plastic = a plasticYdynamic ,

{III.A.39}

Ydynamic Ystatic =

H , 3
1/ 2

{III.A.40}

t plastic.loading

m = 2RY 2 1 dynamic

{III.A.41}

In equations {III.A.39} through {III.A.41}, ap, is the radius of the post impact dent on the projectile, Y, is the flow stress of the projectile and, H, is the hardness of the projectile. The theoretical and experimental contact times, quoted from Chaudhri and Walley [1978] are for a steel projectile, were 1.9 s and 1.5 0.3 s, respectively. As these values reflect, the calculated contact times are 0.1 s higher than the range on the experimental contact times. This disagreement indicates that while the theory gives reasonably accurate estimates of the contact time, it is not exact.
A.2.2. Results of Previous Experiments

In all of the experiments on soda lime glass, the same types of cracking patterns were observed. The impact velocity has a large effect on which of these types of cracking patterns are dominant. This holds true for the 1 mm, 3 mm and 5 mm diameter projectiles, with critical speeds for changes between the different cracking patterns depending upon the diameter of the projectile. Cracks begin to be seen at a threshold velocity of about 50-100 m/s. Between the threshold velocity and about 200 m/s conical cracks are dominant with smaller radial and lateral cracks. This threshold velocity varies with projectile diameter. It is about 100 m/s for balls of diameter 1 mm and 50 m/s for balls of diameter 5 mm [Field, 1988; Chaudhri and Walley, 1978]. The effect of kinetic energy is therefore an important parameter in the altering cracking patterns. There is also a

65 hemispherical crushed contact zone under the impact site. As the velocity increases beyond 100-140 m/s the radial and lateral cracks begin to increase in size. Eventually, at velocities between 180 and 200 m/s they dominate the cracking [Field, 1989; Chaudhri and Walley, 1978]. The crushed zone increases in size and material ejects from the surface increasing the severity of the radial cracks. Eventually the cone crack becomes unclear and difficult to see. In one paper it was remarked that with the limitation on frame rate they could no longer see the conical cracks [Chaudhri and Walley, 1978]. Finally, at velocities at and above 250 m/s stress wave behavior becomes dominant [Field, 1988; Field, 1989]. Once the stress waves dominate internal cracking patterns are disordered, no quasi-static cracking feature are visible, and the Rayleigh surface waves produce cracking patterns on the surface [Field, 1988; Field, 1989]. The conical cracks form during the loading of the specimen by the projectile. They initiate at ring cracks just outside of the contact zone [Field, 1988]. The radius of the ring cracks appears constant with changing velocity because the stresses, which generate them, are mainly a function of contact area. The cracks then propagate along a conical path shaped by principal stresses [Field, 1988]. With increasing velocity, the cone cracks angle to the impact axis decreases. This is due to the final contact area of the projectile being larger. If the glass is thin enough, then the crack curves to meet the rear surface at right angles due to interaction with the reflecting waves [Field, 1988]. If there is sufficient thickness, oscillations in the crack size can be seen as it attempts to reach an equilibrium state. This can cause the crack to angle back towards the front face in a skirt [Chaudhri and Walley, 1978]. This type of crack dominates at speeds of about 100 to 180 m/s for the 1 mm steel projectiles [Field, 1989; Chaudhri and Walley, 1978]. For the larger steel projectiles, other studies show that the cones angle to the impact axis decreases with velocity for increasing velocity for similar materials such as Pyrex and fused silica glasses [Field, 1988]. Also, with the larger 5 mm diameter projectiles, the cone produced by the crack ejects from the specimen at speeds as low as 30 m/s [Field, 1989].

66 The other major types of cracks are radial cracks, which propagate on the surface and lateral cracks which propagate initially parallel to the surface, and then curve up to meet it. The radial cracks form during the loading process and, because there is no plastic deformation, are caused by circumferential stresses generated from frictional forces in the surface ring cracks due to the surface deformation [Chaudhri and Walley, 1978]. The lateral cracks form during unloading from tensile stresses at the contact zone interface, which cause small tensile cracks to initiate [Field, 1988]. They can also form from the turn up of the skirt of the conical crack [Chaudhri and Walley, 1978]. Both types of cracks are present from about 100 m/s; they appear at significant sizes in the impact velocity range between 100 and 140 m/s. The smaller the projectile, the larger the impact velocity that is needed to produce the same size crack. This cracking occurs at a speed of 207 m/s for a 5 mm steel projectile that impacted an 8 mm thick specimen. With such a large projectile diameter to plate thickness ratio, the radial cracking was caused by plate bending [Field, 1989]. Another type of cracking is observed for smaller projectiles. Knight et al. [1977] observed the impact of 0.8 mm and 1 mm diameter steel spheres on soda lime glass and Pyrex at speeds between 30 and 300 m/s. The soda lime glass impacts showed cracking patterns similar to that of a quasi-static indentation of a point load. Point indentations, create a Boussinesq stress field. Instead of just the cone crack, a group of splinter cracks beneath the indentation point is observed to initiate. These cracks grow normal to the principal tensile stress direction [Knight, Swain and Chaudhri, 1977]. The cone crack was observed in some cases, starting to grow before the splinter cracks. Crushing of the contact area or plastic deformation at the projectile are suggested as possible causes of the splinter cracks. These cracks propagate at the maximum crack velocity for soda lime glass ~ 1500 m/s [Knight, Swain and Chaudhri, 1977]. During the unloading process, the cracks closest to the surface can keep growing and bend towards the impact face in the same manner as lateral cracks. Other cracks closer to the impact axis continue to grow towards

67 the rear surface for a distance on the order of millimeters. However, with increasing velocity more of the splinter cracks alter direction. The unloading cracks travel at the much slower speed of about 300m/s [Knight, Swain and Chaudhri, 1977]. Additionally in this study, measurements of the strain produced by the stress waves on the surface of the specimens occurred. These strain records indicated that dynamic cracking is reflected in changes to the strain records. In this case, compressive strain pulses were associated with well developed conical cracks [Knight, Swain and Chaudhri, 1977]. In the present study, the relationship between cracking patterns and the stress waves in the material, which in turn alter the surface strains, is further explored. At speeds of about 250 m/s, for the 3 mm and 5 mm diameter steel projectiles, stress waves begin to dominate [Field, 1988; Field, 1989]. The internal fractures become very chaotic and the surface cracks are caused by Rayleigh surface waves. Back surface damage is caused by the reflection of the compressive loading wave as a tensile wave from the back surface [Field, 1989]. A specimen impacted by a 5 mm diameter projectile at 402 m/s shows short circumferential bands of surface cracks. These cracks are the result of reinforcement of the Rayleigh surface wave and reflected tensile waves from the back surface. This is a different category of cracking than at lower speeds where the same patterns that exist in quasi-static indentation testing are observed [Field, 1988]. The current study examines these impact velocities, using 1 mm diameter steel projectiles. To summarize, the cracking patterns of the specimens indicate a distinct dependence on both impact velocity and specimen thickness. In research conducted by Chaudhri and Walley [1978], the specimens impacted by 1 mm diameter projectiles exhibit dominant Hertzian cone cracks for velocities up to about 180 m/s. Studies by Field [1988] showed that conical cracks are seen to initiate at front surface ring cracks and then propagate along a path shaped by the principle stresses. At velocities above 100 to 120 m/s radial and lateral

68 cracking formed and grew more dominant with increasing velocity in impacts of 1 mm balls [Chaudhri, 1978]. Front surface radial cracks formed during the loading and unloading process due to tensile surface circumferential stresses [Chaudhri, 1978]. Rear surface radial cracking was seen to be caused by the reflecting longitudinal stress wave occurred in specimens impacted with 3 mm balls [Field, 1989]. The lateral cracks formed during unloading from tensile stresses on the contact zone interface from small tensile cracks [Field, 1988]. Front surface chipping represents either lateral cracks that have ejected from the front surface or, at higher velocities, damage due to the stress waves [Field, 1989]. In cases of small balls impacting, 0.8-1 mm diameter, instead of just the cone crack, a set of splinter cracks beneath a crushed indentation zone were seen to initiate in a study by Knight et al. [1977]. These cracks initiated at tensile stresses produced by the crushing of the contact zone and grew normal to the principal tensile stress direction. A cone crack was observed in some cases, starting to grow before the splinter cracks [Knight, 1977]. At high velocities, the larger 3 mm and 5 mm projectiles produced chaotic internal damage and different forms of surface damage [Field, 1988; Field, 1989]. The current study expands upon these observations and studies the relationship between impact velocity and thickness on the cracking patterns.

69
B. Design of Soda Lime Glass Particle Experiments

Within the soda lime glass phase of the project, experiments were designed to vary the thickness of the specimen and the velocity of the impact. Four velocities and four plates of varying thickness were studied in this phase. The velocities in the experiments are designated in four categories: low or around 150 m/s, medium low or around 230 m/s, medium high or around 300 m/s, and high or around 350 m/s. The glass targets are 3 mm, 5 mm, 15 mm and 25.4 mm thick. Also factored into the experimental matrix are the methods of measurement. The experimental apparatus is that discussed in chapter II. A preliminary set of tests provided the loading pressures necessary to achieve the desired impact velocities. Two sets of measurement setups were employed. The basic quantities configuration (BQ) and supplemental experiments (SP) contained all but two of the experiments. The bending configuration (BN) measures the rear surface vibrations along orthogonal directions. The basic quantities (BQ) set of experiments is designed to measure several quantities. The contact time, the impact velocity, the rebound velocity, and the general cracking pattern are determined from the IMACON high speed cameras. To ensure that the entire impact and rebound is recorded, a low framing rate of 150,000 frames per second is used. The strain gages on the surface measure the dynamic wave patterns of the radial and hoop strains. These strain records can be compared with numerical and theoretical strain wave estimates. Supplemental experiments (SP) were performed with the IMACON 200 camera using a framing rate of 400,000 frames per second. The greater framing rate was used for greater accuracy of velocity measurement. The strain gages were also used in the supplemental experiments. The second series of experiments (BN) observed vibrations on the rear surface. Strains in two orthogonal directions were observed. This configuration in called bending because the plate flexes due to these strains. Since the strain gages in this set are located on the rear surface along the axis created by the impact of the projectile, the vibrations of the rear surface can be measured. The same

70 camera system is used as in the BQ experiments to obtain impact and rebound velocities as well as general cracking patterns. The vibrations in the strains on the back surface represent the dynamic flexing of the entire specimen. By using high-speed data collection the oscillations in the strain can be measured. Two oscilloscopes collect data; one with a 50 microsecond recording window and one with a 500 microsecond window. This is to gain a clear picture of the dynamic vibration history of the specimen. These two experiments were carried out after the BQ experiments and were targeted at specific velocity and thickness combinations that would produce the best data on dynamic specimen flexing. Each experiment is designated by the order in which it was performed and by a configuration type. For example BQ-12 indicates the twelfth experiment performed in the set. Most experiments were conducted using the basic quantities setup. The bending setup was used only to measure in two experiments. These are designated BN-01 and BN-02 respectively. Table 3.1 contains the set of experiments sorted primarily by type, impact velocity, and finally by specimen thickness. The experiments are divided up in the table by two different parameters because each different experimental setup has experiments with varying impact velocities and thickness. For the basic quantities, four different velocity ranges are explored for each of the four thickness plates. These different velocities represent various regimes of behavior in the previously studied soda lime glass experiments. The four thickness plates represent a variation from thin panels to near half space thickness. Some experiments were repeated due to either clarification issues or unclear measurements. This design of experiments provided the ability to observe the transient and equilibrium behavior of soda lime glass under various impact conditions.

71
C. Results and Analysis of the Soda Lime Glass Experiments

In the present study, several series of experiments were performed. The first, series involved a set of experiments to determine the pressures required to obtain the desired projectile velocities. The basic quantities (BQ) set of experiments was the second series conducted. Following this was the third experimental group, the bending (BN) experiments. Finally, the supplemental experiments (SP) were performed. From the collected data, various impact properties of the material can be deduced. These properties can be divided into four categories. First, the contact times, which can be inferred from high speed camera images, are compared with the theoretical Hertzian contact times and the numerical contact times from LS-DYNA [Nathenson et al., 2005]. Second, the realm of the stress waves and material behavior before equilibrium is reached is examined by means of the strain gage signals. These are compared with the theoretical estimates from Lamb and Mitra, as well as strain waves generated by the LSDYNA numerical studies [Nathenson et al., 2005]. Also from the bending experiments, the vibration of the specimens is analyzed. Third, the kinetics of the impact, the elastic energy absorbed by the specimen and the coefficient of restitution of the glass are discussed. Fourth, the records of the cracking patterns are examined with respect to changes in the thickness and the impact velocity. By examining these areas a comprehensive picture of soda lime glass behavior under high velocity particle impact is assembled.
C.1. Projectile Velocity vs. Firing Chamber Pressure Calibration

Before the projectile accelerator could be used for impact studies, the loading pressure to firing velocity relationship had to be determined. The projectile was accelerated using various loading pressures and removed from its sabot. The velocity of the projectiles were determined using the high-speed camera images. The position of each projectile in succeeding frames was used along with the time between the frames to estimate the velocity. In order to measure the sabot velocity following its exit from the gun barrel, the laser system was used. The sabot velocity is determined by placing two laser beams a fixed distance apart

72 across the path of the sabot and using a diode to measure the time of the intensity drop caused by the passage of the sabot though each beam. The high speed camera is triggered by the laser velocity system when the first beam is passed. This laser then triggers the flash and the ULTRA 17 camera that is set to take exposures of ten nanoseconds duration at the frame rate of 150,000 frames per second. This camera takes seventeen pictures, which is more than sufficient to establish the velocity. The first exposure begins after a delay determined by the distance between the laser velocity system and the specimen surface, allowing for the recording of images to commence when the projectile passes in range of the camera. In order to insure that the camera accurately measures the distance moved by the projectile between two frames, an object of known length in one of the frames is measured. This object was either the width of the front prong of the specimen or the projectile diameter. The distance traveled by the projectile, and the interframe time allows the cameras dedicated computer to determine the velocity by selecting one point in space at two different times that is, two separate frames. This method yielded the curve for sabot and projectile velocity versus firing pressure. Measurements were taken at pressures of 100, 150, 200, 250, 300, 350 and 400 pounds per square inch. As shown, velocities as high as 311 meters per second were achieved (Figure 3.4). By curve fitting it was determined that the equation with the greatest R2 value for firing pressure in terms of desired velocity is a polynomial function. This relationship is specified by the following equation: Pressure = 0.0019 Velocity2.1209 . {III.C.1}

This equation has a R2 value of 0.871. With this equation, the necessary pressure for a given projectile velocity can be determined. The full set of experiments is also shown in Figure 3.4.

73
C.2. Time of Contact Between the Sphere and the Plate

The contact times were obtained from the experiments using the camera, from the theoretical Hertzian approximation [Goldsmith, 2001], and the numerical study using LS-DYNA [Nathenson et al., 2005]. The accuracy of the camera contact times was limited accurate due to the 6.6 microsecond or 2.5 microsecond inter-frame time and debris obscuring the point at which the ball left the surface. Accepting this limitation, the contact time estimates from the camera are contrasted with those estimates using the theoretical Hertzian approximation and the LS-DYNA numerical model. All of the contact time values are tabulated in Table 3.2. The Hertzian approximation [Goldsmith, 2001] shows a decrease in contact time from 2.61 s at 150 m/s to 2.20 s at 350 m/s (Figure 3.5). This trend of decreasing contact time is also reflected in the LS-DYNA numerical modeling, where the contact times are estimated to decrease from 2.61 s at 150 m/s to 2.16 s at 350 m/s. According to the LS-DYNA results at higher impact velocities, the 3 mm thick plates have slightly smaller contact times, on the order of 0.05 microseconds when compared with the other three thickness plates [Nathenson et al., 2005]. The three thicker specimens have nearly identical contact times according to the numerical model. Specifically, the simulation at 150 m/s in the 5 mm plate was off by 0.01 s from the simulation of the 15 mm and 25.4 mm plates. The other velocity simulations matched exactly for all three thicker plates. Including the 3 mm plates, the numerical model contact times are all within 3% of the theoretical Hertzian contact times. Both the model and Hertzian theory are elastic, with no cracking so this congruence is expected. This close agreement confirms that the numerical model is accurately predicting the theoretical behavior. In examining the camera records, it is evident that the contact times can be estimated to within an accuracy of 6 s, as this is the inter-frame time for the ULTRA 17, and to within 2.5 s for the IMACON 200. The time at which the specimen leaves the surface is uncertain in several of the experimental images due

74 to debris ejecting from the surface. This makes identifying contact time to be uncertain to within 1-2 frames. However, in most of the images, the contact times can be seen to be within one frame. The resulting contact times for the experiments where the IMACON 200 is employed are between 0 and 2.5 s (Figure 3.5). This estimate fits well with the Hertzian estimate of contact times between 2.61 s and 2.20 s. The literature also shows close but not perfect agreement, within 0.4 s, between theoretically and experimentally estimated contact times [Chaudhri and Walley, 1978]. In summary, close agreement is found between the contact times estimated from the experimental camera images and the two models. The theoretical Hertzian equation {III.A.9} and the LS-DYNA numerical model contact time data (Table 3.2) indicate agreement within 0.1 s, or 3 %, at each velocity for all of the thickness plates. The experimental estimate of 1 frame, or 2.5 s falls within the range of both the theoretical and numerical estimates. This agreement of the contact time estimates confirms that the Hertzian theory is applicable to impacts of hardened chrome steel on soda lime glass and that the LS-DYNA numerical model is accurate.
C.3. Examination of the Impact Surface Strain Gage Records

Using stacked tee rosette strain gages from Vishay Measurements Group, the dynamic surface strain wave profiles were captured. Specifically, the specimens the radial and hoop strains are measured. These experimental results show distinct variations with specimen thickness, with the same trend for each velocity. The resulting records also show similar characteristics to the LS-DYNA numerical and the theoretical results from Lamb and Mitra.
C.3.1. Strain Wave Components Explained Using Theory and Numerical Methods

In the numerical study performed by Guodong Chen [Nathenson et al., 2005] in conjunction with the current set of experiments, the material stresses and deformations were examined using LS-YDNA 3D. In this study, the effects of a

75 steel ball hitting three different cases of soda lime glass plates were studied. The response of a half space, plates of finite thickness but infinite lateral size, and rectangular geometry that corresponded to the experimental specimens were examined. The strains in the LS-DYNA model were measured at a point 10 mm from the impact location in order to simulate the results observed in the experimental study. Lambs [1904] and Mitras [1964] theoretical models were also solved for the 10 mm radial distance and are compared to the half space LSDYNA simulation. In the LS-DYNA numerical study on the half space the structure of the radial wave is that of a pulse first in tension, then compression. Smaller reverberations follow this large pulse (Figure 3.6). The LS-DYNA simulation includes all three of the stress waves: the longitudinal wave, the shear wave, and the Rayleigh wave. The strains generated from the equations by Lamb [1904] (Figure 3.2) and Mitra [1964] (Figure 3.3) are both compatible in form with the LS-DYNA solution at 350 m/s impact velocity. Lambs solution specifies that the Rayleigh wave is dominant. The solution in Figure 3.2 has the same general form as the LSDYNA solution for the half space in Figure 3.6. The initial compressive strain pulse is followed by a tensile strain pulse of greater amplitude and then a compressive strain pulse. Mitras [1964] solution includes the longitudinal and shear waves. Comparing Mitras solution (Figure 3.3) to the approximation to the LS-DYNA study at 350 m/s (Figure 3.6), shows a similar initial pulse, this time attributed to the longitudinal wave. The pulses maximum strain is normalized to the largest maximum experimental strain in the plot. The plot of Mitras solution at this radial distance contains a large longitudinal pulse and much smaller shear and Rayleigh wave pulses, which is consistent with the experimental strain records.

76

Both theories yield solutions that are of similar form to the LS-DYNA simulation. However, Mitras solution (Figure 3.3) is based on more realistic assumptions and loading conditions. Mitras solution also appears more similar to the numeric results for the half space (Figure 3.6). The initial compressive pulse is smaller than that of Lambs (Figure 3.2). The LS-DYNA strain arrives at the time for the longitudinal wave, not the Rayleigh wave. However, some differences are visible from the records. For example, the fact that the Mitra solution includes an impulse forcing function with zero duration and infinite magnitude, and the numeric solutions do not. The greater duration and lower absolute magnitude of the experimental and numeric pulses are evidence of this difference. For the case of finite thickness, variations exist within the strain records. The changes in the strain occur around the time of the arrival of the reflected waves from the rear surface or following their arrival (Figure 3.7). In the simulation of the 3 mm and 5 mm specimens the tensile portion of the strain pulse is increased, while the compressive part disappears. This is not present in the 15 mm and 25.4 mm thick specimens, where the reflected wave does not return until after the pulse has passed. In all cases, the reverberations following the pulse were altered somewhat due to the passing of the reflected wave. In the case of a finite rectangular geometry, some further variations in the reverberations following the pulse were measured (Figure 3.8). Because the wave reflecting from the lateral boundary did not return until the main strain pulse passed the simulated recording location, no major effect was observed on the surface strain. Minor deviations are visible in the reverberations following the passage of this wave. The same thickness effect was observed as in the simulations with infinite boundaries.

77
C.3.2. Description of Experimental Strain Records

All of the experimental radial strain records showed one distinct spike and a smaller spike in the hoop strain record. The highest velocity experiments show the largest amplitude strains. Looking at one experiment from each thickness at the 350 m/s range (Figure 3.9) it can be seen that as the thickness increases from 3 mm to 15 mm, the peak amplitude decreases. This is due to the fact that the thicker materials more closely approximate the half space assumption of the calculations. The 15 mm and 25.4 mm thick plates have very similar strain profiles. The radial profiles show a tensile spike in the case of the thicker specimens, 15 mm and 25.4 mm, that quickly returns to a near zero level with some reverberations. In the thinner two specimens, this tensile pulse is of much longer duration. Before these large tensile pulses, there is always a small compressive pulse. These effects of thickness are also seen in the numerical and theoretical work. In the 3 mm experiments, at all impact velocities, the strain pulses are observed in the radial direction to have an average duration of 34 microseconds (Table 3.3). The smallest velocity (BQ-18) has a duration of 44 microseconds, while the 300 m/s (SP-1) has a duration of 27 microseconds, and the maximum velocity has a duration of 33 microseconds. The pulse magnitude increases from 67 micro-strains at 148 m/s (BQ-18) to 151 micro-strains at 300 m/s (SP-1). However, the magnitude decreases to 93 micro-strains at 345 m/s (BQ-26). In the 5 mm thick specimens, the strain pulses have an average duration of 13 microseconds (Table 3.3). At 174 m/s (SP-8) the duration is 6.2 microseconds. This increases to 26 microseconds at 323 m/s (BQ-13). The average magnitude is 49 micro-strains. At 174 m/s (SP-8) the amplitude is 46 micro-strains and this increases to 57 micro-strains at 299 m/s (SP-9) before returning to 48 microstrains at 323 m/s (BQ-13).

78 Pulses of shorter duration occur in the 15 mm thick specimens (Table 3.3). Here the average duration is 3.28 microseconds and the average amplitude is 26 microstrains. The shortest pulse occurs at 164 m/s (SP-6) and is 1.6 microseconds long. The longest pulse is 6.2 microseconds long and occurs at 370 m/s (BQ-25). The maximum strain of 37 micro-strains occurs at 164 m/s (SP-6). The minimum strain is 20 micro-strains and is at 231 m/s (SP-13). Similar behavior is observed in the 25.4 mm thick specimens where the average duration is 4.5 microseconds (Table 3.3). The corresponding average peak strain is 25 micro-strains. The smallest pulse is 2.6 microseconds long in the 287 m/s experiment (SP-11). The largest pulse happens in the 336 m/s experiment (BQ11) and is 9 microseconds long. The highest peak strain is 29 micro-strains and occurs in the 287 m/s experiment (SP-11). The smallest peak strain occurs at 336 m/s and is 21 micro-strains. There is considerable variation in the amplitude and pulse duration recorded by the strain gages. However, certain conclusions can be drawn from the data. The average amplitude of the stress pulses decrease with thickness from 104 microstrains to 25 micro-strains when the thickness is increased from 3 mm to 25.4 mm (Figure 3.10a). The average duration of the stress pulses also decreases over the same thickness change from 34 microseconds to 4.5 microseconds (Figure 3.10b). The strain amplitude shows an increasing trend over the velocity range, but with wide variation (Figure 3.10c). The duration of the pulse shows no trend with increasing velocity (Figure 3.10d). Therefore, varying the specimen thickness correlates with a quantifiable effect on both the strain magnitude and pulse duration while varying the impact velocity does not. In light of the LS-DYNA numerical study, certain conclusions can be drawn from the experimental strain gage data. In all of the experiments, the strain profile is that of a small compression dip followed by a large tensile pulse. This is in agreement with the LS-DYNA simulations. Unlike the simulations, there is no rebound into compression in the experiments, suggesting that the materials

79 fracture strength is low enough that cracking occurs, preventing compressive strain from occurring on the surface. The low fracture strength of the specimen is also responsible for the fact that the numeric peak strains are an order of magnitude greater than that of the experiments. Additionally, in the LS-DYNA simulations at 350 m/s, the tensile strain pulse with the 3 mm plate is seen to have twice the duration as in the 25.4 mm plate (Figure 3.8). The experimental tensile strain pulse in the 3 mm specimen has 4.5 times the duration of the strain pulse in the 25.4 mm experiment for the same velocity range (Figure 3.9). This additional distortion of the duration of the strain pulse in the experiment is also due to cracking.
C.3.3. Correlation Between Rear Surface Vibration and Crack Oscillation

Measuring the strains of the rear surface of the specimen with the strain gages allows for observation of dynamic oscillations in the strain gage profiles. The oscillation in the strain profiles reflect the dynamic flexing or bending of the plate. The measurements were conducted on two specimens; one of which was 5 mm thick and the other was 15 mm thick. These impacts occurred at speeds of 284 m/s and 290 m/s, respectively. These are designated by experiment numbers BN-1 and BN-2. The experiments were conducted at these speeds in order to examine the regime of velocity and thickness where the cracking patterns are most evident. That is, experiments in which the largest size cone cracks were observed to form and oscillate. There is a correlation between the oscillations in the strain histories and the cracking pattern oscillations observed using the high speed camera. The strain gage histories indicate oscillations, always in the tensile range during the first 50 s of the experiment. Both the 5 mm (BN-1) and 15 mm (BN-2) thick strains have several wave oscillations during this time. The 5 mm experiment (BN-1) has longer and more irregular oscillations of about 12-18 s (Figure 3.11). The overall crack in the 5 mm experiment (BN-1) is a crushed zone with many small splinter cracks. These splinter cracks appear to oscillate with a frequency of 6-12 s. The 15 mm experiment (BN-2) has shorter oscillations of about 6-12 s

80 (Figure 3.12). The 15 mm experiment (BN-2) shows a crushed zone and a partially formed cone crack. This cracks edges distinctly show oscillations of 1224 s. In the camera images, the cracking patterns in these two experiments are somewhat less severe than those in the BQ series of experiments of approximately the same velocities. This may be due to the strain gage glued to the rear surfaces. The added strength imparted by the glue may reduce stresses and retard crack growth. The cracking patterns and the strain records in both experiments show oscillations. In both BN-01 and BN-02, the cracking patterns oscillate at the same period that the strain gage records vibrate, within the experimental error. It is apparent from the geometry of the specimens that cracks, which are within the specimen, should not be subject to exactly the same stresses as the rear surface. However, given the thin nature of the specimens, any oscillatory behavior of the stresses and hence the strains should be similar. In fact, this is seen in the experiments. Therefore, it is likely that the dynamic vibration caused by the reflection of stress waves from the rear surfaces is connected to the oscillations in the cracks.

81

C.4. Impact Energies and Coefficients of Restitution

The impact energy imparted to the specimen is dissipated by elastic and inelastic processes. A certain portion of this impact energy is dispersed through elastic processes such as stress waves and vibrations, while most of the remaining energy either goes into crack generation or the rebound of the projectile. This section discusses quantifying the relationships between these different forms of energy. First, the experimental method for determining the impact energy and the corresponding coefficient of restitution is explained. Next, these quantities are compared with the elastic energy lost and resulting coefficient of restitution estimated from the numeric simulations [Nathenson et al., 2005]. The energy dissipated due to stress waves is also estimated using theory [Hunter, 1957].
C.4.1. Determination of Experimental Impact Energy and Coefficient of Restitution

The impact velocity was determined in two ways. First, the laser velocity system measured the sabot velocity by determining the time of travel between two laser beams set a fixed 5.9 mm apart. These beams produce a set of velocities before the sabot stripper. Error in this measurement comes from both the exact spacing of the laser beams, and the definition of the drops in the pulse measurement (Figure 3.13). The measurements of the pulses are taken from the beginning, at the end, and at the center of the first drop to the same point on the second drop. The second velocity measurement comes from the high speed camera. By measuring the distance traveled by the projectile between two of the camera frames, a fixed time interval, the velocity is measured. This measurement is based on a calibrated distance, specifically, the projectiles diameter. This is known to be 1.588 mm for the hardened chrome steel ball bearings. In order to improve accuracy, four sets of two frames are used. These measurements are averaged in order to compute the impact velocity.

82 The rebound velocity is computed in the same manner as the camera computation of the impact velocity. In this case, debris from the impacts of the specimens can result in obscuring of the rebounding projectile. Thus, usually fewer sets of frames are used in this computation. Because the camera image is two dimensional, any sideways motion of the projectile is not measurable. However, the measurement of rebound velocity is still reasonably accurate because any motion perpendicular to the axis of impact is minimal given proper alignment of the specimen. The laser impact velocity, camera impact velocity, and rebound velocity are all measured from each experiment. The actual impact velocities indicate a spread around the intended velocities (Figure 3.4), but this spread is acceptable as the velocities still resolve into four distinct categories. By measuring the average impact and rebound velocities and then comparing them, the kinetic energy that is absorbed by the specimen, K .E . , can be calculated. This is done by means of the following equation: 1 2 m1 (v 2 vR ) . 2

K .E . =

{III.C.2}

In equation {III.C.2}, m1, is the mass of the projectile, v, is the impact velocity and, vR, is the rebound velocity. This equation is used because the specimen can be treated as not moving for the duration of the contact. Also, the rebound energy can be normalized to the initial impact energy:
2 K .E .rebound vR = 2. K .E .impact v

{III.C.3}

The projectiles rebound kinetic energy decreases with increasing impact velocity (Figure 3.14). This indicates that as the impact velocity increases, a larger percentage of the impact energy is absorbed by the specimen. The ratio of rebound kinetic energy to impact kinetic energy corresponds to a coefficient of restitution which, is expressed in the following formula:

83

eE =

vR

v.

{III.C.4}

The coefficient of restitution is symbolized by, eE. These values are tabulated (Table 3.4). With increasing impact velocity, the coefficients of restitution decreases (Figure 3.15). The thickness of the impacted specimen has no effect on the coefficient of restitution.

84
C.4.2. Elastic Stress Wave and Vibrations as Portions of the Impact Energy

As discussed above, the portion of the impact energy absorbed by the specimen manifests in elastic deformation and cracking processes. The dissipation of elastic energy occurs in two main modes: stress waves and vibrations. The amount of energy that goes into the stress waves can be estimated by means of the analysis presenter by Hunter [1957]. In equation {III.A.32} this relationship is given. The resultant values are plotted in Figure 3.14. The corresponding coefficient of restitution is given in equation {III.A.35} and these values are plotted in Figure 3.15. Furthermore, results of the numerical simulations conducted using LSDYNA enable the estimation of the change in kinetic energy and the corresponding coefficient of restitution. However, no cracking is considered in the numerical model, and thus these estimates only include the effects of elastic processes. These two estimates and the experimental values for the coefficients and change in kinetic energies are also tabulated in Table 3.4 and plotted in Figures 3.14, and 3.15. The coefficient of restitution, based on the measurements recorded via the highspeed camera, show a drop over the four different thickness plates from 0.50 at 150 m/s to 0.19 at 350 m/s (Table 3.4). Hunters [1957] estimates for the stress wave coefficient of restitution show a drop from an average over the thickness of 0.94 at 150 m/s to 0.89 at 350 m/s (Table 3.4). The LS-DYNA estimates for the elastic coefficient of restitution also exhibit some thickness dependence (Figure 3.15). The three thicker specimens have near identical coefficients as they decrease from 0.89 at 150 m/s to 0.85 at 350 m/s (Table 3.4). The maximum difference between the estimates for the three thicker specimens at the same velocity is 0.01. The 3 mm specimens decrease from 0.81 to 0.77 as the velocity increases over the same range (Table 3.4). No such correlation between thickness and the experimental coefficient of restitution is visible due to scatter in the experimental data.

85 Similar patterns are observed for the particles kinetic energy change due to the impact (Figure 3.14). For the estimated coefficient of restitution using Hunters [1957] equations, the rebound kinetic energy drops from 87 % of the input energy at 150 m/s to 78 % at 350 m/s. The LS-DYNA simulations for 5 mm, 15 mm, and 25.4 mm exhibit similar decreases in rebound kinetic energy over the same impact velocity range from an average of 80 % to an average of 72 %. The deviation among these values is 1 %. The numeric estimate of the rebound energy is 66 % at 150 m/s and 59 % at 350 m/s in the 3 mm thick specimens. There is a larger decrease in the rebound kinetic energy in the experiments. At impact velocities around 150 m/s the average rebound kinetic energy is 25 %. When the impact velocity is about 350 m/s, the average rebound energy is 3.8 %. The absolute rebound energy is smaller in absolute terms, and the rate of decrease in the rebound energy with increasing impact velocity is much greater than that for the LS-DYNA numeric simulation or Hunters [1957] stress wave computation (Table 3.4). The coefficient of restitution and percent rebound energy both decrease with increasing velocity in the experiments. Additionally, most of the impact energy is unaccounted for by elastic processes. The ratio of the inelastic energy to the elastic energy increases with increasing velocity. Glass is a brittle material, and almost no plasticity occurs during the impact. Therefore it can be assumed that the inelastic energy is applied to the crushing of the contact zone and the various cracking patterns. These cracking patterns are discussed in the following section.
C.5. Cracking Pattern Behavior

Cracking patterns were examined both in situ and post test. According to high speed camera observations the crack propagation speed is about 1500 m/s [Field, Sun, and Townsend, 1989]. This is much slower than the stress waves traveling through the material. The longitudinal wave speed is 5600 100 m/s [Field, 1988]. As a result the stress waves themselves may alter the stress distributions within the material, and thus cause alterations in the cracking patterns. The

86 sections below describe the observations of the cracking patterns and the variations due to thickness and impact velocity.
C.5.1. Details of Cracking Following Experiments

Following each experiment an investigation of the cracking patterns was carried out. The cracking patterns were found to change significantly with both impact velocity and thickness. The variations in these patterns with thickness and velocity are detailed for representative experiments in Table 3.5. For specimens impacted below about 130 m/s, either no damage occurred at the impact site, or only a small circular indentation of 1 mm to 3 mm diameter appeared. This indentation size is of the same order as the 1.588 mm diameter projectiles. At speeds of about 130-150 m/s a crushed zone appeared in all specimens. From this, it can be argued that the threshold velocity of crack initiation for this geometry of projectile and these impact velocity ranges is therefore 140 10 m/s. More severe forms of damage accumulated with velocity. The resultant cracking patterns are described below. The experiment with a 3 mm thick plate, at 149 m/s impact velocity (BQ-18) resulted in a 2 mm diameter, 1 mm deep crushed zone. 3 mm diameter chipping occurred on the front surface. According to previous studies, a series of ring cracks should appear outside of the crushed zone [Field, 1988]. In the current experiments, surface chipping usually, but not always occurred on the front surface obscuring the area where the ring cracks should be located. This made determination of the ring crack diameters impossible. In BQ-18, three short radial cracks appeared. For the purposes of this discussion, the definition of a short radial crack is that it has not propagated completely through the specimen, while the long radial cracks have propagated completely through the specimen. Also, a partial cone crack with maximum diameter of 9 mm appeared. At higher impact velocities, the variation in cracking patterns due to thickness became more apparent. For the 3 mm specimens, the impacts produced radial cracks of increasing number and severity as the impact velocity increased. A

87 Hertzian cone was also ejected from the rear surface. At 238 m/s (BQ-3), in addition to ring cracks and surface chipping of maximum diameter 9 mm, there were four long and three short radial cracks. The ejected cone crack had a diameter of 10 mm. At 292 m/s (BQ-9) 8 long radial cracks and 4 short radial cracks were present. Again, there were ring cracks and surface chipping this time 8mm in diameter. The cone that is ejected had a 10 mm diameter. At 300 m/s (SP-1) the cone crack had a 6 mm base diameter. Six large and 2 small radial cracks were observed. The 345 m/s experiment (BQ-26) resulted in 4 small and 4 large radial cracks. There was no observed surface ring cracks or chipping. The ejected cone had a diameter of 11 mm. The 5 mm specimen produced a crushed zone 2 mm in diameter and 1 mm deep when impacted at 155 m/s (BQ-14) and shallow ring cracks and chipping of 6 mm diameter. A lobe of partial lateral cracking appeared, spreading from the crushed zone. A maximum radius of 2.5 mm from the axis of impact was observed. Also, 1 mm long splinter cracks were observed extending from the base of the crushed zone. These splinter cracks extended perpendicular from the surface of the crushed zone. They formed at or near the bottom of the crushed zone. At an impact velocity of 174 m/s (SP-8) a crushed zone of 1 mm diameter, an 8 mm diameter lateral crack and 4 mm surface chipping formed. Also splinter cracking was observed with a maximum length of 1.5 mm. Some of the smaller splinter cracks curved towards the surface of the specimen, while the largest, 1.5 mm, splinter cracks remained perpendicular to the crushed zone. At 234 m/s impact velocity (BQ-8), the crushed zone deepened to 1.5 mm while the diameter remained 2 mm. Ring cracking was not observed, but some chipping of 9 mm diameter did occur from the front surface. One short radial crack and partial lateral cracking formed. The lateral crack is considered partial because it did not completely surround the crushed zone. Its largest radius was 4.5 mm from the impact point. Splinter cracks of 1 mm length were also seen.

88 In BQ-7, at 300 m/s impact velocity, the crushed zone increased in diameter to 2.5 mm and depth to 2.5 mm. Also, 5 mm diameter surface chipping was seen. Four short radial cracks and a cone crack with maximum diameter of 8 mm were visible in the 5 mm thick specimen, but the cone crack did not eject. No lateral cracking or splinter cracks were observed. At 295 m/s (SP-9) a crushed zone of 2.5 mm diameter was also observed with 12 mm diameter surface chipping. Lateral cracks of 15 mm diameter were seen and 3 small radial cracks initiated at the rear surface. At an impact velocity of 311 m/s, the crushed zone was observed to be 2.5 mm in diameter and 1 mm deep. A partial cone crack 6 mm in diameter formed but did not eject. Two large radial cracks also initiated from the rear surface. At 323 m/s (BQ-13), the 5 mm specimen exhibited a crushed zone of diameter 2 mm and depth 2 mm, large partial shallow ring cracks, and chipping from the front surface of 16.5 mm diameter. Also visible were one large radial crack, two cone cracks, one full cone crack but not ejected of diameter 8 mm and one partial cone crack. Partial lateral cracking of 5 mm radius could be seen. No splinter cracks were visible. Different damage patterns occurred in the 15 mm specimens in the same velocity range than in the 3 mm and 5 mm thick specimens. At 137 m/s impact velocity (BQ-19) there was a crushed zone of diameter 3 mm and depth 1 mm. The surface ring cracks and front surface chipping extended to 5 mm diameter. No radial, conical, or lateral cracks were seen. There were splinter cracks of length 1 mm. At 164 m/s (SP-6), a crushed zone of 1.5 mm diameter and 1 mm deep was observed. Surface chipping of 9 mm diameter occurred. Also 1 mm deep splinter cracks were visible. At 230 m/s impact velocity (BQ-6) the crushed zone was 2.5 mm in diameter and 1.5 mm deep. Material was ejected from the front surface and there was shallow surface ring cracking and chipping of 8 mm diameter. There were again splinter cracks of length 1 mm but no radial, cone or lateral cracks. Another

89 experiment at the velocity of 231 m/s (SP-13) resulted in a 1 mm diameter crushed zone. A 9 mm completely circular lateral crack was observed along with 7 mm radius partial surface chipping. Splinter cracks of 1 mm length also occurred. At 239 m/s (SP-5) a crushed zone that was 2.5 mm diameter and 1.5 mm deep was seen. Some surface chipping occurred with a 3 mm radius. Five partial lateral cracks initiated but only propagated 1 mm. Also, 1 mm deep splinter cracking was observed. In the 15 mm specimens at higher impact velocities, more severe cracking is observed. At 279 m/s (SP-7) a 1.5 mm diameter crushed zone is generated by the impact. A lateral crack results in surface chipping of 12 mm diameter. At 300 m/s (SP-4) a crushed zone of 2 mm diameter is observed. Surface chipping of 14 mm diameter occurs. Additionally partial lateral cracks of 5 mm and 10 mm radius are seen. At 307 m/s (BQ-5) the crushed zone was 3 mm in diameter by 1 mm deep. The surface cracking was partial but 10.5 mm in diameter. No radial or cone cracks were observed, but a complete lateral crack of diameter 13 mm that surrounded the crushed zone and traveled in a saucer like fashion towards the surface was seen. Splinter cracks of length 0.5 mm were also seen. At 371 m/s (BQ-25) the impact resulted in a crushed zone of 3 mm diameter and 2 mm depth. Large shallow ring cracks and chipping of 20.5 mm diameter were observed on the front surface. A complete lateral crack of 11 mm diameter was seen. No radial cracks or vertical splinter cracks were observed. The largest, 25.4 mm thick, specimens showed the smallest amount of cracking. At 139 m/s (BQ-21) a crushed zone 3 mm in diameter and 1 mm deep was present. 5 mm diameter chipping existed. A partial lateral crack of 2.5 mm radius appeared. Small splinter cracks 0.5 mm long appeared. No radial or conical cracking occurred. An impact at the velocity of 172 m/s (SP-10) resulted in a crushed zone of 2 mm diameter and 1 mm depth. 3 mm diameter surface chipping was observed. 1 mm deep splinter cracks also occurred. At 229 m/s (SP-12) a 2.5 mm diameter crushed zone occurred. There was 10 mm long chipping to one side of the impact point in a semi-circular pattern. When

90 measured through the point of the impact the maximum chipping diameter was also about 10 mm. A 9 mm diameter lateral crack was seen and 1 mm deep splinter cracks occurred. At 242 m/s (BQ-4), there was a crushed zone of 2 mm diameter and 1 mm deep and chipping of 10 mm diameter from the front surface. No radial, cone, or lateral cracking was present. However, splinter cracking 1 mm in length occurred. At 287 m/s (SP-11) a 2.5 mm diameter crushed zone was observed. A maximum diameter of 11 mm was observed in the surface chipping. 11 mm was also the size of the lateral crack diameter. Splinter cracks of 1 mm depth were also seen. Increasing the impact velocity had little effect on the cracking patterns in thickest specimens. At an impact velocity of 307 m/s (BQ-10) a crushed zone of diameter 2 mm and depth of 1.5 mm was seen with shallow surface ring cracks and material chipping of 8.5 mm diameter from the front surface. As at lower velocities 1 mm long splinter cracks were seen. However, increasing the impact velocity to 336 m/s (BQ-11) did affect the cracking pattern. In experiment BQ11 the crushed zone was 3 mm in diameter and 2 mm deep, while the surface ring cracks increased to 18 mm diameter. Chipping occurred from the surface. Both 0.5 mm long splinter cracks and a sub surface lateral crack 6.5 mm radius formed. To summarize, several types of cracking are found in the specimens following the impacts. Front surface damage includes ring cracks and chipping. Internal damage consists of a hemispherical crushed zone as well as other forms of cracking including radial, conical, lateral, and splinter cracks. The types of cracking patterns that dominate and the size and shape of these patterns are effected by both specimen thickness and impact velocity. It is apparent that the conical and radial cracks were most prevalent in thinner specimens and caused more damage at higher velocities. Rear surface radial cracking appeared where the dynamic bending stresses were greatest, that is in the thinner specimens.

91 The lateral cracks and surface chipping, on the other hand do not show an effect of thickness, but increase in diameter with velocity. The splinter cracks were visible in thinner and thicker specimens at low velocity, but as velocity increased, were visible only in the thicker specimens. Changing the thickness and velocity has different effects on the internal stress state during the impact, which causes the variations in the cracking systems.
C.5.2. Dynamic Evolution of Cracking

As discussed above, the impacts produced several different types of cracking including radial, conical, and lateral cracks. During the experiments, high-speed observations were made of the cracking patterns. These images allow for the formation of cracking to be examined with respect to variations in the impact velocity and specimen thickness. The dynamic crack observations are summarized below. In all cases above a threshold impact velocity, a roughly hemispherical crushed zone formed first directly below the impact site. This crushed zone varied from 1 mm to 3 mm in size. Multiple ring cracks also formed on the surface, although this was difficult to see from the camera images. Debris ejected from the surface was clearly visible in all cases. Most of this debris was generated by chipping that occurred from the surfaces of the specimens. The amount of debris increased with velocity. Cone cracks were formed for specific cases where the thickness of the specimen is small and the impact velocity is large. These cracks were seen to originate either at the surface of the specimen at the ring cracks (BQ-3, BQ-26) or towards the bottom of the crushed zone (BQ-18, BQ-13, and BQ-7) and to proceed downwards towards the rear surface following the smallest principle stress trajectory as in Fields paper [1988]. They formed within 6 s of the impact. The result of the cone crack is a section of the specimen ejected from the rear surface when sufficient impact velocity is reached. For the three higher impact velocities of the 3 mm plates (BQ-3, BQ-9, and BQ-26) the cone crack did eject.

92 The minimum impact velocity in these experiments was 238 m/s (BQ-3). For two of the 5 mm specimens (BQ-13, BQ-7) and for the lowest impact velocity 3 mm (BQ-18) the cone formed but did not completely eject. This is referred to from now on as partial cone formation. In fact two partial cones formed for the 5 mm at 323 m/s (BQ-13) case. Oscillations were seen to occur during formation of the partial cone crack at 5 mm and 300 m/s (BQ-7) and in the outer of the two cone cracks in the 5 mm, 323 m/s (BQ-13) specimen. The radial cracks also appear in the 3 mm specimens at all velocities and in some 5 mm thick specimens (BQ-8, BQ-7, and BQ-13). The slowest 3 mm (BQ-18) and the 5 mm, 234 m/s (BQ-8) & 300 m/s (BQ-7) impacts resulted in small radial cracks, cracks that did not propagate through the entire specimen. The 5 mm, 323 m/s (BQ-13) impact produced one large radial crack that propagated to the side boundary of the specimen. The remaining impacts (BQ-3, BQ-9, and BQ26), generated multiple long radial cracks and fractured. In some cases the radial cracks originated at the front surface. These impacts include the 5 mm, 234 m/s (BQ-8) where the cracks appeared within 6 s & the 5 mm, 300 m/s (BQ-7) where the cracks were seen to appear after 54 s. These cracks did not propagate to the sides. On the other hand, most of the radial cracks appear to originate on the rear surface. These include the rest of the above mentioned specimens. The radial cracks appeared at the following times following the 3 mm impacts: 149 m/s (BQ-18), 36 s; 238 m/s (BQ-3), 12 s; 300 m/s (BQ-9), 12 s; 345 m/s (BQ-26), 12 s. The radial crack in the 5 mm, 323 m/s (BQ-13) specimen was out of the focal plane and thus unseen by the camera. All of the radial cracks opened in a half penny fashion leaving visible marks of this shape on the crack surfaces. In most cases the radial cracking propagated for far greater distances than the conical cracks. The lateral cracks, like those discussed in the literature [Field, 1988] formed upon unloading and nucleated under the surface at the crushed zone boundary. The cracks formed within 12 s of the impact in all cases where they were observed. These cases are: 5 mm at 155 m/s (BQ-14), 234 m/s (BQ-8), & 323 m/s (BQ-13);

93 15 mm at 137 m/s (BQ-19), 307 m/s (BQ-5), & 370 m/s (BQ-25) and 25 mm at 139 m/s (BQ-21) & 336 m/s (BQ-11). Lateral cracks were seen to propagate in saucer like patterns angling towards the front surface. In some cases, they reached the surface, resulting in chipping from the front surface. Also, in several experiments, the lateral cracks were seen to oscillate. Specifically, oscillations occurred in the following experiments: 5 mm at 323 m/s (BQ-13); 15 mm at 307 m/s (BQ-5) & 370 m/s (BQ-25); and 25mm at 336 m/s (BQ-11). The lateral and cone cracks oscillated because of stress wave reflections from the surfaces of the specimens. This is discussed further in the bending section. Splinter cracks were seen in several experiments. They propagated from near the bottom of the crushed zone, as mentioned in the static section. The sizes of the splinter cracks are typically 0-1 mm. The cracks started out traveling perpendicular to the point in the crushed zone that they nucleate from. Some of them curved back towards the front face, but the ones that propagate from the very bottom of the crushed zone remained straight. These cracks are usually observed in the thicker specimens at all velocities. Approximately 1 mm long splinter cracks were seen in the following specimens: 5 mm at 155 m/s (BQ-14) & 234 m/s (BQ-8), 15 mm at 137 m/s (BQ-19) & 230 m/s (BQ-6), and 25 mm at 242 (BQ-4) & 307 m/s (BQ-10). Smaller, about 0.5 mm long splinter cracks were observed in the 15 mm thick specimens at 307 m/s (BQ-5) and the 25 mm thick specimens at 139 m/s (BQ-21) & 336 m/s (BQ-11). Some of these cracks (BQ-10) were observed to oscillate. These cracks while less severe than other forms of damage are still distinct. Inaccuracies in this measurement develop because of the narrow field of focus and the 6.6 or 2.5 microsecond minimum time between camera images. The observations of the dynamic cracking behavior are limited by the inter-frame time. The dynamic evolution of two cone cracks can be seen in Figure 3.16. Radial cracking from both the front and rear faces were observed in all the thinner (3 & 5 mm) specimens and became more severe as the velocity increased. In the 3 mm specimens, the cracks all initiated from the rear surface, and in the

94 three experiments at higher velocities (BQ-3, BQ-9, and BQ-26) specimen fractured into several pieces. The nature of the rear surface cracking can be seen in Figure 3.17. In the 5 mm specimens, no radial cracks were observed in the slowest impact (BQ-14), radial cracking from the front was observed in the middle velocities (BQ-7, BQ-8), and 1 radial crack was observed from the rear surface in the fastest velocity impact (BQ-13). Lateral cracks did not appear in the 3 mm impacts, but occurred in most of the 5 mm, 15 mm, and 25.4 mm impacts. One full lateral crack can be seen in Figure 3.18. The cracks appear to increase in size and fullness with velocity, but not thickness. Surface chipping, which appeared in almost every experiment, appeared to increase in diameter with velocity but not thickness. Splinter cracks are visible in the 5 mm, 15 mm and 25.4 mm thick specimens at the slowest (BQ-14, BQ-19) and medium slow velocities (BQ-6, BQ-8). As the velocity increases, the splinter cracks are only seen in the 15 mm and 25.4 mm thick specimens for the medium high velocity (BQ-5, BQ-10) and only in the 25.4 mm for the highest velocity (BQ-11) impacts.

95
D. Significance of Results

The current study on soda lime glass has resulted in a number of correlations between the experimental data and estimates from theoretical equations and the LS-DYNA numerical model. First, the contact time provides a means of ensuring that the experiments are accurately predicted by estimates from Hertzian theory and the numerical model. Second, the numerical model and theoretical predictions for this wave are compared to the experimental records. The models and the experiments provide information about the effects of impact velocity and specimen thickness on the strain. Third, observations concerning the kinetics of the impact provide information about the coefficient of restitution of the specimen as well as the partitioning of impact energy. This allows for an examination of the relative severity of the variety of damage modes. Finally, the cracking patterns provide an insight into the differing state of damage in the material due to changes in thickness and impact velocity. Knowledge of the types of cracking and their specific patterns can provide insights into the weakening of the material. The experimental contact times are within 3 % of both the LS-DYNA numerical simulations and Hertzian theory. This precise of a correlation indicates a strong match between the three methods. This fact indicates that the assumptions upon which Hertzian theory and the LS-DYNA model are based are applicable to the impact of steel ball bearings on soda lime glass blocks. As a result, both the theory and numerical methods can be used to predict the elastic impact behavior as observed in the experiments. The front and rear surface strains were measured and compared to the LS-DYNA simulations and plots of theoretical equation. In the experiments, the strain on the front surface consists of one large tensile pulse. The initial pulse increases in size and duration with decreasing specimen thickness. The complete pulse can be approximated by the solution provided by Mitra [1964] for the impact over a circular area on the surface of a half space. Additionally, LS-DYNA numerical simulations using the same geometry as the specimens indicate that cracking and

96 the material strength place a limiting factor on the elastic deformation. The LSDYNA numeric simulations with different boundary conditions show that the strain profile at a specific location on the front surface is altered by the return of waves from these boundaries. Additionally, the simulations suggest that the reflections of the stress waves have an effect on the crack formation. Specifically, the dynamic vibrations of the rear surface show oscillation periods in tension that are within error to the oscillation periods of the cracks observed by the high speed camera. This indicates that the stress waves influence the crack propagation. In the LS-DYNA numerical model for an impact on an infinite half space the strain is observed at a location on the front face of the specimen. This record indicates the arrival of a single pulse and several smaller reverberations. Modifying the model geometry to include a finite and decreasing thickness has the following effects on the radial strain: an increasing magnitude of the tensile part of the pulse, the disappearance of the compressive part of the pulse, and alteration of the reverberations. These alterations occur around and after the time at which the stress pulse should reflect from the rear surface and arrive at the measurement location. Adding a finite lateral boundary to the LS-DYNA model geometry has lesser effect on the surface strains and only changes the reverberations following the major strain pulse. The experimental surface strain profiles agree with the LS-DYNA numerical simulations and Mitras [1964] and Lambs [1904] theories on both the structure and the nature of the waves. Experimental strain measurements in the thicker plates have a similar tensile portion of the stress pulse, but no compressive portion. The thinner specimens show an increased tensile magnitude as compared to the thicker specimens, and also no compressive strain. In the thin specimens, the experimental pulses have a greater duration than the models because of the effects of cracking. The magnitude of the experimental strains is an order of magnitude lower than the numerical elastic strains suggesting that

97 cracking prevents the material from straining as much as it would under pure elastic conditions. The kinetics analysis reveals the long time impact behavior of the material to be within the regime of quasi-static understanding. The experimental coefficients of restitution are always much lower than the theoretical or LS-YNDA elastic coefficients and they decrease at a greater rate than the models. The same pattern is seen where the fraction of rebound energy over impact kinetic energy is plotted. Here, increasing velocity correlates with decreasing rebound kinetic energy, more of which is taken up by cracking and elastic energy dissipation at higher velocities. Inelastic processes, such as crack formation account for most of the loss of energy and the amount of inelasticity increases with increasing velocity. Within the elastic portion of the impact, theoretical and numerical estimations show that part of the energy goes to the formation of stress waves. Most of the remainder of the energy goes into vibrations. The experiments show that the radial and lateral cracks propagate the farthest distance parallel to the impact surface and that the conical cracks typically penetrate to the rear surface of the thin specimens causing cone ejection. Of these three cracking patterns, the radial cracks are observed to cause the most severe damage. The conical and radial cracks are most prevalent in the thinner specimens, while the lateral cracks did not appear in the 3 mm thick specimens but occurred in the other specimens. All three types of cracking increased in severity with velocity. Surface chipping and vertical splinter cracking were also observed, although they did not cause the as much damage as the other three cracking types. It is apparent that altering the impact velocity and the geometry of the specimens has effects on the surface strains and the cracking patterns. Cracking patterns present in previous experimental studies are clearly observed exist in the current study. This study provides a description of the changes in cracking patterns due to changes in impact velocity and thickness. The dynamic stress

98 waves, do not dominate the internal cracking patterns at these impact velocities, although they have an effect as seen by oscillations in some cracking patterns. The stress waves do cause surface strains, which are also estimated by numerical simulations in LS-DYNA and by theory. These results indicate that quasi-static and stress wave theories both apply to impacts of this geometry.

99
References
Chaudhri, M. M. and S. M. Walley, 1978. Damage to Glass Surfaces by the Impact of Small Glass and Steel Spheres. Philosophical Magazine. A 37(2): 153-165. Field, J. E., 1988. Investigation of the Impact Performance of Various Glass and Ceramic Systems. U.S. Army: Final Technical Report Contract Number DAJA45-85-C-0021. Field, J. E., Q. Sun, and D. Townsend, 1989. Ballistic Impact of Materials. Inst. Phys Conf. Ser. No 102: Session 7. Oxford. Goldsmith, W. 2001. Impact: The Theory and Physical Behavior of Colliding Solids, Dover Publications. Goodier J. N.; W. E. Jahsman; and E. A. Ripperger. 1959. An Experimental Surface-Wave Method for Recording Force-Time Curves in Elastic Impacts. Journal of Applied Mechanics. 3-7. Hunter, S. G. 1957. Energy Absorbed by Elastic Waves During Impact. Journal of the Mechanics and Physics of Solids. 5 162-171. Knight, C. G.; M. V. Swain; and M. M. Chaudhri. 1977. Impact of Small Steel Spheres on Glass Surfaces. Journal of Materials Science. 12: 1573-1586. Lamb, H., 1904. On the Propagation of Tremors over the Surface of an Elastic Solid. Philosophical Transactions of the Royal Society. A 203 1-42. Love, A. E. H. 1944. The Mathematical Theory of Elasticity 4th Ed. Cambridge University Press: London. Matweb. 2005. www.matweb.com Mitra, M. 1964. Disturbance Produced in an Elastic Half-Space by Impulsive Normal Pressure. Proceedings of the Cambridge Philosophical Society. 69: 683-696. Nathenson, D., G. Chen, and V. Prakash. 2005. Dynamic Response of Soda Lime Glass to Small Particle Impacts. Society for Experimental Mechanics Conference Proceedings. Spath, W. 1961. Impact Testing of Materials. Gordon and Breach: New York.

100
Tillet J. P. A. 1954. A Study of the Impact of Spheres on Plates. Proceedings of the Physical Society. B 67: 677-88.

101
Tables
Table 3.1: Experiment List Including Specimen Impact Velocity and Thickness. Experiment # BQ-3 BQ-4 BQ-5 BQ-6 BQ-7 BQ-8 BQ-9 BQ-10 BQ-11 BQ-13 BQ-14 BQ-18 BQ-19 BQ-21 BQ-25 BQ-26 BN-1 BN-2 SP-1 SP-4 SP-5 SP-6 SP-7 SP-8 SP-9 SP-10 SP-11 SP-12 SP-13 Thickness (mm) 3 25.4 15 15 5 5 3 25.4 25.4 5 5 3 15 25.4 15 3 5 15 3 15 15 15 15 5 5 25.4 25.4 25.4 15 Impact Velocity (m/s) 238 242 307 230 300 234 292 307 336 323 155 149 137 139 371 345 284 290 300 300 239 164 279 174 295 172 287 229 231

102
Table 3.2: Experimental, Theoretical, and Numeric Peak Contact Times. Experiment # BQ-3 BQ-4 BQ-5 BQ-6 BQ-7 BQ-8 BQ-9 BQ-10 BQ-11 BQ-13 BQ-14 BQ-18 BQ-19 BQ-21 BQ-25 BQ-26 SP-1 SP-6 SP-7 SP-8 SP-9 SP-10 SP-11 SP-12 SP-13 Contact Time (s) 6.67 6.67 6.67 6.67 6.67 6.67 6.67 6.67 6.67 6.67 6.67 6.67 6.67 6.67 6.67 6.67 2.50 2.50 2.50 2.50 2.50 2.50 2.50 2.50 2.50 Calculation Label Hertzian 150 m/s Hertzian 230 m/s Hertzian 300 m/s Hertzian 350 m/s Num. 3 mm 150 m/s Num. 3 mm 230 m/s Num. 3 mm 300 m/s Num. 3 mm 350 m/s Num. 5 mm 150 m/s Num. 5 mm 230 m/s Num. 5 mm 300 m/s Num. 5 mm 350 m/s Num. 15 mm 150 m/s Num. 15 mm 230 m/s Num. 15 mm 300 m/s Num. 15 mm 350 m/s Num. 25.4 mm 150 m/s Num. 25.4 mm 230 m/s Num. 25.4 mm 300 m/s Num. 25.4 mm 350 m/s Contact Time (s) 2.61 2.40 2.27 2.20 2.61 2.32 2.22 2.13 2.60 2.37 2.28 2.18 2.61 2.37 2.28 2.18 2.61 2.37 2.28 2.18

103
Table 3.3: Experimental Strain Measurements. Experiment # BQ-3 BQ-4 BQ-5 BQ-6 BQ-7 BQ-8 BQ-9 BQ-10 BQ-11 BQ-13 BQ-14 BQ-18 BQ-19 BQ-21 BQ-25 BQ-26 SP-1 SP-6 SP-7 SP-8 SP-9 SP-10 SP-11 SP-12 SP-13 Amplitude () 107 16.0 128 26.7 58.7 45.3 115 32.0 21.3 48.0 128 66.7 13.3 16.0 24.0 93.3 151 37.1 22.7 45.8 57.1 22.7 28.6 25.7 20.0 Duration (s) 31.2 4.63 3.90 2.10 6.10 10.7 16.1 4.40 9.00 26.0 9.61 44.0 4.20 5.00 6.20 32.9 26.8 3.56 2.60 2.90 2.40 1.63 2.90 6.24 8.00

104
Table 3.4: Impact Velocities, Rebound Velocity, Kinetic Energies, and Coefficients of Restitution. Experiment # BQ-3 BQ-4 BQ-5 BQ-6 BQ-7 BQ-8 BQ-9 BQ-10 BQ-11 BQ-13 BQ-14 BQ-18 BQ-19 BQ-21 BQ-25 BQ-26 SP-6 SP-8 SP-9 SP-10 Impact Velocity (m/s) 238 242 307 230 300 234 292 307 336 323 155 149 137 139 371 345 164 174 295 172 Rebound Velocity (m/s) 71.7 87.3 65.7 83.9 43.0 58.0 70.7 94.5 84.8 61.9 72.4 46.2 94.5 63.2 78.9 38.1 98.0 88.6 92.0 81.9 Rebound Energy/ Impact Energy 0.091 0.13 0.046 0.13 0.021 0.061 0.059 0.095 0.064 0.037 0.22 0.097 0.48 0.22 0.045 0.012 0.36 0.26 0.097 0.23 Coefficient of Restitution 0.30 0.36 0.21 0.36 0.14 0.25 0.24 0.31 0.25 0.19 0.47 0.31 0.69 0.46 0.21 0.11 0.60 0.51 0.31 0.48

Calculation Label Hertzian 150 Hertzian 230 Hertzian 300 Hertzian 350 Num. 3 mm 150 Num. 3 mm 230 Num. 3 mm 300 Num. 3 mm 350 Num. 5 mm 150 Num. 5 mm 230

Energy Ratio 0.87 0.83 0.80 0.78 0.66 0.62 0.53 0.59 0.79 0.74

Coefficient 0.94 0.91 0.90 0.89 0.81 0.79 0.73 0.77 0.89 0.86

Calculation Label Num. 5 mm 300 Num. 5 mm 350 Num. 15 mm 150 Num. 15 mm 230 Num. 15 mm 300 Num. 15 mm 350 Num. 25.4 mm 150 Num. 25.4 mm 230 Num. 25.4 mm 300 Num. 25.4 mm 350

E. Ratio 0.72 0.72 0.80 0.74 0.71 0.71 0.80 0.74 0.72 0.72

Coefficient 0.85 0.85 0.89 0.86 0.84 0.84 0.89 0.86 0.85 0.85

105
Table 3.5: Post Impact Static Cracking Pattern Details

106
Figures

1
Peak Force / Max Force for 350 m/s sim.

Target 5 mm 350 Target 5 mm 300 Target 5 mm 230 Target 5 mm 150 Hertz 350 m/s Hertz 300 m/s Hertz 230 m/s Hertz 150 m/s
Sphere Diameter: 1/16" Plate Thickness: 5 mm

m/s m/s m/s m/s

0.75

0.5

0.25

0.5

1.5

Time / Arrival of Longitudinal Wave at 10 mm Strain Gage


Figure 3.1: Hertzian approximation [Goldsmith, 2001] and numerical simulations for impact force Curve [Nathenson, 2005]. The solid Target curves are the asymmetrical simulations, while the dotted Hertzian curves are symmetric.

107

1 Lamb

Strain / Max Strain

0.5

Arrival of Rayleigh Wave

-0.5

0.98

0.99

1.01

1.02

1.03

1.04

Time / Arrival of Rayleigh Wave


Figure 3.2: Surface strain profile for Rayleigh wave of the general point impact of a half space, using the equations from Lamb [1904]. The numerical differentiation of the theoretical equation

{III.A.10} is shown here. The strain is normalized to the peak tensile strain and the time is
normalized to the arrival of the longitudinal wave.

108

1
Longitudinal Wave Arrival (t1)

Mitra
Longitudinal Wave

Strain / Max Strain

0.5

Shear Wave Arrival (t2)

Rayleigh Wave Arrival (t3)

-0.5

1.5

Time / Arrival of Longitudinal Wave


Figure 3.3: Strain profile for impulsive impact of a half space in a circular area, after Mitra [1964]. The strain is normalized to the peak tensile strain in the longitudinal wave. The longitudinal wave arrives at t1, the shear wave at t2 and the Rayleigh wave at t3. These correspond to limits on the calculation in equations {III.A.18} and {III.A.19} as follows:

t1 =

R , 3 R ,

{III.F.1}

t2 =

{III.F.2}

t3 =

r a . V

{III.F.3}

The symbols are the same as those from equations {III.A.18} and {III.A.19}.

109

600 550 500 450 Bending Basic Quantities Calibration Supplamental Exps. Best Fit Equation
2.1209

Pressure (psi)

400 350 300 250 200 150 100 50 0 0 50 100 150 200 250 300 350 400

Pressure = 0.0019 * Velocity

Velocity (m/s)
Figure 3.4: Calibration curve for projectile impact including all experiments. Also specified is the best fit curve, which happens to be polynomial. The symbols indicate the range of velocities for each target velocity.

110

2.75

Contact Time (s)

2.5

2.25

1.75

Hertz Theory Numeric 3 mm Thick Numeric 5 mm Thick Numeric 15 mm Thick Numeric 25.4 mm Thick Average of Experiments

1.5

50

100

150

200

250

300

350

400

Impact Velocity (m/s)


Figure 3.5: Plot of experimental, theoretical [Goldmsith, 2001], and numerical [Nathenson, 2005] contact time variation with velocity. The numeric values for 5 mm, 15 mm, and 25.4 mm are within 0.01 s for all impact velocities. The accuracy of the experimental contact times are limited by the fastest camera framing rate. The plot shows the maximum contact time for the experiments to be 2.5 s.

111

1 0.75 0.5 Radial Strain 350 m/s Hoop Strain 350 m/s

Strain / Max Strain

0.25 0
Rayleigh Wave Arrival

-0.25 -0.5 -0.75 -1

Longitudinal Wave Arrival

Shear Wave Arrival

Time / Arrival Time of Longitudinal Wave


Figure 3.6: LS-DYNA simulations of the surface strains profiles at 10 mm from the impact location on a half space [Nathenson, 2005]. The impact speed is 350 m/s. This strain is similar in form to the theoretical Mitra wave (Figure 3.3), except that the strain wave has a greater duration and encompasses all three stress waves. This is a result of a finite loading time rather than the impulsive loading used in Mitras computations.

112

0.9
Strain / (Max Normal Strain for 3 mm Plates)

Rayleigh Wave

Shear Wave

1 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 -0.2 -0.3 -0.4 -0.5 0 1 -0.1

Radial 3 mm 350 m/s Radial 5 mm 350 m/s Radial 15 mm 350 m/s Radial 25.4 350 m/s

Finite Thickness Simulations

Rear Surface Wave 3 mm Rear Surface Wave 5 mm

Rear Surface Wave 15 mm

Rear Surface Wave 25.4 mm

Time / (Arrival of Longitudinal Wave)


Figure 3.7: LS-DYNA simulations of the surface strain profiles at 10 mm on plates with infinite lateral boundaries but finite thickness [Nathenson, 2005]. Impact speed is 350 m/s. The 25.4 mm and 15 mm signals overlap during the initial pulse, while the 5 mm and 3 mm initial pulses have greater peak strains and longer duration.

113

0.9
Strain / (Max Strain for 3 mm Plate Simulation)

Rayleigh Wave

Shear Wave

1 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 -0.1 -0.3 -0.4 -0.5 0 1 -0.2

Radial 3 mm Radial 5 mm Radial 15 mm Radial 25.4 mm

Finite Thickness and Lateral Boundaries Impact Velocity: 350 m/s

Lateral Boundary Wave

Rear Surface Wave 3 mm

Rear Surface Wave 5 mm

Rear Surface Wave 25.4 mm

Rear Surface Wave 15 mm

Time / (Arrival of Longitudinal Wave)


Figure 3.8: LS-DYNA simulations of the surface strain profiles at 10 mm on plates with finite rectangular geometries identical to the experiments [Nathenson, 2005]. The impact velocity is 350 m/s. Until the arrival of the lateral boundary wave, no differences are seen from Figure 3.5.

114

Strain / (Max Strain of 3 mm Plate Simulation)

0.1

0.09 0.08 0.06 0.05 0.04 0.03 0.02 0.01 0


Shear Wave

0.07

Rayleigh Wave

Exp 3 mm 345 m/s Exp 5 mm 323 m/s Exp 15 mm 371 m/s Exp 25.4 mm 336 m/s

Experimental Radial Strain


Lateral Boundary Wave

Rear Surface Wave 3 mm

Rear Surface Wave 5 mm

-0.02 -0.03 -0.04 0 1

Rear Surface Wave 15 mm

Rear Surface Wave 25.4 mm

-0.01

Time / (Arrival of Longitudinal Wave)


Figure 3.9: Comparison plot of experimental strain profiles in 350 m/s range for all four thickness specimens. No compressive strains are visible. As in the LS-DYNA simulations, the 3 mm and 5 mm experiments show a greater rise in the peak strain than the 15 mm and 25.4 mm specimens. However, the duration of the 3 mm and 5 mm pulses is much greater than in the numeric simulations. Additionally, the absolute magnitude of the strains is a factor of 10 lower than predicted by the numerical simulations.

115

200

50

(a)

(b)

Primary Tensile Pulse Duration (s)


0 5 10 15 20 25 30

175

Maximum Amplitude (-strain)

40

150 125 100 75 50 25 0

30

20

10

10

15

20

25

30

Specimen Thickness (mm)

Specimen Thickness (mm)

200

50

(c)

(d)

Primary Tensile Pulse Duration (s)


0 50 100 150 200 250 300 350 400

175

Maximum Amplitude (-strain)

40

150 125 100 75 50 25 0

30

20

10

50

100

150

200

250

300

350

400

Impact Velocity (m/s)

Impact Velocity (m/s)

Figure 3.10: (a) The strain amplitude decreases with increasing thickness. (b) Also, the pulse duration decreases with specimen thickness. (c) Variation of strain amplitude with impact velocity. An increasing trend is visible in the strain amplitudes. (d) Plot of pulse duration with impact velocity. No conclusions can be drawn from this plot.

116

0.25 0.2

Voltage (V)

0.15 0.1 0.05 0 -0.05 -0.1

10

12

14

16

Camera Frames

Figure 3.11: Rear surface strains from a bending series impact of a 5 mm specimen at 284 m/s. Two oscilloscopes recorded the strains that register tensile strains for at least the first 50 microseconds (7.5 camera frames). Note the period of the oscillating strains are about 12 to 18 microseconds.

117

0.1 0.075 0.05

Voltage (V)

0.025 0

-0.025 -0.05 -0.075 -0.1

10

12

14

16

Time (s)

Figure 3.12: Rear surface strains from a bending series impact of a 15 mm specimen at 290 m/s. Two oscilloscopes recorded the tensile strains that, considering an offset of about 25 e, register tensile strains for the entire recording period. Note the period of the oscillating strains are about 6 to 12 microseconds.

118

-0.09

Start of First Drop


-0.1

Midpoint of First Drop

Beam Intensity (Volts)

-0.11

End of First Drop


-0.12

Start of Second Drop Midpoint of Second Drop End of Second Drop

-0.13

-0.14

-0.15

-0.16

1E-05

2E-05

3E-05

4E-05

Time (seconds)
Figure 3.13: Example of oscilloscope output from laser velocity triggering system. The three locations in each of the two drops where the time is measured are indicated. These are divided by a fixed distance of 5.9 mm and the resulting speeds are averaged.

119

1 0.9
Hunter's Theory Numeric 3 mm Numeric 5 mm Numeric 15 mm Numeric 25.4 mm Exp. 3 mm Exp. 5 mm Exp. 15 mm Exp. 25.4 mm

Rebound Energy / Impact Energy

0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0

50

100

150

200

250

300

350

400

Impact Velocity (m/s)


Figure 3.14: Variation in ratio of rebound kinetic energy to impact kinetic energy with velocity. The experimental rebound energy is lower than the LS-DYNA or the stress wave estimations, which are both elastic.

120

1
Hunter's Theory Numeric 3 mm Numeric 5 mm Numeric 15 mm Numeric 25.4 mm Exp. 3 mm Exp. 5 mm Exp. 15 mm Exp. 25.4 mm

Coefficient of Restitution

0.75

0.5

0.25

50

100

150

200

250

300

350

400

Impact Velocity (m/s)


Figure 3.15: Variation in coefficient of restitution with velocity. The numeric estimates for the thicker three specimens are within 0.01. As in Figure 3.14, the experimental coefficients are much less than those estimated elastically.

121

Figure 3.16: Impact of 5 mm thick specimen at 323 m/s (BQ-13) showing two conical cracks. Images are 6.67 microseconds apart. Note the formation of a lateral crack near the front surface above the crushed zone.

122

Figure 3.17 Impact of 3 mm thick specimen at 345 m/s (BQ-26) showing radial cracks propagating from the rear surface. Images are 6.67 microseconds apart. Note the cone crack that forms in frame 2 and ejects from the rear surface. Also, developing radial cracks that start in frame 3 have half-penny shape squished by the small thickness of the specimen.

123

Figure 3.18: Impact of 15 mm thick specimen at 371 m/s (BQ-25) showing lateral cracks. Images are 6.67 microseconds apart.

124

Chapter IV Planar Impact Experiments: Configuration and Procedures High velocity planar impacts were carried out using the single stage gas gun facility in the Department of Mechanical and Aerospace Engineering at Case Western Reserve University. These experiments examined the effects of dynamic stress loading on AS800 silicon nitride and soda lime glass under plain strain conditions. The single stage gas gun is designed to accelerate projectiles using compressed air or helium, down an 82.5 mm gun barrel. The specimen is located in an impact chamber; an expansion chamber facilitates soft recovery of post impact debris. Both pure compression and pressure shear experiments at multiple impact velocities were conducted. A Velocity Interferometer System for Any Reflector (VISAR), contact pin tilt system, and laser velocity system are used to obtain the projectile velocity at impact, the particle velocity of the free surface of the target plate and the impact parallelism during the impact process. From these measurements, the dynamic spall strength and Hugoniot Elastic Limit of the material was determined. The effect of impact velocity and skew angle on spall strength was also examined. In this chapter, the single stage gas gun is described, the method of preparing the specimen, projectile and the gas gun for the experiments are explained, and the procedures for conducting the experiments are discussed. A. The Single Stage Gas Gun and Observation Systems The single-stage gas-gun in the D. K. Wright Laboratory was employed in the planar impact experiments (Figure 4.1). This gas gun utilizes compressed nitrogen or helium gas to accelerate the flyer plate up to 550 m/s by means of a fiberglass projectile down the gun barrel. The barrel and impact chamber are evacuated to a pressure of 50 m of Hg to avoid the effects of air shock waves inside the target chamber. The specimen is located in a target alignment fixture within an airtight impact chamber. The impact velocity of the flyer plate is measured by using a laser velocity system, the targets free-surface particle velocity is measured by a laser reflected from this surface and analyzed using the

125

VISAR, and the parallelism between the flyer and target at impact by the tilt pin system. Following impact, soft recovery of debris is accomplished by means of a momentum balance and de-acceleration in soft cloths to absorb the impact. A steel cylinder, which is the momentum balance, catches the debris in the expansion chamber. It transforms the linear kinetic energy of the projectile and the impacted specimen into rotational kinetic energy and gravitational potential energy. This section describes the physical components of the single stage gas gun and the laser and tilt pin based diagnostics. A.1. Description of the Single Stage Gas Gun The mechanism for rapidly accelerating projectiles using high pressures consists of a piston in a seal and firing chamber, connected by a rod to a cap in the loading chamber (Figure 4.2). The seal chamber is pressurized to seal the cap against the breach between the loading chamber and the gun barrel. The load chamber is then pressurized with the gas that will accelerate the projectile. During the firing sequence, the firing chamber gas is released from a reservoir to create a pressure gradient that causes the piston to open the breach. This process is discussed further in the procedure section below. The setup is controlled externally by means of a series of valves in a control panel (Figure 4.2). This system allows the firing to be operated from a distance. The single stage gas gun barrel has an 82.5 mm inner diameter and is 4.57 meters long. It is honed and has a square cross section key-way, which traverses the length of the bore. The fiberglass projectile is accelerated down this barrel, and a Teflon key in the key-way prevents the projectile from rotating while it is accelerating down the gun barrel. A rubber o-ring, at the rear of the projectile, creates an airtight seal between the projectile and the gun barrel. This barrel is secured to the ground by means of a concrete block that prevents gun barrel recoil during the experiment. The gun barrel is connected to both the impact chamber and high pressure chamber by means of bolted flanges. These bolts and

126

o-ring seals maintain a vacuum within the gun barrel and the impact chamber during the experiment. The impact chamber contains the specimen and its holder. It is rectangular in shape and constructed of inch thick steel plates welded together (Figure 4.1). Internal to this chamber is a shelf to support the specimen holder, a tray to catch falling debris and a rod to hold the alignment system. The specimen holder consists of a base plate that adjusts for the skew angle and three rectangular rings, which enable the tilt alignment of the specimen. Glass windows allow for laser measurements to be taken within the chamber. The measurement systems include the VISAR interferometer, laser-based velocity system, and a series of pins that measure tilt between the flyer and the target and indicate the onset of impact. Vacuum feed-throughs are employed for the VISAR interferometers fiber optic cables and the triggering pins relay cables. These feed-throughs are necessary because prior to impact, the gun barrel and impact chamber are evacuated to a pressure of 50 m of Hg. A square guide tube is also bolted to the shelf behind the specimen holder so as to contain and guide the projectile and the impact debris towards an expansion chamber at the end of the impact chamber. A circular opening in the rear of the impact chamber, covered with a Mylar sheet during the experiment, enables the post impact debris to enter the expansion chamber while maintaining a vacuum in the impact chamber during the experiment. The Mylar sheet is destroyed during each experiment. Moreover, the expansion chamber is not evacuated during the experiment. The expansion chamber is considerably larger than the impact chamber. It is also rectangular and made of inch thick steel plates welded together (Figure 4.1). Inside of the chamber, a cylinder, which weighs on the order of 2000 lbs., is suspended from the top plate of the chamber. This balance is filled with cloth rags that, along with a transfer of kinetic energy, absorb the impact and allow soft recovery of the sample. Unlike the rest of the gun, the expansion chamber is not evacuated prior to the experiment. The experiment takes place in the impact chamber, so it is unnecessary a vacuum to be maintained in the expansion

127

chamber. A transitional flange projects from the front of this chamber. This flange prevents debris from ricocheting outside of the expansion chamber during the transfer of the speeding debris. The expansion chamber is mounted on wheeled tracks, which allow the chamber to move away from the impact chamber during setup to allow access to the rear of impact chamber. Prior to the experiments, the transitional flange of the expansion chamber is placed against the rear wall of the impact chamber, covering the Mylar sheet. Post impact specimen recovery is through the transitional flange on the front end of the expansion chamber and also through a circular access hatch on the rear side. A.2. Laser Velocity, VISAR and Tilt Pin Measurement Systems Measurements for the gas gun consist of three types. The first data type is the impact velocity, recorded by means of a set of parallel laser beams across the projectiles path. The second type is the determination of the parallelism of the specimen and flyer interface, as indicated by the staggered triggering times of the tilt pins. The third type of measurement is of the free surface velocity, carried out by interferometer readings. The details of the observation mechanisms for each of these measurements are discussed in this section. A.2.1. Laser Velocity System The velocity of the projectile is measured by the time the projectile takes to traverse the distance between three laser beams set a distance of 9 mm apart. A UNIPHASE Helium-Neon 5mW laser (Model 1125p) is used to generate these three beams. These beams enter and exit the impact chamber by means of a pair of glass windows. The beams are focused by a lens and collected using a high frequency photo-diode. The path of the beams is across the path of the projectile. The beams are interrupted in sequence by the passing projectile. The signal is amplified and the history of the drop in intensity as the projectile cuts each beam is recorded on an oscilloscope (Figure 4.5).

128

The time for the projectile to pass between beams is measured from the plot of the beam intensity versus time (Figure 4.5). This measurement is repeated six times; once for the beginning and end of each of the three intensity drops. The times between the initial and final drop as well as the times between the middle drop and the two other drops are recorded. Dividing the appropriate distances by the recorded times gives six different velocity measurements. These six velocities are averaged in order to specify the impact velocity. A.2.2. Tilt Measurement and Triggering System In order to determine the exact moment of impact between the flyer and specimen, contact pins are employed. The electrical circuit that is completed when contact occurs also triggers the recording systems for the VISAR. The triggering system consists of five conductive pins placed into machined holes in the specimen holding ring. Four of the pins are electrically isolated from the aluminum ring, while the fifth acts as a ground. The pins are lapped until they are flush with the specimen surface. Once any pin is in contact with the flyer ring, which is also in contact with the specimen ring, the electrical circuit is complete. The other three isolated pins however are not impacted at the exact same time as the pin that triggers the system. This delay, measured in tens of nanoseconds indicates the parallelism of the flyer and specimen at the time of impact. Therefore, the triggering system doubles as a tilt measuring system for the purpose of determining any deviation from a parallel impact of the flyer and target plates. The optical alignment procedure imparts an alignment of 0.2 milliradians [Kim et al., 1977]. However, a deviation of less than 0.5 milli-radians is acceptable [Prakash, 1998]. The tilt is measured and checked after each experiment to ensure that the deviation from the parallel remains below this level.

129

A.2.3. VISAR Interferometer The Valyn VISAR interferometer system is used to measure the rear, or the free surface particle velocity of the specimen. The acronym VISAR abbreviates the full name: Velocity Interferometer System for Any Reflector. This system uses a COHERENT VERDI 5 Watt solid-state diode-pumped frequency doubled Nd:YVO4 CW laser. This laser is monochromatic and has a wavelength of 532 nm. The original VISAR interferometer dates from 1972 and was created by Barker and Hollenbach [1972]. Their design enabled the collection of velocity measurements from surfaces that were both specular and diffuse. The current generation of interferometer is multi-beam, and up to seven independent points can be measured simultaneously during impact [Barker, 2000]. The VISAR system uses four glass etalons, which provide a range of velocity per fringe constants from 99.2 m/s/fringe to 1874 m/s/fringe. The minimum resolution is approximately 2 m/s using the 99.2 m/s/fringe etalon setup. One of the features of this system is that it uses fiber optic cable to channel the laser beam between the laser and the probe and the probe and the interferometer. A beamsplitter directs the laser light into a 60 micrometer fiber, which is connected, through a fiber splice into an optical fiber attached to a probe. The probe focuses the light on to a pre-determined spot on the target surface 30 mm away from the probe tip. The motion of the specimen causes a Doppler shift in the light. When positioned at this height over a diffuse or specular surface, the Doppler shifted light returning to the probe is collected with a lens into a 300 micrometer optical fiber. This fiber is connected to a second fiber through another splice. The use of splices allows a probe to be destroyed in each experiment, without damaging the fibers permanently attached to the lasers beamsplitter and the interferometer. [Barker, 2000]. The fiber carrying the Doppler shifted light enters the interferometer, which is mounted on an aluminum frame to keep the components aligned (Figure 4.3). The incoming light is first filtered through an iris. The intensity of the laser

130

beam is monitored using a small portion of the light that is removed with a beamsplitter. This observation ensures that any errors due to a loss of beam intensity during the experiment are recognized. Possible causes of loss of beam intensity include poor alignment of the probe to the specimen, and damage to the free surface of the specimen. Following this, the primary laser beam is split using a 50/50 beamsplitter into two beams. One beam is reflected off a mirror mounted on a piezoelectric aligner/translator (PZAT) that vibrates, creating a motion that allows the interferometer to be aligned, using the bulls eye technique. Three potentiometers within a Burleigh RG-93 Programmable Ramp Generator control the bias of three separate crystals that comprise the PZAT system [Barker, 2000]. The second beam travels through glass etalons, which delay the beam by an integer number of wavelengths. Combinations of four different length etalons enable the resolution of the VISAR to be altered from 99.2 m/s per fringe to over 1 km/s per fringe. A 1/8th wave plate, which only retards the p-orientation of the light but not the s-orientation, is placed in front of a mirror off which the beam reflects. These orientations are perpendicular. The retardation caused in the p-component of the beam is one quarter of a wavelength. This beam then traverses the etalons a second time. The two primary beams are interfered and the resulting beam is again split into two components using the 50/50 beamsplitter [Barker, 2000]. In the alignment process, the PZAT provides pulses in only the one beam that does not pass through the etalons. During the experiment, both of the beams contain temporal pulses, but the pulses are not synchronized because of the delay influence of the etalons. The resulting light and dark fringes are directly related to the rear surface particle velocity. The beams are optimized by the bulls eye method to provide the best fringe contrast. Then, they are each fed into a beamsplitter that directs the p and s components into different fibers. Because the p-wave was retarded twice by the 1/8th wave plate, the p-wave beam is 90o out of phase with respect to the s-wave. Each of these four components enters a

131

fiber and is directed to the HP oscilloscope. The scope allows the beam intensities to be optimized by plotting them as Lissajous curves [Barker, 2000]. The optimized beams are subtracted to increase the signal to noise ratio [Hemsing, 1979]. The s and p oriented components are recorded separately in order to remove the inaccuracy caused by the ambiguity of in the sign of the acceleration and to increase the accuracy of the data reduction. This is known as quadrature coding. The signals are strengthened using 1.2 GHz bandwidth amplifiers and are recorded on a Tektronix oscilloscope [Barker, 2000] with a sampling rate of 2 GHz. Post impact, these signals are analyzed using proprietary software enabling the determination of the particle velocity from the fringe patterns (Figure 4.4a) [Barker, 2000]. The signals indicate an abrupt jump in free surface particle velocity of the sample corresponding to the arrival of the first compressive wave (Figure 4.4b). A deviation from this sharp rise indicates the elastic limit has been exceeded and the material is deforming plastically. In the spall experiments, after the signal has reached a plateau, a dip in the free surface particle velocity is recorded indicating the spall of the material. The depth of this dip indicates the spall strength. The shape of the dip is also indicative of the amount of damage present in the material. The effects of impact velocity and skew angle on the spall strength of silicon nitride and soda lime glass will be discussed further in the Chapters V, VI, and VII.

132

B. Experimental Procedures The experimental apparatus described above was employed in experiments on AS800 silicon nitride and soda lime glass. The VISAR rear surface particle velocity measurements enabled the determination the Hugoniot Elastic Limit, dynamic spall strength, and other strength observations as described in Chapters V through VII. In order for the experiments to be carried out, a standardized procedure was developed for the preparation of the specimen, flyer, and the gas gun itself. The specimen and the flyer were placed within aluminum 7075-T6 rings. The projectile must be assembled. The gas gun must be properly cleaned and prepared for the experiment. Following the setup and the alignment, the experiment is performed using high pressure air or helium to fire the gas gun.

B.1. Specimen and Flyer Preparation In Chapters V through VII, the planar impact experiments on AS800 grade silicon nitride and soda lime glass are described. In these experiments, GC103 tungsten carbide and aluminum 6061-T6 were employed as additional flyer and target materials. In Section 1.1, the configuration of these flyers and specimens is briefly described. More extensive discussions of the materials including the material wave speeds and densities are reserved for the individual chapters. Section 1.2 explains the process for assembling any of the configurations of specimens and flyers. B.1.1. Materials for Specimens and Flyers The experiments described in Chapter V were conducted on AS800 grade silicon nitride. Specifically, for the shock compression and pressure-shear experiments at a 12 degree skew angle, single layer flyers of 3 mm or 4.5 mm thick silicon nitride or 2.83 mm thick CG103 tungsten carbide impacted 8 mm and 5.5 mm thick silicon nitride plates. In the pressure shear experiments, a 12 mm thick

133

polymethylmethacrylate (PMMA) window was attached to the free surface of the specimen. For the multiple shock experiments, as described in Chapter VI, the target (specimen) plate was impacted by dual flyer plates. The flyers consisted of 2.83 mm thick GC103 tungsten carbide or 3 mm thick aluminum 6061-T6 glued to a 3 mm thick silicon nitride flyer. The glue that was used was a TRA-CON binary epoxy. The specimens were 4.5 mm thick silicon nitride with 10 mm thick and 25.4 mm diameter silica glass windows. The reverberation experiments employed 4.5 and 5.5 mm thick silicon nitride flyers and 0.5 mm thick tungsten carbide specimens. In Chapter VII, the soda-lime glass planar-impact experiments are described. In this series of experiments both shock compression and combined pressure and shear at a skew angle of 18 degrees, were conducted. These experiments used soda-lime glass and tungsten-carbide as both flyer and target plates. The experiments employed 5.9 mm thick aluminum 6061-T6, 6 mm thick tungstencarbide, and 12.5 mm thick soda-lime glass as flyer disks. The specimens were 6.5 mm and 12.5 soda-lime glass, or 4 mm thick tungsten-carbide disks. All of these disks were 62.6 mm in diameter. Also, in these experiments, the specimens and flyers were single layers. B.1.2. Assembly of the Projectiles and Specimens The specimens free surfaces are coated with a 150 nm layer of aluminum using the AUTO 306 vacuum coating machine in order to provide for a reflective surface. If the specimen and flyer are layers of multiple plates they are glued together using TRA-CON BIPAX BA-2115 binary epoxy and set to cure in a vise for 24 hours. Both the specimen and the flyer need to be encircled by aluminum 7075-T6 rings in order to facilitate their attachment to the target holder and projectile respectively (Figure 4.6). Holes are drilled around the edges of the specimen ring for the four plastic holding pins, for the alignment bolts, and a

134

tapped hole for a -20 bolt is drilled through the ring. This bolt enables the attachment of the VISAR probe. The specimen ring is also machined with five holes in which metallic pins are inserted for the triggering and tilt measurement system. One is a ground pin and the other four are electrically isolated from the ring. These pins are glued into place in the holding ring using Hardman Red 04001 binary epoxy, which cures in 45 minutes. The impact faces of both the flyer and target rings are then lapped on a Lapmaster machine in order to produce a flatness of 1 to 3 light rings. The surface finish is held to the order of 15 microns by the diamond slurry. A high strength binary epoxy is then used to glue the flyer and specimen to their respective aluminum rings. This is done in another vise arrangement so that the specimen and flyer are aligned precisely in the center of their rings. Before the rings and specimens are secured, Loctite-1711 mold release is used to ensure that the vice is not glued to the specimens. The binary epoxy consists of 35.0 ml of Loctite-9412 Hysol resin and 10.0 ml of Loctite-9412 Hysol hardener. After curing period of 24 hours, the impact surfaces are lapped again on the Lapmaster in order to ensure that the impact surfaces, the epoxy surfaces, and the trigger pins still have the prescribed flatness. Wires of 16 inch length are soldered to the trigger pins on the side away from the impact face. Four plastic holding pins are then glued into the holes in the outer surface of the specimen ring. These pins are of lengths ranging from 39 mm to 23 mm and are designed to align the target to the center of the target holder (Figure 4.6). These pins break during the experiment preventing damage from occurring to the permanent components of the target holder. The projectile is a hollow tube of fiber glass with diameter 3.25 inches and length 1 foot. The outer surface of this projectile is machined to a tolerance of +0 and 0.005 inches in order to allow for smooth fitting in the barrel of the gun. Angled projectiles are marked indicating the orientation of the key-way in the barrel with respect to the major axis of the elliptical front of the projectile. An

135

aluminum cap is glued using the Hardman binary epoxy into the circular groove machined into one end of the projectiles inner diameter. This cap is aligned to the location of the key-way. A machined o-ring grove encircles the cap. Additionally, a hook allows the projectile to be pulled through the barrel. Once this glue has cured, four small holes are machined in the projectile. Two of these holes are along the line of the key-way near to the end cap and are designed equalize pressure within the projectile. The other two holes are opposite each other and are located about 5 mm from the open end of the projectile. These are used in the alignment procedure. One hole is machined 45 degrees off the key-way line and the other is 180 degrees opposite the first hole. Following this, the outside of the cap and projectile are sanded to remove any glue and to reduce the diameter of the projectile near the capped end, which typically swells slightly once the cap has been glued into it. The projectile is then pulled through the barrel to ensure to ensure a tolerance on the fit that allows the projectile to slide freely, but not slip. Following this, the flyer ring is glued with the Hardman binary epoxy to the front of the projectile ensuring that no glue contaminates the outside of the projectile. After this glue has cured, a rubber oring is placed in the o-ring grove on the cap, and the machined Teflon key is placed in the appropriate slot on the cap. The key and o-ring are secured with vacuum grease. B.2. Gas Gun Setup Sequence The single stage gas gun is now prepared. First, acetone soaked cloths are pulled through the barrel until any dirt or debris from previous experiments is removed. The VISAR interferometer probe is then threaded through the vacuum feedthrough plate and the probe is then inserted into the probe holder, which in turn is screwed into the appropriate hole on the specimen ring. This probe holder allows the VISAR probe to be aligned to the center of the free surface, with the proper distance of 30 mm between the probe tip and the free surface. The VISAR probe is adjusted until the maximum return signal intensity is achieved.

136

As mentioned above, the free surface is coated with aluminum so as to be more reflective and provide a stronger return signal. The projectile and flyer assembly is then placed into the impact chamber end of the barrel with a rope connected so that it can be pulled back through the barrel. The key on the cap is inserted into the barrels key-way ensuring that the flyer will remain in the same orientation throughout the alignment process. The specimen is attached to the specimen holder. The specimen holder is then placed inside of the chamber and bolted down. The trigger wires are attached to the appropriate connections on the holder, which connects them to the triggering system. Once this is complete, the alignment of the specimen and the flyer commences. The flyer and specimen are aligned to each other by using circular paper guides and adjusting the specimens vertical and horizontal location in the specimen holder. Once the specimen and flyer are centered on each other, the alignment pins are locked down holding the specimen assembly in place. The paper guides are then removed, the specimen and flyer impact surfaces are cleaned with acetone, and the projectile is pushed into the barrel until it is protruding only enough to let the two holes drilled opposite each other to be accessible. Rubber bands are stretched from screws along these holes and the corresponding ones in the specimen ring. Dust free disk mirrors are placed in these rubber bands over the flyer and projectile. A mirror prism on an adjustable stand is then placed in the chamber along with a free standing mirror. Together, this system allows the angular alignment of the specimen with the flyer in three stages. First, a light bulb is employed for rough alignment. Then, two auto-collimators are employed successively to bring the angular alignment to 210-5 radians. This optical alignment process employs a technique that was pioneered by Kim et al. [1977]. The trigger and velocity systems are tested. The intensity of the VISAR signal is also checked.

137

Following this, a square guide tube is placed into the chamber to ensure that once fired, the projectile will be directed into the expansion chamber. This tube is bolted into place. The projectile is then pulled through the barrel to the breach end. After this, the breach between the high pressure chamber and the barrel, which is sealed with an o-ring, is bolted closed. The rear wall of the impact chamber is sealed, by bolting down a circular plate, secured with an oring, and a Mylar sheet. This Mylar sheet covers the opening that allows the projectile and flyer to proceed into the expansion chamber following the impact without causing damage to a permanent component of the gas gun. completes the setup procedures for the system. This seal also permits a vacuum in the impact chamber during the experiment. This

B.3. Firing Sequence The pre-firing sequence consists of vacuuming the chamber and setting up the measurement systems. First, the PRAXAIR compressed air or helium tank is connected to the regulator of the control system (Figure 4.2). The regulator is not opened yet. All three of the vent valves; the seal vent valve, fire/reservoir vent valve, and load vent valve; are opened and any remaining air from previous tests is allowed to escape. The firing valve, and the three regulator valves; the seal regulator valve, fire/reservoir regulator valve, and load regulator valve; are also opened. The system is thus clear of any remaining pressurized air from the previous experiments. The seal vent valve is closed first, followed by the load vent valve. The fire/reservoir vent valve is left open for the moment. The seal regulator, fire/reservoir regulator and load regulator valves are also closed. The fire valve is ensured to be open. The compressed air tank is then opened and the tank pressure is recorded. The air tank regulator valve is then opened to above the desired loading pressure. This pressure is also recorded. The seal regulator valve is then opened until the desired seal pressure, usually 60 psi, is reached. This causes the piston to seal the breach between the loading chamber and the gun

138

barrel (Figure 4.2). The seal regulator valve is then closed. The impact chamber pressure measurement panel power is engaged at this point. The vacuuming of air from the impact chamber is the next phase of the setup. First, the exhaust valve on the top of the impact chamber is closed. Then, the check valve between the impact chamber and the impact chamber vacuum gage is opened. The vacuum pump is then filled with oil to the appropriate level and the pump is started. After a few moments, the check valve between the vacuum pump and the gun barrel is opened. This is done first in order to keep the projectile from being pulled down the barrel towards the impact chamber by the pressure differential. A minute later, the check valve between the impact chamber and the vacuum pump is opened. The pressure drop rate is monitored by the impact chamber vacuum gage and if it slows below a certain rate, the chamber is checked for leaks. Tightening bolts, or adding additional vacuum grease typically stops any leaks. Once the chamber pressure is down to around 100 milli-torr, the VISAR system is recalibrated. Following this, the velocity system is set up using its oscilloscope to check for proper beam intensity and contrast. The next step is to pressurize the load chamber. First the air tank pressure is confirmed to be sufficient for the desired loading pressure. Then the load regulator valve is opened to the desired pressure, which is recorded. The pressure necessary for a certain velocity is a factor of flyer weight and projectile tightness in the barrel. As a result, each experiment requires an estimation based upon past experimental velocities and pressures. For example, at a speed of 65 m/s, with an average amount of friction between the barrel and projectile, a load pressure of 30 psi is typically used. The final firing procedure commences after the loading chamber is pressurized. First the firing valve is closed and then the fire/reservoir vent valve is closed. After the vacuum pressure is at firing level, which is 80 milli-torr or less, the fire/reservoir regulator valve is opened and the pressure in the auxiliary reservoir

139

chamber is set equal or higher than the load pressure. This pressure is noted. Then the fire/reservoir valve is closed. The compressed air tank main valve is then closed. The seal pressure is dumped by opening the seal vent valve, which removes most of the pressure holding the breach closed. The remaining pressure comes from the loading chamber pressure. The check valve between the vacuum gage and the impact chamber is closed. Then the check valve between the vacuum pump and the gun barrel is closed. All of the optical and electronic recording systems are confirmed to be aligned and armed at this time. The check valve between the vacuum pump and the gun barrel is closed. The vacuum pump is then shut off. A second check of all systems is performed ensuring that all recording devices are armed, that the tilt system is functioning, and that the VISAR intensity as measured by the Lissajous figure on the oscilloscope is a circle of greater than 100 mV diameter. This ensures a strong VISAR signal. After a five count the firing valve is pulled. This releases the air in the firing reservoir into the firing chamber, which moves the piston, allowing the air in the load chamber to escape down the barrel. Following the experiment, all of the records are transferred to disk and then analyzed by means of dedicated programs. The VISAR software produces velocity versus time curves for the rear surface velocity (Figure 4.4). The laser velocity system measures beam intensity, which is translated into velocity by observing the drops in intensity (Figure 4.5). The tilt system is recorded and the drops in voltage indicate the amount of parallelism between the flyer and the target. As mentioned above, the acceptable parallelism is 0.5 milli-radians or fewer [Prakash, 1998]. The data analysis will be discussed further in the next three chapters.

140

C. Summary The single stage gas gun is employed to create plane strain shock compression and pressure shear impacts on silicon nitride and soda lime glass. This gas gun consists of a high pressure chamber, a 4.572 m long barrel with an 82.55 mm diameter bore, an impact chamber, and an expansion chamber. The impact chamber contains a specimen holder, which is used to align the specimen and flyer. The experiment is conducted in vacuum of less than 80 milli-torr to avoid the effects of shock waves in air. Measuring systems include the VISAR interferometer, a laser velocity system, and a tilt system built into the triggering mechanism. The VISAR uses a probe system, which allows precise alignment to the free surface for the strongest signal. The experimental procedure involves the preparation of a projectile using the flyer plate, preparation and alignment of the specimen, and operation of the gas gun. The flyer and specimen are placed in aluminum 7075-T6 rings and secured using epoxies. These rings enable the flyer to be attached with the fiber glass projectile and the specimen to be connected to the specimen holder and the VISAR probe. The gas gun is cleaned, and the specimen and flyer are aligned to within 0.5 milli-radians before the vacuuming process. The firing sequence involves a rapid release of the loading gas, which expands into the gun barrel and accelerates the projectile. Planar impact experiments were chosen because of the goals of this study, among which the determination of the effects on the spall strength and the dynamic strength of various impact velocities and skew angles are the most important. By means of the VISAR, these quantities were determined. The results, which are discussed in the next three chapters, provide insights into the behavior of the AS800 silicon nitride and soda lime glass under plane strain shock compression and pressure-shear.

141

References
Barker L.M, Barker V.J., Barker Z.B., 2000. Valyn VISAR Users Handbook. Albuquerque, New Mexico, USA. Barker L.M., Hollenbach R. E., 1972. Laser interferometer for measuring high velocities of any reflecting surface. Journal of Applied Physics, 43(11), 4669-4675 Hemsing W.F., 1979. Velocity sensing interferometer (VISAR) modification. Review of scientific Instrumentation, 50 (1), 73-78. Kim K. S., Clifton R.J., Kumar P., 1977. A combined normal and transverse displacement interferometer with an application to impact of Y-cut Quartz. Journal of Applied Physics, 48, 4132-4139. Prakash V., 1998. Time-resolved friction with applications to high speed machining: experimental observations. Tribology Transactions, 41 (2), 189-198.

142

Figures

Figure 4.1: Overview of the single stage gas gun. The high pressure chamber, gun barrel, impact chamber and expansion chamber are shown. Inside the impact chamber, the specimen holder and guide box are indicated. The momentum balance inside the expansion chamber is shown. The control and observation equipment is also indicated.

143

Figure 4.2: Firing chamber and air supply diagram. The loading paths from the compressed gas tank to the seal, fire/reservoir, and load chambers are indicated along with the pressure gages. The regulator and vent valves are also shown.

144

Figure 4.3: Valyn VISAR interferometer design. The fiber optic cables are terminated in the laser collimator. Part of the beam is split off and directed into the beam intensity monitor. The main beam travels through the large beamsplitter, and is split 50/50. One path reflects from the mirror fixed to the piezo-electric crystals. The other path is delayed in the etalons and the 1/8th wave retardation plate further delays the p-orientation of this beam. Next, recombining and splitting the beams into the left and right exit beams occurs. The polarizing beamsplitters separate the s and p orientations of the beams. The resulting four beams are collected by output fibers.

145

Figure 4.4: The output of the VISAR interferometers is shown. (a) The fringe patterns for a representative experiment as recorded by the oscilloscopes. (b) The processed velocity curve for this representative experiment, using the VISAR analysis program.

Figure 4.5: The velocity measurement system beam intensity plot for a representative experiment. The blue circles represent the points at which the time is measured at the beginning and end of each intensity drop. The time between succeeding drops and between the first and last drops are used to calculate the impact velocity.

146

Figure 4.6: The specimen and flyer ring configurations for a representative experiment. The images show a top view and a cross section view through the center of the assemblies. A representative single stage flyer is shown on the left side. The aluminum 7075-T6 ring is secured to the flyer by Hysol resin. On the right side is a representative specimen. The specimen has a window attached by TRA-CON binary epoxy. The specimen is secured to the ring with Hysol resin. The large tapped hole in the lower left corner of the aluminum ring is for the probe holder. The remaining five holes are for the triggering system pins. Six holes are drilled in the side of the ring. The four plastic pins, which connect the specimen assembly to the specimen holder are shown in their holes. The remaining two holes are for the bolts which anchor the rubber bands during the fine alignment of the specimen and target.

147

Chapter V The Shock Response of AS800 Grade Silicon Nitride


Plate impact shock wave experiments were conducted on AS800 aerospace grade silicon nitride in order to examine the effects of the state of stress on its dynamic material behavior. The use of nominally plane wave loading conditions on the impacted specimens enable quantitative determination of the material response. In order to understand the motivation for the experiments described herein, previous theoretical and experimental work on AS800 silicon nitride is first discussed. The equipment and procedures involved in conducting these experiments were discussed in detail in Chapter IV. However, the specific theoretical background that is useful in the analysis of these experiments is discussed in detail in Section A of this chapter. Following this, in Section B, the specimen and flyer properties are provided and the experiments are categorized. In Section C, details of the experimental results and analysis are provided. The experimental data include the elastic and hydrodynamic Hugoniot state parameters, the spall strength, and the post impact microstructure. This set of dynamic material properties provides the basis for characterizing the dynamic material response of AS800 silicon nitride under intense shock wave loading conditions.

148

A. Background on Shock Compression Experimentation In order to discuss the dynamic response of the material, first the basic theory of high speed planar impacts is reviewed. This includes the theory of stress wave propagation from impacting surfaces under both elastic and elastic-plastic cases. In this discussion, the definitions of the Hugoniot Elastic Limit and dynamic yield stress are provided. Also, the process of spallation, or the delamination of the material due to internal tensile stresses resulting from rarefaction waves is discussed. Previous shock compression experiments on silicon nitrides from the literature are examined. Additionally, the motivation for the current study is developed.

A.1.

The Theory of Planar Shock Compression

The impact of two plates under planar shock compression results in the propagation of stress waves in the two materials. The amplitude of the resulting shock waves and the particle velocities are proportional to the impact velocity. Eventually, if failure does not occur, the reverberations of the stress waves within the two plates result in an equilibrium state of stress. According to one dimensional wave theory, the stress waves cause the material stress and particle velocity states to alter in a predictable fashion, which enables the computation of the state of stress in the material from the measurement of the particle velocity at the free surface of the target plate. This predictability allows the determination of the dynamic material properties under both elastic and elasticplastic conditions. The equations that describe both of these regimes and the Hugoniot Elastic Limit that marks their separation are discussed in Sections A.1.1 and A.1.2.[VP2] A.1.1. Elastic Impact Theory The study of the transient stress waves in solids is its material response before stress equilibrium is reached. Because this is the case, the material response is determined by the propagating compression/tension and/or rarefaction waves within the solid. The velocity at which these waves propagate is unique to each

149

material. For example, when two material plates with parallel surfaces impact each other in the elastic regime, both longitudinal compression and shear waves are generated. The longitudinal compression waves propagate through the material at the longitudinal wave speed, CL, which is calculated using the elastic modulus, E, density, , and Poissons ratio, v, through the well known relation
E (1 - ) . (1 - 2 )(1 + )

CL =

{V.A.1}

The shear wave velocity, CT, is given by


CT = CL . (2(1 - v )/(1 - 2v ))

{V.A.2}

Because these waves have well defined propagation velocities within the material, a time-distance diagram (T-X), as shown in Figure 5.1a, can be produced which determines the wave propagation in both time and depth in the material. The TX diagram contains information on the temporal and spatial wave propagation in the flyer and specimen [Boslough, 1993].[VP3] Figure 5.1a shows the time distance diagram in a material subjected to combined compression and shear loading. The longitudinal waves, which travel faster than the shear waves, have lesser slope because the slopes of the lines in this diagram are equal to the inverse of the propagation wave velocity. The waves propagate both forward (into the specimen) and backwards (into the flyer) from the point of impact, causing the two material plates to compress. The free surfaces at the far ends of the two plates cause the compression waves to reflect as tensile rarefaction waves, which unloads the two impacting plates [Boslough, 1993]. Since the flyer plate is thinner than the target plate (specimen), the two rarefaction waves meet inside of the specimen and can result in spall of the target plate if the stress level is high enough [Grady, 1988]. This spall process and the critical conditions for it to occur are discussed in more detail in the next section.

150

In a pressure-shear experiment, the shear waves also propagate from the impact point and reflect from the free surfaces. However, because of the elastic nature of impact, the shear waves and the longitudinal waves propagate independently of each other. That is, when the waves intersect, neither is affected by the other. The longitudinal and shear waves each produce a stress consistent with the wave type. The longitudinal wave causes compression, and the shear wave causes shear. These stresses combine to produce the overall stress state in the material. The stresses produced in the materials can be determined by characteristic equations
[VP5]

which are governed by one-dimensional elastic hyperbolic wave

theory [Achenbach, 1975]. These characteristic equations are also referred to in this chapter as line equations because the hyperbolic wave theory can be approximated as linear relations between the stress and the particle velocity. Continuum theory states that the sum of the stress, , and the quantity materials acoustic impedance times the particle velocity, V, remains constant during the propagation of the stress wave. The acoustic impedance is defined by the wave speed, c, multiplied by the density, . The line equation can be written as

cV = constant .

{V.A.3}

If this equation is applied to each of the stress states in the T-X diagram (Figure 5.1a), a set of stresses and particle velocities can be built up. The stress states before impact can be obtained from the flyer (State 1: v= impact velocity, = 0)

and the specimen (State 2: v = 0, = 0). Using the equation twice results in a pair of equations for the impacted state in both the flyer and the specimen, known as state 3: 1 - cV1 = 3 - cV3 , 2 + cV2 = 3 + cV3 . {V.A.4} {V.A.5}

151

Solving this pair of equations for the two unknowns yields the stress state in the impacted material. This stress state is also known as the Hugoniot state. In theory, this process can be repeated for each different stress state resulting in a stress velocity pairing for each state which can be plotted on a stress versus particle velocity ( S-V) diagram (Figure 5.1b). Here, the slopes of the lines correspond to the materials impedances. If the flyer and target have two different impedances, the slope will be different, as shown in Figure 5.1b. To reiterate, the stress and particle velocity states of the material that are depicted in Figure 5.1b are caused by the propagation of stress waves shown in Figure 5.1a. The initial states of the material before impact and after impact, but before the longitudinal wave fonts arrive are State 1 for the flyer and State 2 for the specimen. These are represented in the triangular regions in Figure 5.1a under the first pair of lines representing stress waves. State 3 is the compressive state that occurs following the impact in the interior of both plates. State 4 occurs on the free surface of the specimen where measurements can take place. State 5 is the corresponding state at the free surface of the flyer. Both of these states are completely unloaded and have zero stress in the absence of shear. State 6 is created by the combination of the two rarefaction waves. This is the state where spallation may occur. State 7 is the state that is created on the specimens free surface by the interaction of the two stress waves at State 6.. Any spall in State 6 will affect the particle velocity and the stress in State 7. These states define the material behavior. The stresses and particle velocities of these states can be determined if the free surface particle velocity time history and flyer impact velocity are known from experiments.

A.1.2. Elastic-Plastic Impact Theory and Experiments

The characteristic line equations can also be used for describing the material deformation in the regime of elastic-plastic deformation. However, the elasticplastic wave speed is slower than the corresponding elastic wave speed. As a result, the plastic impedance is smaller than the elastic impedance. The plastic

152

waves on the time-distance diagram propagate from the impact point in a similar manner to the elastic waves, with a smaller slope due to the slower plastic wave speed. The Hugoniot Elastic Limit is the point where the material response reaches the elastic limit under dynamic plane strain loading conditions. When this stress level is exceeded on the stress-velocity diagrams, the impedance difference produces a slope change. This point, hereafter referred to as the HEL, is directly related to the dynamic yield strength of the material by the expression 1 2 . 1

YO = HEL

{V.A.6}

In this equation, the materials HEL, HEL , is related to the dynamic yield stress, Y0, by a function of the Poissons ratio, v [Reinhart and Chhabildas, 2002]. The HEL represents the dynamic tensile strength in plane strain. In the plate impact experiments, there is no strain orthogonal to the impact direction. The dynamic yield strength Y0, can be measured directly under plane stress conditions, such as those obtained in a Split Hopkinson bar type experiment. Additionally, the HEL can vary with grain size [Mashimo, 1994] [Mashimo, 1998] and material density [Nahme, 1994]. In the present study, three different monolithic silicon nitrides were examined. The specimens varied in grain size and porosity, but the porosity variation was small. All the materials had Al2O3, Y2O3, and other binders in their makeup. The first group of specimens had a grain diameter between 0.5 m and 1.2 m, an un-shocked density of about 3.16 g/cc, a porosity of 4%, and an HEL between 10 GPa and 12.5 GPa. The second group had a grain size between 0.25 m and 0.4 m, an un-shocked density of about 3.22 g/cc, a porosity of 1%, and a HEL between 14 GPa and 16.5 GPa. The third set had grain diameters of 0.15 m to 0.3 m, the same density as the second group, a porosity of 1% and a HEL between 17 GPa and 20 GPa. Mashimos [1998] study suggests that the larger the grain size, the smaller is the HEL.[VP7]

153

It is important to note that the HEL manifests itself as a kink in the rise time of the free surface particle velocity versus time profile during the loading from State 2 to State 4. The HEL in a silicon nitride of similar density to AS800 (3.15 g/cm3) was observed to be 12.1 GPa [Nahme, 1994]. A less dense silicon nitride (2.28 g/cm3), in contrast, had a lower HEL of 1.9 GPa. At applied compressive stresses greater than 5.2 GPa, a second higher slope particle velocity region was also observed. The authors state that this is similar to the double-HEL behavior of TiB2. The collapse of voids is suggested as an explanation for this phenomenon [Nahme, 1994]. The elastic-plastic impedance can be estimated by employing the bulk modulus, K, of a material. The bulk modulus is related to the elastic Youngs modulus E, and Poissons ratio , as K = E [3 (1 2 )] . {V.A.7}

In terms of the bulk modulus, the plastic or the bulk wave speed, cP, can be defined as

cP = K 2 .

{V.A.8}

In Equation {V.A.8}, is the density of the material. The elastic-plastic impedance is calculated by multiplying by the density with the plastic wave speed as Z P = cP . {V.A.9}

This elastic-plastic impedance can be used to estimate the stress and the particle velocity states when the state of stress is close to the HEL. This alters the single line estimation of the loading of the silicon nitride from State 2 to State 3 on the stress and velocity diagram (Figure 5.1b) to a double line (Figure 5.2b). The line

154

connecting State 2 to the HEL has a slope equal to the elastic impedance. The line connecting the HEL to State 3 has the slope equal to the elastic-plastic impedance. Note that the tungsten carbide flyers depicted in both diagrams are represented by a double lines while going from State 1 to State 3. This is because the HEL of tungsten carbide is smaller than that of silicon nitride. The tungsten carbide flyer plates are employed in the experiments due to the higher impedance of tungsten carbide when compared with silicon nitride. This produced a higher stress level for the same impact velocity. The Hugoniot Elastic Limit of tungsten carbide has been reported to be 7.2 0.8 GPa [Dandekar, 2004]. To use the double lines to predict State 3, the HEL must first be determined. Knowing the HEL, the particle velocity corresponding to the HEL can be calculated using VHEL = 2HEL / ZT . {V.A.10}

In Equation {V.A.10} the target impedance is represented by ZT. For the silicon nitride described by Nahme [1994] with a similar density to AS800, the HEL velocity is 682 m/s. For the tungsten carbide in Dandekars [2004] work the HEL velocity is 136 m/s. Using the [VP9]stress and particle velocity state of (VHEL ,

HEL ) instead of the zero stress and zero particle velocity for State 2, the particle
velocity V3, and the stress 3, in State 3 can be determined as V3 =(ZF V1 +ZT-PHEL/ZT -HEL )/(ZT-P +ZF ) , {V.A.11} {V.A.12}

3 =ZT-P V3 +(HEL -ZT-PHEL/ZT ) .

In Equations {V.A.11} and {V.A.12}, the impact velocity from State 1 is V1, the flyer impedance is ZF, and the elastic-plastic impedance of the target is ZT-P. From these estimates of the stress and particle velocity in State 3 particle velocity and stress estimates in the remaining states can be determined. It should

155

be noted that the unloading to State 4 is also elastic-plastic, but the remaining states are elastic (Figure 5.2b).
A.2. The Failure in Dynamic Tension due to the Spallation of Silicon Nitride

As stated above, when the rarefaction induced tensile stress reaches a level beyond a materials ability to resist decohesion, it fails in a process known as spallation. This occurrence represents a dominant failure mode in materials subjected to dynamic loading in dynamic tension. In section A.2.1, the process of spall failure [Dremin and Molodets, 1990] and estimation of the spall strength [Grady, 1988] are discussed. Following this, in Section A.2.2, previous experimental works conducted to determine the spall strengths of tungsten carbide [Dandekar, 2004] and silicon nitride [Nahme et al., 1994] are discussed.

A.2.1. The Theory of Material Spall

The spall of a material is the decohesion of a material during its interaction with tensile stress waves. The spall of a material results in the formation of new material surfaces in the interior of the specimen (for example, State 6 in the T-X and S-V diagrams (Figure 5.1)). Here, two studies are presented that describe the spall process. The first [Dremin and Molodets, 1990] indicates that the spall in metals occurs as nucleation and propagation of flaws. The second [Grady, 1988] takes a more mathematical approach and describes the criterion that must exist for spall to occur. A. N. Dremin and A. M. Molodets discuss spall as a two step process [1990]. They describe the process as one where nucleation and unstable propagation are divided into distinct stages where the stress is increasing and decreasing respectively. Then, they discuss the micro-mechanism of damage. The nucleation of microscopic flaws occurs irregularly via a thermo-activation process with local stresses causing a decrease in the inter-atomic bond activation energy. The propagation of spall in metals was observed to occur by voids in aluminum

156

and copper and by micro-cracking in iron and low carbon steels. Either, the voids or the micro-cracks coalesce to form the spall plane [Dremin and Molodets, 1990]. In Gradys paper, spall is acknowledged to involve both energy methods and micro-structural details [Grady, 1988]. The requirements for the initiation of spall according to both energy and microstructure based models are discussed. Flaw induced spall, where sufficient energy is available for the spall process, but the lack of material flaws is the limiting factor is one model. In this case, the stress waves impart sufficient energy in the material to drive the generation of the new surfaces, which are created by the spall process. However, the spall process can only initiate when a flaw such as a void or a micro-crack is created. Energy induced spall, where the microstructure is assumed to be flawed such that when the critical energy criterion is reached the material immediately spalls is the main focus of Gradys [1988] paper.[VP12] This criterion assumes that flaws already exist in the material to such an extent that when the stress waves have imparted enough energy for the formation of the new surfaces, spall initiates at one of these pre-existing flaws. The spall of a material is defined by Grady [1988] to be an internal fracture caused by a dynamic stress state higher than the material tensile stress. A theoretical spall strength, Pth, can be determined by equating the cohesive energy with the elastic energy that is imparted to the lattice due to the tensile loading
2U coh Bo . vo

Pth =

{V.A.15}

This theoretical spall strength is defined in terms of the bulk modulus, Bo, the specific cohesive energy, Ucoh, and the specific volume, vo [Grady, 1988]. The spall strength of brittle materials is limited by either a minimum energy condition or by a maximum pre-existing flaw size. The energy criterion states

157

that the fracture surface energy must be equaled or exceeded by the sum of the kinetic energy and the elastic energy. A minimum time criterion, ignoring the kinetic energy, which is of an order of magnitude smaller than the elastic energy, U, results in equations for the brittle spall strength, Ps, time to spall, ts, and fragment size, s. The values used for fracture toughness, Kc, were quasi-static. The equations are as follows

Ps = (3co Kc2 ) , ts = 1 co

13

{V.A.16}

3Kc co

23

{V.A.17}

s =2

3Kc co

23

{V.A.18}

In Equations {V.A.16} through {V.A.18}[VP13], , is the density, c0, is the longitudinal wave speed, and, , is the strain loading rate [Grady, 1988]. The criterion for energy limited spall can be written directly as U t 3 / co , Where, = Kc2 2co . {V.A.20} {V.A.19}

In Equations {V.A.19} and {V.A.20} the material spall value is independent of the loading strain rate. The energy limited spall criterion is met when the elastic energy times the spall time, U t , meets this value. This indicates a minimum value for the spall, which necessitates a favorable condition of microstructural flaws. That is, that a flaw of sufficient size exists such that spall occurs as soon as the spall criterion is met. Experimental comparisons show overlap between experimental and calculated data for this criterion [Grady, 1988]. This indicates

158

that this energy criterion is a good predictor of the spall strength for brittle materials. Ductile spall on the other hand occurs through micro void coalescence and the fracture energy, W, can be found approximately by the multiple of the flow stress

Y, times a critical void volume fraction, c . This ductile fracture energy has no
dependence on fragment size due to the ductile process. Equations can be written for the spall strength Ps, fracture time ts, and fragment size s, as for the brittle case in terms of the flow stress and critical void volume fraction

Ps = (2coY c ) , ts =
1 2 2Y c co 2 co

12

{V.A.21}

12

{V.A.22}

s = 8Y c 2

12

{V.A.23}

Also noted in Gradys paper [1988] are effects due to strain rate, temperature, scale, and pre-compression. The strain rate during the loading and the energy of fracture effects the spall properties, but this effect is small. The flow stress increases with strain rate slowly while the fracture toughness decreases slowly. Analogous effects on Kc, and Y, are caused by increasing temperature. Also, if the spall process occurs on the same length scale as the process zone ahead of the crack, the energy release rate can be altered. Elastic-plastic compressive shock loading can cause material damage prior to spall, which affects the spall strength. These factors can all effect the material spall properties [Grady, 1988]. However, these effects are small when compared with the effects of impact velocity and the skew angle of impact during combined pressure and shear loading.

159

A.2.2. Experimental Determination of Spall Strength

In previous studies, the spall strength of materials was found to vary with impact velocity, or equivalently the maximum compressive stress. A study by Dandekar et al. [2004] in tungsten carbide found that the spall stress decreases with impact velocity. This study found that variation of the skew angle or the duration of the stress pulse produced no effect on the spall stress.[VP15] The dynamic spall stress of tungsten carbide was also much greater than the quasi-static spall stress [Dandekar, 2004]. However, the presence of high hydrostatic pressures in the shock experiments is likely to cause the increasing of tensile strengths [Dandekar, 2004].[VP16] The quasi-static spall strength of silicon nitride is reported by Nahme et al. [1994] to be 0.59 0.06 GPa. This value is within the range of reported dynamic spall strengths of 0.5 GPa and 0.8 GPa [Nahme, 1994]. Nahme et al. [1994] examined the dynamic spall strength of silicon nitride, but did not systematically examine the effect of impact velocity. The high velocity shock properties of tungsten carbide were examined in depth by Dandekar [2004]. Hot pressed tungsten carbide made by Cercom that is 97.2% WC and 2.8% W2C by weight was employed. A hexagonal structure exists for both of these materials. The density was 15.530 Mg/m3, the elastic wave speed was 7.05 km/s and the elastic shear wave speed was 4.32 km/s. The average grain size was 0.9 m and the void volume fraction was 0.01. This material was subjected to planar impact experiments using a 100 mm diameter bore light gas gun. Flyer plates of the same material and of c-cut sapphire and x-cut quartz were employed. A total of seventeen experiments were carried out with maximum compressive stresses between 6.88 GPa and 24.27 GPa. The material spall strength was calculated by measuring the drop in the free surface particle velocity corresponding to the time when the two reflected rarefaction waves meet and cause State 6 (Figure 5.1). The velocity difference between the initial State 3 value and the bottom of the dip was examined. An elastic release impedance was assumed. The spall strength was observed to decline rapidly from 2.06 0.08

160

GPa at a compressive stress of 3.4 GPa to a spall strength of 1.38 GPa at a compressive stress of 7 GPa, after which the spall strength is observed to decrease more slowly. The experiments indicated an elastic spall strength of 1.22
0.45 GPa at around a compressive strength of 24 GPa. This behavior is

understood to be the result of crack blunting and shock induced plastic deformation, which decreases the weakening of the spall strength caused by compressive stress levels greater than the HEL. The experiments also indicate that there is not a large dependence on the duration of the tensile stress loading between 0.5 s and 1.1 s. Shock pressure-shear experimentation did not show a major difference in the spall strength at a skew angle of twelve degrees [Dandekar, 2004]. Nahme, Hohler, & Stilip subjected silicon nitride specimens of two different densities to both planar impact and depth of penetration experiments with the goal of increasing the available knowledge of the relationship between high speed impact behavior and material microstructure [Nahme et al., 1994]. The two silicon nitride materials had densities of 3.15 g/cm3 and 2.28 g/cm3 and had longitudinal wave speeds of 10.6 km/s and 8.6 km/s. These materials were subjected to plate impacts using steel disks in a 70 mm diameter bore gas gun. A VISARTM interferometer was employed to measure free surface particle velocities. Tungsten sinter alloy rods were used in the depth of penetration experiments, which hammered the ceramic plates into a steel backing. Among the results from the plate impact experiments was the determination of the spall strength of the materials. The spall strength was observed to be between 0.5 GPa and 0.8 GPa. Several experiments were carried out at varying velocities, but this range is the only quoted value [Nahme et al., 1994]. SEM analysis of the impacted fragments indicated that the two silicon nitrides had quite similar features at the scale of 1 micron and above[VP18]. The particles were mostly less than 1 m in diameter. Only a very few particles larger than 10 m in diameter were seen. Voids, with diameters between 10 m and 100 m

161

were understood to be the main cause for the variation in density. Areas of glassy resemblance were observed in both materials, with the denser silicon nitride having larger regions. Also, at velocities less than 400 m/s inter-granular fracture was the preferred failure mechanism, with only a very few transcrystalline cracks [Nahme et al., 1994]. The depth of penetration experiments indicated that both grades of silicon nitride offered the same level of resistance to penetration based on areal density as silicon carbide, and were only slightly worse than TiB2. Both silicon nitride grades were observed to have a greater resistance than Al2O3. This paper fails to find a single parameter that can explain the differences in dynamic and terminal ballistic behavior, but does contribute by suggesting that the similarities in microstructure above the level of a few microns may be related to the similarity in terminal impact behavior [Nahme et al., 1994].

162

B. Experimental Design

In this series of experiments both normal and combined compression and shear plate-impact experiments were conducted to investigate the HEL of Si3N4 and also to examine the effects of the state of stress on its spall strength. The normal plate impact experiments were conducted in the impact velocity range of 50 m/s to 550 m/s. The 550 m/s impact velocity with a tungsten carbide flyer plate corresponds to a predicted elastic stress of 14.8 GPa, which is above the HEL of Si3N4. In the combined compression and shear set of the plate impact experiments, a twelve degree skew angle was utilized. The impact velocity for this set of experiments was kept between 115 m/s and 300 m/s; the state of stress corresponding to this range of impact velocities and the 12 degree skew angle was such so as to not cause premature disintegration of the specimen in tension due to the pure-shear state of stress in Si3N4. These sets of experimental parameters had not been previously investigated for this grade of silicon nitride.

B.1.

Description of Materials Under Investigation

The material under study is AS800 grade silicon nitride, manufactured by Honeywell and provided by the NASA Glenn Research Center. This material was obtained in disks of 50 mm diameter that had thickness of 5.5 mm or 8 mm. These disks were impacted by either Si3N4 or tungsten carbide flyer plates. The silicon nitride has a longitudinal wave speed, CT, of 10.9 km/s and a density, , of 3.27 g/cm3 [Choi, 2002]. Multiplying the density times the wave speed gives the acoustic/longitudinal impedance, ZT, which for silicon nitride is 35.5x106 Pa/(m/s). The corresponding shear impedance is found by multiplying the density times the shear wave speed, CTS, which is 6.10 km/s. This results in a shear impedance, ZTS, of 19.9 x106 Pa/(m/s). The elastic-plastic wave speed of silicon nitride is 8.27 km/s. The corresponding elastic-plastic impedance, ZTP, is 27.1x106 Pa/(m/s). Two materials were used for these flyers: the AS800 grade silicon nitride and a CG103 tungsten carbide. This grade of tungsten carbide has 3% by weight of

163

cobalt as a binder. Tungsten carbide has a longitudinal wave speed of 6.99 km/s and a density of 15.2 g/cm3 which resulted in a longitudinal impedance, ZF, of 106x106 Pa/(m/s) [Dandekar, 2004]. The shear wave speed in tungsten carbide is 4.19 km/s, which results in a shear impedance of 63.6 x106 Pa/(m/s). This is under elastic conditions. The elastic-plastic wave speed of tungsten carbide is 5.05 km/s. The elastic-plastic impedance, ZFP, is thus 76.6x106 Pa/(m/s). For the shear experiments, a polymethylmethacrylate (PMMA) window material is employed in order to reduce the tensile stress level during refraction. This PMMA has a density of 1.19 g/cm3 and a longitudinal wave speed of 2.00 km/s. This gives it an elastic material impedance, ZW, of 2.38x106 Pa/(m/s). The shear impedance, ZWS, is 1.14 x106 Pa/(m/s), which is found from the shear wave speed of 0.96 km/s. There are two reasons why this material is used for a window. The low impedance allows a portion of the compressive wave to propagate, while the remainder is reflected from the boundary as a tensile wave. The low wave speed means that the wave propagation through the material is slow enough that the waves returning from the free surface of the window do not return during the measurement time.
B.2. Experimental Matrix for the Shock Compression Tests

Nine experiments were conducted in shock compression in order to determine the particle velocity of the free surface and thus the state of stress in the interior of the specimen (Figure 5.3). These experiments are summarized in Table 5.1. From these experiments, the Hugoniot Elastic Limit (HEL) and the spall strength as a function of shock compression could be determined. The characteristics of the dip in the particle velocity profile caused by the release waves emitted during the spall event also indicate the mode of material failure that occurs. The significance of these quantities will be discussed later. The free surface velocity records for all the experiments are summarized in Figure 5.3. In these experiments, 5.5 mm thick silicon nitride flyer plates were used to

164

impact 8 mm thick silicon nitride target plates at impact velocities of 65 m/s, 66 m/s, 107 m/s, 208 m/s, 355 m/s, and 528 m/s . These experiments were labeled SC-1 through SC-5 and SC-8. Two experiments were run at low impact speeds, 65 (SC-8) and 66 m/s (SC-5), in order to examine the velocity time profiles at stress levels close to the minimum spall strength. Besides this set of experiments, three experiments were conducted at higher impact speeds using the tungsten carbide flyer plates in order to examine the inelastic behavior of Si3N4. Tungsten carbide, because of its higher acoustic impedance when compared to silicon nitride, produces a higher compressive stress for the same impact velocity. Two of these experiments were conducted at impact velocities of 494 m/s (SC-7) and 546 m/s (SC-6) that resulted in free surface particle velocities of 708 m/s, and 792 m/s respectively. The free surface particle velocity profile for the experiment conducted at 792m/s (SC-6) indicates a kink during the rise time of the shock wave. This kink in the particle velocity indicates a transition from elastic to inelastic response during its shock compression. One experiment was conducted using a tungsten carbide flyer plate at 417 m/s (SC-9), inducing a rear surface particle velocity of 599 m/s. This experiment was performed to examine the differences in stress pulse duration between the two flyer types. The time which the initial stress wave takes to reflect off of the flyer free surface and return to the impact face is known as the pulse duration.[VP19] The silicon nitride flyers of 5.5 mm thickness and tungsten carbide flyers of 2.83 mm thickness generated wave pulses with durations of 1.03 s and 0.81 s respectively.

B.3.

Experimental Matrix for the Pressure-shear Impact Tests

A set of four plate impact experiments were conducted under combined pressureshear conditions. In these experiments tungsten carbide flyer plates of thickness 2.83 mm were used to impact silicon nitride disks of 5.5 mm thickness. The primary objective of these experiments was to examine the effects of combined

165

pressure and shear loading on the spall strength of Si3N4. The experiments were conducted at impact velocities in the range from 115 m/s to 300 m/s (Figure 5.4), and were labeled SC-10 through SC-13. One of the concerns in conducting the plate impact spall experiments under combined pressure and shear loading is the possibility of material disintegration prior to the spall event due to development of a state of pure shear within the specimen. Because the normal stress wave travels faster than the shear stress wave, there is a region after the reflection of the incident longitudinal stress pulse but before the reflection of the shear wave from the free surface of the specimen, where a part of the specimen is in a state of pure shear. This state is shown as State 4* in Figure 5.5. The state of pure shear can lead to the generation of principal tensile stresses at 45 degrees to the impact direction that are large enough to cause material failure. A Mohrs circle analysis was used to determine the maximum velocity at which the tensile stress in State 6 would exceed the expected spall strength in Si3N4 but the maximum tensile stress in state 4* would not. The stress-velocity analysis technique of the line equations works equally well for both normal and shear stresses. Therefore, from the impact velocity, the normal and shear velocities can be obtained as

V1N = VI cos , V1S = VI sin .

{V.B.1} {V.B.2}

In the above equations, the skew angle, , is 12 degrees. Using the line equations, the normal stress in state four is found in terms of the normal impedance values for the specimen, ZT, the flyer, ZF, the window, ZW, and the normal impact velocity, V1N is given as

166

Z 2 + 2Z Z 2Z Z Z 2 F T W T F 4N = T ZWV1N . (ZT + Z F )(ZT ZW )

{V.B.3}

This is a compressive stress, which is taken to be positive in this analysis. The shear stress can be determined in a similar fashion to be ZTS Z FS V1S . ZTS + Z FS

4-No Window =

{V.B.4}

In Equation {V.B.4} the shear impedances, as specified in Section B.1 are employed. Using these stress values, the maximum principal stress for the material with and without the window can be calculated as
2 X + Y Y 2 + XY . X 2 2

1, 2 =

{V.B.5}

In Equation {V.B.5} X is the normal stress, and XY represents the shear stress. The solutions to equations {V.B.3} and {V.B.4} are used for to obtain the magnitudes for X and XY, respectively. Moreover, Y is assumed to be zero for this calculation. Thus, the principal stress formula takes the form
2 X X 2 + XY . 2 = 2 2

{V.B.6}

Substituting in the normal stress value, Equation {V.B.3} and the shear stress value, Equation {V.B.4} enables the calculation of the maximum tensile stress for a given impact velocity. These values are listed in Table 5.2. They are compared with estimations of the spall strength of the material at a given normal velocity calculated using Equation {V.C.35}. When the spall strength was exceeded by the principal tensile stress, the material disintegrated in State 4*.

167

To avoid disintegration of the specimen due to tension in pure shear, a low impedance transparent window made from PMMA is glued to the rear of the specimen. The presence of the relatively low impedance PMMA window causes a part of the initial compressive wave to transmit into the PMMA window, such that the Si3N4 specimen adjacent to the PMMA window does not reach a state of pure shear. The normal stress in State 4* can still be calculated using Equation {V.B.3}. However, the shear stress in State 4* changes to Z 2 + Z Z + 2Z Z 2Z Z FS TS FS WS WS TS 4-Window = TS ZWSV1S . ZTS + Z FS )(ZTS ZWS ) (

{V.B.7}

Using the principal stress Equation {V.B.6}, the new tensile principal stress can be determined. Based on the above analysis and the estimate for the spall strength of Si3N4 under nominal shock compression conditions, an estimate for the limiting impact velocity under combined compression and shear impact conditions that does not lead to disintegration of the specimen can be made for both with and without the presence of the window. When no window is used, the principal tensile stress is larger than the spall strength at impact velocities of 250 m/s and higher. For the case of the tungsten carbide flyer, this corresponds to a maximum 350 m/s free surface particle velocity. However, in presence of the PMMA window, the principal tensile stresses are insufficient to cause spall in Si3N4 up to an impact velocity of 300 m/s and higher. This impact velocity corresponds to a 420 m/s free surface particle velocity. In view of these impact velocity constraints, the chosen impact velocities for these experiments were 115 m/s, 186 m/s, 233 m/s, and 300 m/s. The impact angle was maintained at 12 degrees in all experiments. In this way, the matrix for the pressure-shear experiments was determined.

168

C. Experimental Analysis and Results

Using the single stage gas gun, the two series of experiments described above were performed on AS800 silicon nitride. The experiments yield several results that elucidate the material behavior under dynamic tensile conditions. The equation of state and Hugoniot curve are investigated. The HEL of the material is determined by analyzing the free surface particle velocity of the target plate by impacting it to a level of shock compressive stress that is higher than the expected level for inelastic deformation in Si3N4 . From the HEL the dynamic yield stress of Si3N4 under uniaxial loading conditions can be determined. Also, by observing the spall signal in the free surface particle velocity versus time the spall strength in Si3N4 could be determined. In these experiments both the effects of increasing impact velocity and state of stress (combined pressure and shear loading) on the spall strength were investigated.
C.1. Elastic and Hydrodynamic Relationships in the Hugoniot State

Observing the state of stress of the initial shocked state of the material enables the determination of the equation of state and the Hugoniot curve. Specifically, the material shock velocity was computed and related to the particle velocity to determine the EOS of Si3N4 that is applicable to the range of the impact stresses employed in the present experiments. Additionally, the maximum compressive stress and strain are compared to obtain the Hugoniot curve. The results of these experiments offer insight into the suitability of AS800 grade silicon nitride for use in impact related aerospace applications.

C.1.1. Shock Velocity vs. Particle Velocity in the Hugoniot State

Figure 5.3 summarizes the free surface velocity profiles for all of the shock compression plate impact experiments. In examining the shock response of a material, an important parameter is the equation of state, which gives the relationship between the shock velocity and the particle velocity under hydrodynamic loading conditions. Looking at the S-V diagram for the shock

169

compression case (Figures 5.1b, 5.2b, 5.5b), the particle velocity in State 3, hereafter referred to as uP, is one-half of the rear surface particle velocity in State 4, V4 , i.e.
uP = V4 2 .

{V.C.1}

The shock velocity, US, represents the rate at which the shock propagates through the material. In principle this is the same as the elastic wave speed for impacts below the HEL. However, there is some variation (Table 5.1) due to the uncertainty in the measurement of the shock arrival times. The shock velocity is given by
U S = 2T tarrival .

{V.C.2}

The thickness of the specimen T, is measured by a micrometer prior to the assembly of the specimen and flyer. The time when the flyer impacts the specimen is the trigger time for the experiment and corresponds to the zero time on the particle velocity record. The arrival of the stress wave also represents the beginning of State 3. The arrival time, tarrival, is the difference between these two moments. Figure 5.6 shows the plot between the shock speed and the measured particle velocity obtained from the present series of experiments. From the plot it can be seen that the shock velocity is nearly a constant and is equal to the elastic wave speed of longitudinal waves in Si3N4. This is to be expected in the present study since Si3N4 remains nearly elastic in all the plate impact experiments except for the two highest impact velocity experiments conducted with the tungsten carbide flyer plates. In fact, the average of the shock velocities is 10.7 km/s, which is within 1.8% of the elastic wave speed, 10.9 km/s. This close value indicates good agreement with the wave speed of AS800 silicon nitride calculated by equation {V.A.1} from the density, elastic modulus, and Poissons ratio [Choi, 2002].

170

C.1.2. Difference in Hugoniot State Variables due to Pressure-Shear Loading

Figure 5.4 summarizes the measured surface velocity profiles of all the pressureshear plate impact experiments conducted in the present study. Because a PMMA window was employed in the pressure-shear experiments, the stresses within the material need to be evaluated keeping the new experimental configuration in mind (Figure 5.1b). The Hugoniot state, State 3 remains the same as without the PMMA window. However, the State 4 does not return to zero compressive stress. This is because the measured surface of the silicon nitride is not a free surface, but the interface between the Si3N4 and the PMMA. By calculating the intersection of the line with slope ZW, and connecting the line from State 3 to what was State 4 without the window, the new relationship between the stress and particle velocity in States 3 and 4 can be determined. The evaluation of the particle velocity in State 3 from the free surface particle velocity can therefore be carried out by using the relationship

uP = V4

(ZW + ZT )
2ZT

{V.C.3}

The rest of the calculations of the shock velocity, the Hugoniot stress and strain remain the same as for the normal shock compression case. The resulting experiments are summarized in Table 5.1.

C.1.3. Elastic Hugoniot Stress and Strain Relationship

The Hugoniot curve is a stress versus strain relationship that describes the material deformation above the HEL. It is technically, the set of all possible end states to the shock process in a given material. The material actually deforms along the Rayleigh line from the HEL to the final shocked state. In the current study, the curve cannot be determined in its entirety because only the elastic

171

regime is covered in depth by the experiments. However, an elastic stress versus strain relationship can be estimated from the data. The elastic stress and strain represent the material behavior under elastic deformation conditions. By means of the line equations, this stress can be found simply by multiplying the particle velocity, uP, by the material impedance, ZT:

E = OCT uP = ZT uP .

{V.C.4}

In Equation {V.C.4}, , is the density. The corresponding strain value is simply the ratio of the particle velocity to the elastic wave speed, CT:
E = uP CT .

{V.C.5}

The elastic compressive stresses and strains for each experiment can be determined by these equations. Combining Equations {V.C.4} and {V.C.5}, the elastic stress and strain are directly related by
2 E = OCT E = 386E (GPa).

{V.C.6}

This Hugoniot stress and strain relationship represents the loci of all the shocked states in Si3N4 under elastic compression (Figure 5.7). The resulting values are tabulated for the experiments in Table 5.1. The onset of plasticity alters the stress and strain away from the elastic case. This is discussed in more detail in Sections C.1.4 and C.1.5.
C.1.4. Calculation of the HEL

In the event that the impact velocity is high enough to cause either the flyer or the specimen to exceed its HEL, the loading shock profile shows a sharp change of slope (kink) in the free surface particle velocity diagram at a point defined by the HEL. The stress at the HEL can be determined from the corresponding free

172

surface particle velocity by multiplying by the impedance of the target plate and dividing by two, that is by reversing Equation {V.A.10}. As can be seen from Figure 5.8, in Experiment SC-6, where the free surface particle velocity reaches 792 m/s, the curve reduces slope at about 692 m/s. The elastic wave speed is used to estimate the acoustic impedance and calculate the HEL by the well established formula [Nahme et al., 1994]:
HEL = ZTV4 2 .

{V.C.7}

The HEL estimated by Equation {V.C.7} is approximately 12 GPa. This is consistent with the material behavior documented in other studies, which show similar grades of silicon nitride to have HELs of 12.1 GPa and 10-12.5 GPa respectively [Nahme et al., 1994; Mashimo, 1998]. Most of the other experiments conducted in the present study were in the elastic deformation regime and thus show no change in slope in their particle velocity versus time profile. The state of stress in Experiment SC-7 was nearly elastic, and the measured particle velocity versus time profile showed a maximum particle velocity of 708 m/s. This value of the particle velocity was too close to the particle velocity corresponding to the HEL (~692 m/s) found in Experiment SC-6 for Experiment SC-7 to be used in determining the HEL. From the HEL, the dynamic yield strength, Yo, of Si3N4 can be estimated by the well known relationship [Reinhart, 2002] 1 2 . 1

YO = HEL

{V.A.6}

This dynamic yield strength represents the strength of Si3N4 in compression under plane stress conditions beyond which plastic yielding begins, as opposed to the HEL which represents the onset of inelastic deformation under fully planestrain conditions. In Equation {V.A.6} the Poissons ratio, v, of the AS800 silicon

173

nitride is taken to be 0.27 [Choi, 2002], and therefore, the dynamic yield strength of Si3N4 can be estimated to be 7.6 GPa.
C.1.5. Determination of the Hydrodynamic Hugoniot Stress and Strain States

Because the inelastic material behavior alters the relationship between the stress and the particle velocity, Equation {V.C.7} needs to be modified for stresses above the HEL. In the present experiments, all three of the normal impact tungsten carbide flyer experiments exceed the HEL of the tungsten carbide. Also, the two highest velocity impacts exceed the HEL of the silicon nitride as established in the literature [Nahme et al., 1994; Mashimo, 1998]. However, even though both the tungsten carbide and the silicon nitride become inealstic at the highest impact velocity experiment, it can be seen (Figure 5.2) that the compression and release of the silicon nitride is always symmetric, maintaining Equation {V.C.7} as the means for determining the HEL. One method of approximating the Hugoniot stress and strain relationship of the material under elastic-plastic loading is to treat it as if it were deforming in a hydrodynamic fashion. The hydrodynamic Hugoniot stress can be experimentally determined for each experiment by using the shock velocity, US. In the present case Equation {V.C.4} takes the form
H = HEL + OU S (uP uP HEL ) .

{V.C.12}

The hydrodynamic strain at this Hugoniot state can also be found from the shock and particle velocities in State 3. The strain is in fact the ratio of the particle velocity to the shock velocity

H = uP /U S .

{V.C.13}

In Equation , O, is the initial density of Si3N4, which as mentioned above is 3.27 g/cm3.

174

This hydrodynamic stresses and strains differ from the elastic line once plasticity has occurred. This fact is reflected in the experimental results (Table 5.2). The hydrodynamic strain values for the experiments where the HEL is exceeded are plotted versus the stress to yield an equation for the relationship between them (Figure 5.7). coefficient: A relationship can be found for these two points with a correlation

H = 149H + 6.46 (GPa).

{V.C.14}

Another relationship can be determined between the HEL and the highest velocity experiment (SC-6).

H = 87.2H + 9.29 (GPa).

{V.C.15}

These stress versus strain relationships apply only within the range of the elasticplastic experiments. In order to examine the hydrodynamic curve more fully, more experiments must be performed at greater impact velocities. However, this is beyond the capabilities of the single stage gas gun.
C.2. Measurement of the Spall Strength

The wave recorded on the free surface of the target plate is a compilation of the time history of the dynamic material response (Figures 5.3, 5.4) under shock wave loading conditions. Upon reaching State 4, the free surface particle velocity is at its maximum point, Vmax. The rarefaction waves caused by the compression waves reflecting from the free surfaces of the flyer and specimen intersect within the specimen to produce a state of tension, State 6 (Figures 5.1a, 5.2a, and 5.5a). Without spall, the stress and particle velocity levels will reach State 6, as specified by the line equations. This is depicted in Figures 5.1b, 5.2b, and 5.5b as the dotted lines and the circle representing the no spall state. This stress level is not reached if the tensile stress at the spall plane is greater than the materials spall strength, in which case the material delaminates. In the case of

175

the material spall, the tensile stress can only reach the spall strength, which is indicated in Figures 5.1b, 5.2b and 5.5b by the circle showing the intersection of solid blue line in the tensile region and horizontal spall strength line. This state is known as State 6. In the absence of spall, once the intersection of the release waves propagates to the free surface of the material it completely unloads to, State 7. This state is indicated in Figures 5.1b, 5.2b, and 5.5b by the circle marked,Vno spall. If there is spall, the material unloads partially, causing a decrease in free surface particle velocity and then it reloads, which produces a reacceleration. This de-acceleration and re-acceleration produces a dip in the free surface particle velocity, from which the spall strength is calculated. The spall signal causes a decrease in free surface particle velocity from Vmax to Vmin,. This corresponds to the bottom of the dip. As the spall progresses, the free surface velocity accelerates again to another value Vo. In the shock compression experiments Vmax and Vo are the same. Thus, Vo, is not marked in Figures 5.1b or 5.2b. Using two of these three velocities, the materials spall strength, spall, can be calculated by

Spall =

1 ZT (Vo Vmin ) . 2

{V.C.16}

In Equation {V.C.16} ZT represents the specimens acoustic impedance. For the pressure-shear experiments the spall strength calculation must be altered to include the effects of the PMMA window, and is given by 1 Spall = (ZT + ZW )(Vo Vmin ) . 2

{V.C.17}

In the pressure-shear experiments, the two velocities Vo and Vmax are different. The rebound velocity, Vo, provides the more accurate measure of the spall

176

strength. This is because the material does not fail completely during the tensile pulse, but rather exhibits partial or incipient spall.
C.2.1. Derivation of the Spall Strength Formulae

In order to derive formulae {V.C.16} and {V.C.17}, relations must be found linking the spall strength to the particle velocity on the rear surface. These analyses start from the line equations that connect the stress state and particle velocity through the material impedance. The following process is then used to determine the connection between the observations of the free surface particle velocities in States 4 and 7 and the spall strength. This discussion includes the effects of a window covering the free surface of the specimen, which makes the equations valid for the pressure-shear experiments as well as the shock compression experiments. First, it can be noted from Figures 5.1b and 5.2b that the line connecting State 6 (no spall) and State 4 (Vmax) passes through the unknown spall point (Vspall, spall). Thus, the line equation for the spall strength can be written in terms of State 4 where Vmax is obtained from the experiment as

max spall = ZT (Vmax VSpall ) .


Also, using the line between State 7-s
[VP25]

{V.C.18} and State 6,[VP26] {V.C.19}

min spall = ZT (Vmin VSpall ) .

State 7-s is marked on Figures 5.1b, 5.2b, and 5.5b by, Vmin, which is known from the experiment. Next, because State 4 is on a line crossing the origin with a slope equal to the impedance of the window, ZW, the State 4 stress, max, can be written in terms of the state 4 velocity Vmax, as

max = ZWVmax .

{V.C.20}

177

The same applies to State 7-s, because it is also lies on this line, i.e.

min = ZWVmin .

{V.C.21}

Substituting Equations {V.C.20} and {V.C.21} into Equations {V.C.18} and {V.C.19}, yields ZWVmax spall = ZT (Vmax VSpall ) , ZWVmin spall = ZT (Vmin VSpall ) . Rearranging equation {V.C.22} to solve for Vspall, gives {V.C.22}

{V.C.23}

(ZT ZW )Vmax + spall


ZT

= VSpall .

{V.C.24}

Now, substituting Equation {V.C.24} into Equation {V.C.23} yields V (ZT ZW )Vmax + spall . min = ZT ZT

ZWVmin spall

{V.C.25}

Rearranging Equation {V.C.25} leads to 1 (ZW + ZT )Vmin + (ZW ZT )Vmax = spall . 2

{V.C.26}

Because Vmax is greater than Vmin, this will be a negative number, which corresponds to tension. This equation is in terms of Vmax and Vmin, both of which are known quantities from the experiments.

178

Another equation for spall, relating Vo and Vmin, can be found in the following fashion. Using the same arguments as above, the stress o in State 7-e can be found from the velocity at state 7-e,Vo, as

o = ZWVo .

{V.C.27}

Next, the velocity intercept along the line connecting State 7-e to the line between State 4, V2, will be calculated. The intercept is referred to as Point A. Because this point is on the velocity axis, the stress, A, is zero. equation uses the elastic impedance of the specimen, ZT:[VP27] The line

A + ZTVA = o + ZTVo .

{V.C.28}

Since the stress at Point A is zero and substituting Equation {V.C.27} for the stress at State 7-e o, yields
ZTVA = (ZW + ZT )Vo .

{V.C.29}

Also, using the line equations to connect Point A to the spall plane, yields ZTVA = spall ZTVspall . Substituting Equation {V.C.29} into Equation {V.C.30}, yields
(ZW + ZT )Vo = spall ZTVspall .

{V.C.30}

{V.C.31}

Using equation {V.C.23}, which provides the relationship between, Vmin, and the spall strength and velocity, Equation {V.C.23} can be rearranged to give
spall + ZTVspall = (ZW + ZT )Vmin .

{V.C.32}

Adding equations {V.C.32} and {V.C.31} yields,

179

2spall = (ZW + ZT )Vmin (ZW + ZT )Vo .

{V.C.33}

Finally, simplifying Equation {V.C.33}, gives 1 spall = (ZW + ZT )(Vo Vmin ) . 2

{V.C.34}

This is a second equation that can be used to calculate the spall strength. Again, the answer is negative because tension is negative on the S-V diagram. This equation is equivalent to equation {V.C.17}. Note that in the case of the normal impact experiments, the window impedance is zero, and Equation {V.C.34} reduces to Equation {V.C.16}.

C.2.2. Spall Strength and Material Failure Mode during Spall in Pure Shock Compression

SiN exhibited spall in all experiments conducted in this series of experiments and a variation of the spall strength with impact velocity was recorded (Table 5.3). As per the discussion in Section C.3.1, the spall strength can be directly related to the spall dip (Figure 5.3) in the free surface particle velocity profile as 1 ZT (Vo Vmin ) . 2

Spall =

{V.C.16}

In the case of the normal loading in the elastic material range, Si3N4 indicates a drop in the spall strength with increasing impact velocity (Figure 5.9), and the spall strength is observed to decrease from an average of 895 MPa at 65 m/s (SC-8) to 564 MPa at 528 m/s (SC-3). Statistically, ceramic materials possess a large scatter in the material failure data. This variation can also be seen in the present experiments, where spall strengths of 908 MPa and 881 MPa are recorded in two experiments at impact velocities of 65 m/s (SC-8) and 66 m/s (SC-5), respectively. A linear fit to the spall strength data can be represented as

180

spall = -0.0383H + 109 (Pa) ,

{V.C.35}

This linear fit is plotted along with the experimental data in Figure 5.9. The linear fit line had a correlation coefficient of R2 = 0.976. It is interesting to note that the experimental spall strengths were very close to the predictions of the straight line, especially given the fact that ceramic materials tend to have large statistical scatter. As the material behavior becomes inelastic under compression prior to spall, the spall strength reaches a plateau. This effect is seen clearly in the two experiments with the highest impact velocity. In these experiments, the spall strengths were 640 MPa at 494 m/s (SC-7) and 622 MPa at 546 m/s (SC-6) (Figure 5.9). These two values are only 18 MPa apart from each other, which is smaller than the scatter of 27 MPa between experiments SC-5 and SC-8, both conducted at 65.5 0.5 m/s. The near constant level of spall strength in experiments SC-7 and SC-6 suggests that the applied compressive stresses results in some degree of inelasticity in Si3N4 that retards the spall initiation. The pull-back time, which is the time that the velocity takes in recovering from Vmin to VO, changes distinctly with impact velocity (Figure 5.3). In the lowest impact velocity experiment, at 65 m/s (SC-8), this recovery time is about 90 ns. At an impact velocity of 107 m/s (SC-1) the pull-back time drops to 30 ns, which suggests a more brittle type of material failure. The increase in the recovery time with an increase in impact velocity indicates a higher degree of the accumulation of damage during the spallation process. In experiments conducted at the higher impact velocities (528 m/s for Experiment SC-3) the pull-back time was again observed to be 90 ns. Within the four experiments using the tungsten carbide flyers, the pull-back time is also seen to increase. The impact velocity in Experiment SC-9 is 419 m/s and the pull-back time is 80 ns. By an impact velocity of 546 m/s (SC-6) in tungsten

181

carbide, the pull-back time has expanded to 130 ns. This, along with the increase in pull-back time with impact velocity in the experiments with silicon nitride flyers, confirms the trend of increasing pullback time with velocity. This increase in pull-back time suggests a damage accumulation mode that increases in duration with increasing impact velocity rather than an instantaneous brittle fracture. An interesting feature is observed in the effect of stress pulse loading time on the pull-back time. The impact velocity in Experiment SC-9, 419 m/s is lower than in Experiment SC-3, but a tungsten carbide flyer was used, so the compressive stress in Experiment SC-9 is higher than in Experiment SC-3, With a higher compressive stress, the pull-back time in Experiment SC-9 should be higher than in Experiment SC-3, However, in Experiment SC-9, the pull-back time is 80 ns. This is slightly lower than in Experiment SC-3, where the pull-back time is 90 ns. The reason this is occurs is that in the experiments with the tungsten carbide flyer, a shorter compressive pulse duration, i.e. 0.81 microseconds instead of 1.03 microseconds, was attained in the experiments. This difference in pulse duration was due to the differences in the impedance and thickness of the WC flyer when compared to the Si3N4 flyer used in the other spall experiments. Thus, the noticed decrease in pull-back time between experiments SC-3 and SC-9 is due to the shortening of the compressive stress pulse loading time.

C.2.3. Spall Strength and Failure Modes Under Combined Pressure and Shear Impact Loading

The shock compression-shear experiments resulted in substantial degradation of the spall strength at the relatively small angle of obliquity (12 degrees). This degradation of the spall strength is seen by the change in the level of the free surface particle velocity, following the spall dip (Figure 5.4). The presence of a PMMA window results in the velocity pull-back following the spall not recovering to the level of Vmax. This creates a distinction between the

182

pre-spall peak velocity Vmax, and the post-spall peak velocity Vo, which was not evident in the pure shock compression experiments (Figure 5.1b). This being the case, the relation for the spall strength in terms of the pull-back particle velocity Vo, and the minimum velocity Vmin, is 1 Spall = (ZT + ZW )(Vo Vmin ) . 2 {V.C.17}

Table 5.3 provides the details of the experiments and the experimental results obtained from the pressure-shear experiments conducted in the present study. When compared to the plate impact experiments, the spall strength in the presence of a modest amount of shear is observed to drop more rapidly with an increase in impact velocity. The spall strength decreases from 803 MPa at an impact velocity of 115 m/s (SC-11) to 249 MPa at an impact velocity of 233 m/s (SC-12) (Figure 5.10). As the velocity is increased still higher, a complete loss of spall strength in Si3N4 is observed. This is evidenced by the lack of spall recovery in the 299 m/s impact velocity experiment (SC-13). As with the shock compression data, a linear fit was determined for the pressureshear spall strength data
spall = -0.1751H + 109 (Pa) .

{V.C.36}

This fit has a correlation coefficient of R2 = 0.989. As mentioned in Section C.2.2 the R2 value represents a good fit considering the inherent uncertainty in material failure data for ceramics. The slope of the linear fit is almost five times as steep as that of the slope for the shock compression experiments. Also, note that the highest velocity experiment has no spall strength. This can be determined from Figure 5.4, because following the drop in free surface particle velocity due to the spall, there is no pull back in the particle velocity. The exact compressive stress for which the spall strength becomes zero is 5.71 GPa. This is between Experiment SC-13, 299 m/s, and Experiment SC-12, 233 m/s.

183

The steeper drop in slope in Equation {V.C.36} when compared with Equation {V.C.35} indicates that the shear loading creates a severe degradation of the spall strength even at a modest skew angle of twelve degrees (Figure 5.10). It is evident that this shear loading causes the material to lose its strength because the material is damaged in shear prior to the spall. The shear stress, while not enough to cause complete failure in shear, does result in damage in the material, making it easier to spall. Pull-back times during the spall are about 100 ns for the lower impact velocity experiments (SC-10, SC-11), and could not be determined for the other experiments (SC-12, SC-13).

C.3. Material Failure Modes during Dynamic Spall and Fractography of the Spall Surface

The recovered silicon nitride target and the flyer materials were heavily fragmented in all the experiments. The expansion chamber makes possible the soft recovery of these fragments. Moreover, the nature of this fragmentation is altered with the impact velocity and the stress state. At low impact velocities around 65 m/s (SC-8, SC-5), the specimen breaks up along the spall plane and in a star shaped pattern generated by stresses resulting from the boundary waves (Figure 5.11a). The spall plane is directly observable in the fragments from these experiments. At higher impact velocities, such as 107 m/s (SC-1), the specimen breaks into smaller fragments (Figure 5.11b). At still higher impact velocities beyond 356 m/s, the material completely disintegrates, becoming essentially a powder (Figure 5.11c). In order to examine the micro-structural effects of impact velocity and the state of stress on the material, fragments with a face that is a surface of the spall plane were examined under the scanning electron microscope (SEM). Four representative experiments were evaluated. To observe the microstructure in low and high impact velocities, experiments SC-8, SC-2, and SC-4 were chosen. These have impact velocities of 64.9 m/s, 208 m/s, and 355 m/s, respectively.

184

To study the effect of angle, Experiment SC-12 was examined. Experiment SC12 was conducted at an impact velocity of 152 m/s. For the low impact velocity experiment conducted at 65 m/s (SC-8), which corresponds to a particle velocity of 31.8 m/s in the material, the SEM reveals a relatively smooth fracture surface with several large intact hexagonal grains (Figure 5.12). The Hugoniot state particle velocity is relevant here because that is the velocity of the material at the spall plane prior to the arrival of the spall waves. These grains are at their largest greater than 100 m long. At higher magnification of 2000X and 3500X, the fracture mode is seen to be predominantly inter-granular fracture. However, some intragranular fracture through the larger grains is also detected. At 208 m/s impact speed, which corresponds to a particle velocity of 178m/s (SC-2), the overall fracture surface appears to have become rougher (Figure 5.13). No grains of 100 m size were detected, but several holes for grains of about 50 m can be seen. These grains are imaged at 2000 times magnification as well. A larger fraction of the 50 m grains are fractured. This image shows the fracture surface to be rougher on all scales but still brittle; both intergranular and intra-granular fracture modes were observed. Shot SC-4, which has an impact speed of 356 m/s and a particle velocity of 178 m/s, shows two distinct regions of fracture, one rough and one smooth (Figure 5.14). More highly magnified images show a larger grain sticking out of a clump of smaller grains in the rough fracture area. The smooth area is smooth even at 3500 times magnification. The fracture mode in the rough zone is still brittle, though it is difficult to ascertain if it is intra-granular or inter-granular. The smooth zone is likely due to inelastic failure processes. The effect of skew angle on the modes of fracture during spallation can be examined from the recovered specimens in Experiment SC-12 (Figure 5.15). This experiment has an impact velocity of 233 m/s with a skew angle of 12 degrees;

185

the corresponding normal component of the particle velocity in Si3N4 was 152 m/s. The overview at 250 times magnification shows a brittle fracture surface, similar to that in SC-2 (Figure 5.13). A few grains on the order of 50 m are seen. Closer views reveal grains on the order of 10 m long sticking out of the fracture surface, mostly intact. These images can be compared with the SEM pictures of the microstructure of the intact Si3N4 ceramic taken from the surface of a sample that was not impacted (Figure 5.16). The examined face was polished. This surface shows the granular structure of the material at 2000 times magnification. Holes in the surface show where large grains, on the order of 10 m have been pulled out of the material during the polishing process. The SEM images indicate that spallation occurs by a primarily planar intergranular brittle fracture mode at the low impact velocities of ~ 60 m/s. 100 m grains are visible, many clearly intact. However, some of these grains are fractured. At the higher impact velocities, on the order of 200 m/s, the surface becomes rougher and several 50 m grains are no longer intact, suggesting both inter-granular and intra-granular fracture modes to be active. At around 350 m/s impact velocity, the material shows both brittle fracture and smooth areas. These smooth areas could represent inelastic failure processes beginning to occur. At higher velocities, the material is reduced to powder, and so no SEM analysis was possible. The presence of shear stresses in the 200 m/s impact is not reflected by a change in failure mode during spallation. No additional surface features in the shear experiment are noticed and the fracture mode is primarily brittle. This suggests that the addition of shear does not cause a different type of microstructural damage.

186

D. Summary of Planar Results

The shock compression and shock compression shear experiments discussed in this chapter examine the dynamic behavior of silicon nitride using one dimensional wave theory. The elastic and hydrodynamic Hugoniot and the elastic equation of state of the material are observed in the range of impact velocities up to and slightly above the Hugoniot Elastic Limit. The dynamic yield strength of the material is evaluated from the HEL. The spall strength under shock compression and shock compression-shear loadings is also described. The elastic equation of state of silicon nitride was determined, and the portion of the elastic-plastic regime near the HEL was also examined. The average level of the elastic shock velocity, 10.7 km/s, is reasonably close, 1.8%, to the elastic wave speed of the material, 10.9 km/s (Figure 5.6). Likewise, the elastic Hugoniot of the material, or the stress-strain relationship is found for the elastic regime. The hydrodynamic Hugoniot equation, which is used above the HEL, indicates divergence from the elastic equation as the material begins to deform plastically (Figure 5.7). These relationships enable the prediction of the stresses, strains, and shock velocities within the impact velocity range of the experiments. The Hugoniot Elastic Limit can be determined from the VISAR records of the experiments that exceed the predicted HEL from the literature. The HEL is found by examining a sudden kink, a decrease in the loading rate of the initial rise in free surface particle velocity. This indicates the onset of plastic deformation. By observation, the HEL is found to be 12 GPa, which concurs with previously recorded HELs for similar materials of 12.1 GPa and 10-12.5 GPa [Nahme et al., 1994; Mashimo, 1998]. From the HEL the dynamic yield strength in compression can be found to be 7.6 GPa. The compressive elasticplastic loading results in a residual compressive stress within the material. The spall strength is determined by the magnitude of the drop and pull-back in velocity observed in all of the experiments at the time when waves reach the free surface from the intersection of the two tensile rarefaction waves. The

187

experimental Formula {V.C.17} uses the minimum velocity, Vmin, which corresponds to State 7-s and the re-acceleration velocity, Vo, which corresponds to State 7-e. Each experiment has a unique spall strength, which varies both with impact velocity and with the addition of a skew angle of 12 degrees. The pull-back time from the spall increases with increasing velocity, except for the experiments at impact velocity of about 65.5 0.5 m/s (SC-8, SC-5) in which the pull-back times are larger than the pull-back time for experiment SC-1 at an impact velocity of 107 m/s. This indicates that insufficient stress exists in experiments SC-8 and SC-5 to cause sudden brittle fracture. Above 66 m/s, sudden brittle fracture occurs. However, as the velocity increases in the experiments with silicon nitride flyers, the accumulation of damage process resulting in spall becomes slower. Within the group of experiments with tungsten carbide flyers, the pull-back time also increases with increasing impact velocity. However, the alteration of the flyer material from silicon nitride to tungsten carbide and the resulting decrease in compressive stress pulse loading time causes a decrease in the pull-back time. The spall strength under elastic shock compression drops 39% between experiments conducted at impact velocities of 65.5 0.5 m/s (averaging SC-8, SC-5) and the experiment conducted at an impact speed of 528 m/s (SC-3) (Figure 5.9). This drop indicates that with increasing severity of impact stress, the material has lesser ability to resist spall. This indicates that inelastic microcracking is occurring, damaging the silicon nitride even below the HEL. The correlation between spall strength and impact velocity, Equation {V.C.35} indicates a difference in spall strength with velocity that is substantially larger than the scatter expected of a ceramic material. This scatter can be quantified by the 27 MPa difference in the two experiments conducted at an impact velocity of 65.5 0.5 m/s (SC-8, SC-5). Under plastic deformation, the material exhibits a near constant spall strength. In fact, in the two elastic-plastic experiments (SC-6, SC-7) the spall strength only changes by 18 MPa, which is 2.8% of the average spall strength of 631 MPa, and

188

less than the scatter in experiments SC-5 and SC-8. From this, it appears that the compressive residual stress generated by the elastic-plastic loading retards the initiation of spall. The effect of velocity is noticeable at the microscopic level in the SEM images. The planar brittle fracture at the spall plane at an impact velocity ~60 m/s gives way to rough brittle fracture at and impact speed of ~ 200 m/s and to the beginnings of non-elastic fracture at an impact speed of ~350 m/s and higher velocities. The addition of a twelve degree skew angle to the shock compression impacts results in a pressure-shear loading which causes the material to exhibit more severe degradation of spall strength (Figure 5.10). The spall strengths in these experiments are nearly comparable to the shock compression experiments at lower velocities. At an impact velocity of 115 m/s (SC-11), the spall strength is 803 MPa, which is 10% off of 893 MPa, the calculated value from Equation {V.C.35} for the normal spall strength at this impact velocity. However, at an impact velocity of 233 m/s (SC-12), the spall strength is only 249 MPa, which is 68% smaller than the calculated normal spall strength of 780 MPa for this impact velocity. At an impact velocity of 299 m/s (SC-13), the spall strength disappears altogether. Clearly, the addition of even a mild skew angle causes severe spall strength degradation. Although the material does not fail in shear, it does experience brittle mode damage and therefore strength reduction prior to the arrival of the spall wave. This has the opposite effect as the residual compressive stress generated by plasticity. However, the microscopic damage patterns as recorded by the SEM images at the spall plane remain the same as in the shock compression impact case at about 200 m/s impact velocity. This indicates that the same fracture modes are present in shock compression and pressure-shear.

189

To summarize, AS800 grade silicon nitride exhibits a high HEL of 12 GPa. The spall strength decreases under shock compressive loading in the elastic regime. This indicates that some damage is accumulating below the HEL. However, under elastic-plastic loading, the spall strength remains roughly constant, suggesting a moderating trend in the elastic spall strengths decrease with increasing impact velocity due to inelastic shock compression. The high HEL and moderating influence of inelastic shock compression favor AS800 silicon nitride as a potential aircraft engine turbine blade material. Pressure-shear loading, on the other hand causes a five times greater decrease in the spall strength over the same elastic region compared with the shock compression experiments. This strong effect of pressure-shear loading on spall strength must be considered a drawback to using AS800 silicon nitride.

190

References
Achenbach, J. D. 1975. Wave Propagation in Elastic Solids. Elsevier Science Publishers. Boslough M. B. and J. R. Asay. 1993. Basic Principles of Shock Compression. High-Pressure Shock Compression of Solids. Springer-Verlag. 7-42. Choi et al. 2002. Foreign Object Damage of Two Gas-Turbine Grade Silicon Nitrides by Steel Ball Projectiles at Ambient Temperature. NASA/TM-2002-211821. Dandekar, D. 2004. Spall Strength of Tungsten Carbide. U.S. Army Research Laboratory. ARL-TR-3335. Dremin A. N.; A. M. Molodets. 1990. Metal Spall and Fracture Mechanisms. Shock Compression of Condensed Matter. Elsevier Science Publishers. 415-420. Grady, D. E. 1988. The Spall Strength of Condensed Matter. J. Mech. Phys. Solids Vol. 36 No. 3 pp 353-384. Mashimo, T. 1994. Shock Compression of Ceramic Materials: Yielding Property. American Institute of Physics. Mashimo, T. 1998. In: High-pressure shock compression of solids III. L. Davison, M. Shahinpoor, editors New York : Springer 101-146. Nahme, V. Hohler, A. Stilp. 1994. Determination of the Dynamic Material Properties of Shock Loaded Silicon-Nitride. American Institute of Physics. 765-768. Reinhart, W. D.; L. C. Chhabildas. 2002. Strength Properties of Coors AD995 Alumina in the Shocked State. International Journal of Impact Engineering. 29 [1-10] 601-619.

191

Tables
Table 5.1: Experimental results for Hugoniot state shock velocities, US, particle velocities, uP, elastic stress, E, and strain, E, and hydrodynamic stress, H, and strain, H. The table shows the values determined using Equations {V.C.1} through {V.C.6}.
Exp. # SC-1 SC-2 SC-3 SC-4 SC-5 SC-6 SC-7 SC-8 SC-9 SC-10 SC-11 SC-12 SC-13 Impact Velocity 107 m/s 208 m/s 528 m/s 355 m/s 66.0 m/s 546 m/s 494 m/s 64.9 m/s 417 m/s 186 m/s 115 m/s 233 m/s 299 m/s Flyer Type SiN SiN SiN SiN SiN WC WC SiN WC WC WC WC WC 0 0 0 0 0 0 0 0 0 12 12 12 12 1.21 km/s 1.02 km/s 1.10 km/s 9.93 km/s 8.52 km/s 8.57 km/s 9.14 km/s 1.07 km/s 9.60 km/s 1.15 km/s 1.13 km/s 1.74 km/s 9.27 km/s 53.1 m/s 100 m/s 251 m/s 178 m/s 31.5 m/s 394 m/s 353 m/s 31.8 m/s 298 m/s 118 m/s 73.2 m/s 152 m/s 195 m/s 1.88 GPa 3.56 GPa 8.91 GPa 6.32 GPa 1.12 GPa N/A N/A 1.13 GPa 10.5 GPa 4.20 GPa 2.60 GPa 5.38 GPa 6.93 GPa 4.89E-3 9.22E-3 2.31E-2 1.64E-2 2.90E-3 N/A N/A 2.93E-3 2.74E-2 1.09E-2 6.74E-3 1.39E-2 1.80E-2 N/A N/A N/A N/A N/A 13.3 GPa 12.2 GPa N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A 4.60E-2 3.86E-2 N/A N/A N/A N/A N/A N/A US uP E E H H

Table 5.2: Mohr's Circle Stress Calculations. The spall strength as determined using Equation {V.C.35} is contrasted with the tensile principal stress in State 4* from Equation {V.B.6}, which is pure shear without a low impedance window. For the experiment not to fail in shear, the spall strength must be greater than the tensile principal stress with the PMMA window, which uses the same equation, but a different shear stress. For it to fail at the spall plane, the spall strength must be smaller than the maximum tension in State 6. Principal Tension in Impact Velocity (m/s) 100 150 200 250 300 350 Free Surface Velocity (m/s) 140.41 210.61 280.82 351.02 421.23 491.43 Spall Strength (Pa) 8.8470E+8 8.4805E+8 8.1140E+8 7.7475E+8 7.3810E+8 7.0145E+8 State 4*, no window (Pa) 3.16E+08 4.73E+08 6.31E+08 7.89E+08 9.47E+08 1.10E+09 Principal Tension in State 4*, window (Pa) 2.10E+08 3.14E+08 4.19E+08 5.24E+08 6.29E+08 7.37E+08 Maximum tension in State 6 (Pa) 1.91E+09 2.86E+09 3.81E+09 4.76E+09 4.83E+09 6.59E+09

192

Table 5.3: Experimental results for spall stress and pull-back time. The impact velocity, skew angle, , and free surface particle velocities, V4, are used to index the data.
Exp. # SC-8 SC-5 SC-1 SC-2 SC-4 SC-3 SC-9 SC-7 SC-6 SC-11 SC-10 SC-12 SC-13 Impact Velocity 65 m/s 66 m/s 107 m/s 208 m/s 355 m/s 528 m/s 417 m/s 494 m/s 546 m/s 115 m/s 186 m/s 233 m/s 299 m/s Flyer Type SiN SiN SiN SiN SiN SiN WC WC WC WC WC WC WC 0 0 0 0 0 0 0 0 0 12 12 12 12 64.6 m/s 64.9 m/s 107 m/s 208 m/s 356 m/s 502 m/s 599 m/s 708 m/s 792 m/s 147 m/s 239 m/s 306 m/s 396 m/s V4 Spall Stress 908 MPa 881 MPa 890 MPa 835 MPa 709 MPa 572 MPa 564 MPa 640 MPa 622 MPa 803 MPa 557 MPa 249 MPa 0 MPa Pull-Back Time 0.082 s 0.100 s 0.032 s 0.052 s 0.062 s 0.086 s 0.081 s 0.14 s 0.13 s 0.11 s 0.09 s N/A N/A

193

Figures

7,7 6, 6 5 3 1 2 4

Figure 5.1: Calculation plots for elastic compression impact experiments. (a) Time versus distance diagram. The longitudinal stress waves propagate outwards from the point of impact, reflect from the free surfaces as tensile rarefaction waves, and intersect defining the spall plane. On this diagram, the locations of the various stress states are marked. (b) Stress versus velocity diagram. The red line indicates the tungsten carbide which (WC) exceeds its HEL. The blue line indicates the silicon nitride. The lines show the paths between the stress states indicated in (a). The spall strength is measured from velocities on the particle velocity axis and Equation {V.C.16} is based on the relationships in this plot.

Figure 5.2: Calculation plots for elastic-plastic compression impact experiments. The configuration is the same as in Figure 5.1. (a) Time versus distance diagram (b) Stress versus velocity diagram. The HEL of silicon nitride is exceeded and the two lines used to estimate the Hugoniot state (State 3) are depicted. Even with plastic deformation, the loading and unloading of the silicon nitride is symmetric.

194

0.8

SC-6

Free Surface Particle Velocity (km/s)

0.7 0.6 0.5 0.4

SC-7 SC-9 SC-3

SC-4
0.3 0.2 0.1 0

SC-2 SC-1 SC-5, SC-8


0 1 2 3

Time (s)
Figure 5.3: Shock compression free surface particle velocity history plots. The initial rise is due to the arrival of the compressive wave at the free surface. From the plot of SC-6, the HEL of silicon nitride can be seen. The spall signal arrives at the free surface from the intersection of the tensile rarefaction waves at a time determined by the thickness of the flyer and its impedance. The pullback time can be seen to change with free surface particle velocity, which is directly related to impact velocity.

195

400

SC-13
350

Free Surface Velocity (m/s)

300 250 200 150 100 50 0

SC-12 SC-10

SC-11

Time (s)
Figure 5.4: Pressure shear experiment silicon nitride-PMMA interface surface particle velocity history plots. The velocities before and after the spall signal are distinct. As the impact velocity, and thus measured surface particle velocity, increases, the pull-back velocity decreases.

4*

Figure 5.5: Calculation plots for compression-shear impact experiments. (a) Time versus distance diagram. State 4* is shown. It is a state of pure shear following the reflection of the compressive stress wave. (b) Stress versus velocity diagram. The green line indicates the PMMA window.

196

15 14 13 12 11 10 9 8 7 6 5 4 3 2 1 0 0 0.1 0.2 0.3 0.4 0.5

Shock Velocity (km/s)

Elastic Wave Speed 10.9km/s

Particle Velocity (km/s)


Figure 5.6 The shock velocity plotted versus the Hugoniot state particle velocity. as the particle velocity increases, the shock velocity decreases. In the elastic regime, the average shock velocity of 10.7 km/s matches with the elastic wave speed. However,

197

15 14 13 12 11 10

Stress (GPa) = 87.2*Strain + 9.29 HEL


49 n rai St *

.4 +6

Stress (GPa)

Stress (GPa) = 386*Strain


G s( es r St

9 8 7 6 5 4 3 2 1 0 0 0.01 0.02

Pa

1 )=

Elastic Hydrodynamic 0.03 0.04 0.05

Strain
Figure 5.7: Plotting the Hugoniot elastic and hydrodynamic stresses and strains. Above the HEL, the hydrodynamic predictions diverge from the elastic prediction. Two correlations are offered, one from the HEL to the highest velocity experiment (SC-6), and one connecting both experiments above the HEL.

198

Free Surface Particle Velocity (km/s)

0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0.8 0.9 1 1.1 1.2

V FS at HEL

Time (s)
Figure 5.8: Rear surface velocity history plot for 546 m/s impact velocity (SC-6) experiment showing the decrease in slope of the velocity rise which indicates the Hugoniot Elastic Limit. The velocity at which this occurs is indicated on the plot.

199

1000 900 800

(a) Sp all

-0

Normal SiN-SiN Normal WC-SiN

.0

38

3*

Spall Strength (MPa)

700 600 500 400 300 200 100 0 0 0.1 0.2 0.3

St re ss +

1E

+0

9(

Pa

0.4

0.5

0.6

0.7

0.8

0.9

Free Surface Particle Velocity (km/s)


Figure 5.9: Spall strength plotted versus free surface velocity for normal impact experiments. In the elastic regime, there is a linear relationship between stress and free surface particle velocity. However, with the onset of elastic-plastic deformation, the spall strength plateaus.

200

1000 900 800

Shear WC-SiN

S pa

Spall Strength (MPa)

700 600 500 400 300 200 100 0 0 0.1

ll = 0 .1 751 +1 ess * S tr E+0 9 (P a)


0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Free Surface Particle Velocity (km/s)


Figure 5.10: Spall strength plotted versus free surface velocity for pressure shear impact experiments. The lower three velocity experiments show a linear trend. However, the highest velocity experiment (SC-13) had no pull-back following the drop in free surface particle velocity that indicates spall (Figure 5.4). Therefore, the spall strength of experiment SC-13 was negligible.

201

Figure 5.11: Post impact fragment images. The scale is in inches. (a) 65 m/s impact speed (SC8). The specimen is still contained within the specimen ring. The spall plane and star shaped fracture are visible. (b) 107 m/s impact speed (SC-1). The fragments are smaller than those in (a). (c) 355 m/s impact speed (SC-4). The debris is mostly powder, but a few fragments of sufficient size to use the SEM remain.

202

Figure 5.12: SC-8, 31.8 m/s Hugoniot particle velocity. Top left: overview at magnification 250 showing brittle fracture and several 100 m hexagonal grains and many smaller ones. Top right: magnification 2000 showing intra-granular brittle fractures of two large grains in different orientations. Bottom left: magnification 3500 with differing size grains and brittle fracture. Bottom right: magnification 3500 with a hole where a large grain has been removed.

203

Figure 5.13: SC-2, 100 m/s Hugoniot particle velocity. Left, overview at 250 times magnification showing no 100 m grains, but some 50 m grains. Right: 2000 times magnification showing fracture of 50 m smaller grains.

Figure 5.14: SC-4, 178 m/s Hugoniot particle velocity. Top left: overview at 250 X showing two distinct regions, rough and smooth, and no visible large grains. Top right: 2000 X view showing rough region, but no grains as large as 50 m. Bottom left: Close up at 3500 X of large grain in rough region. Bottom right: Close up at 3500 X showing smooth region with no large grains.

204

Figure 5.15: SC-12, 152 m/s Hugoniot particle velocity. Top left: Magnification 250 Overview showing rough brittle fracture, same as SC-2. Top right, 2000 X view of ~50 m grains and brittle fracture surface. Bottom left: 3500 times magnification showing fractured large grains.

205

Figure 5.16: Polished surface of specimen, pre-impact. Top left: 250 X overview, showing no distinguishing features. Top right: magnification 2000 showing polished grain faces and a hole where a ~10 m grain was removed by the polishing process. Bottom left: 3500 times magnification showing polished ~ 20 m grain. Bottom right: 3500 X showing grain hole on the order of 10 m.

206

Chapter VI The Effect of Shock Re-Shock, Shock Unloading, and Stress Reverberations on the Behavior of Silicon Nitride
The propagation of a shock wave through a material can cause inelastic deformation that directly affects material properties such as the dynamic yield strength. The material subjected to a subsequent shock or series of shocks may behave differently from unshocked material. In the present study plate impact shock compression experiments are employed using a dual layer flyer plate to determine the shock-release and shock-re-shock response of silicon nitride. By studying the off-Hugoniot stress states generated in experiments of this type, the material strength behavior of the shocked material can be examined [Asay and Chhabildas, 1981; Reinhart and Chhabildas, 2003]. The determination of the dynamic yield strength depends on inelasticity in both the re-shock and release experiments. A second set of experiments is conducted using a thick flyer of silicon nitride and a thin target of tungsten carbide. Multiple unloading states are generated in the flyer as stress waves reverberate in the target. Using equations developed by Hall et al. [Preliminary], the progressive unloading of stress in the silicon nitride flyer is examined. These experiments also provide observations of the materials behavior in the shocked state.. The theoretical background for these experiments and their applications to specific materials are reviewed in Section A. The assumptions of both elastic and inelastic deformation underlie the specific equations that describe the material behavior in these experiments. Also, equations that use the inelasticity of the material to determine the dynamic yield stress are provided. In Section B, the design of experiments is presented. The two types of experiments discussed in Section A are modified for the silicon nitride experiments in this study. Following this, in Section C, the VISAR results are analyzed and the material response is discussed. The experimental results are restricted by the maximum impact velocity achievable in a single stage gas gun. Within this range of impact

207

velocities, the analysis indicates a strong agreement between the particle velocity and stress state predictions of the equations in Section A and the experimental free surface particle velocities presented in Section C.

208

A. Theory and Computations for Multiple Shock Experiments Observing the effects of an additional shock to an already shocked material can provide insights into inelastic damage within the material generated by the first shock. A dual layer flyer can be utilized to provide a shock re-shock or a shockunload condition. This method has been detailed in Section A.1. The basis for this method is explained in two papers -- Asay and Chhabildas [1981] and Reinhart and Chhabildas [2003]. Additionally, with a shock reverberation plate impact experimental configuration, a series of unloading waves can be designed to propagate through the flyer. From this type of experiment, the release behavior of the material can be studied [Dandekar et al., 2003]. This method is explained more mathematically in the paper by Hall et al. [Preliminary]. A.1. Shock Re-shock and Shock Unload Experiments A method for studying the dynamic behavior of a material involves shocking the material twice-- once to load it to a predetermined shock state and then again shock it either to a higher or lower stress state. From these experiments the dynamic material strength of a material in the shocked state can be evaluated without recourse to the knowledge of the materials dynamic hydrostat. This is accomplished by employing flyers with two layers that have differing impedances. The result of using a higher impedance backing material is that the second wave has a higher compressive stress level, which causes re-shock in the material. A lower impedance backing plate reduces the stress, causing unloading after the initial shock. By employing matching impact velocities in both of these experiments the dynamic yield strength as well as the shear stress in the shocked stste can be estimated [Reinhart and Chhabildas, 2003][Asay and Chhabildas, 1981]. A study on alumina that was performed by Reinhart and Chhabildas [2003] demonstrates the dual flyer technique. Through the use of a pair of experiments, one shock-reshock and one shock-unload, the dynamic yield strength of the alumina was determined. Aluminum oxide (Al2O3) with a density of 3.89 g/cc

209

and with longitudinal and shear wave speeds of 10.56 km/s and 6.24 km/s respectively was employed. A two stage light gas gun with a 28 mm bore was employed to accelerate the specimens to about 8 km/s. These experiments employed a flyer backed with either a higher or lower impedance material, and a window backing the specimen. For the window, lithium fluoride was used. Polymethylpentene or TPX was used as the low impedance flyer backing material for the shock release experiments. Tantalum and copper were used to back the flyer in the shock re-shock experiments. The resulting shock waves created off-Hugoniot conditions which are used to calculate the dynamic hydrostat, the shear stress in the shocked state, and the maximum and the minimum stress levels, and the dynamic material strength [Reinhart and Chhabildas, 2003]. From the knowledge of te impact velocity and the free surface velocity versus time histories, the Hugoniot stress versus strain states can be examined. Using conservation laws, the stress and strain Hugoniot states are given by

H = HEL + OU SP ( uH uHEL ) , H = uHEL CL + ( uH uHEL ) U SP .

{ .A.1} { .A.2}

In Equations { .A.1} and { .A.2}, the Hugoniot stress, strain, and particle velocity are represented by, H, H, and, uH, respectively. The stress and velocity at the HEL are, HEL, and uHEL. CL represents the Lagrangian wave speed and, USP, the plastic shock velocity. These values are determined from the initial velocity jump corresponding to the first shock. One important note about these equations is that the stress in the initial Hugoniot state exceeds the Hugoniot Elastic Limit of the material [Reinhart and Chhabildas, 2003]. Once the second shock has propagated, the material either unloads or loads further. At the point where the elastic wave gives way to the bulk wave, the

210

second state can be defined for each of these experiments. By using incremental forms of the conservation equations, the final stress, , and strain, , can be found as

= O C L u , = u C L .

{ .A.3} { .A.4}

At these points, the Lagrangian wave speed, CL, changes with the changing particle velocity, u [Reinhart and Chhabildas, 2003]. The density is, O. Before the second shock, the Lagrangian wave speed is found by
C L (u fs ) =

T + t (u fs )

{ .A.5}

In Equation { .A.5}, T + t, is the time from the velocity profile, corresponding to each, ufs, which are the free surface velocities. The target thickness is, T. After the second shock, the Lagrangian wave speed is determined by
C L (u fs ) =

T + t (u fs ) 2 F /U eff

{ .A.6}

In Equation { .A.6}, F, is the flyer thickness and, Ueff, is the effective velocity of the second shock wave. The shear stresses generated in the material due to the shocks can be calculated from the longitudinal stresses in the material based on elementary theory [Asay and Chhabildas, 1981]. The planar impact experiments are in a state of uniaxial strain, and the stress can be written as

X Y = 2 .

{ .A.7}

211

Here, X, represents the longitudinal stress, Y, the lateral stress, and, , the resolved shear stress. Also, the mean stress, , for the shocked state can be written as

1 ( X + 2 Y ) . 3

{ .A.8}

From Equations { .A.7} and { .A.8}, the longitudinal stress, X , can be written as a function of the mean and shear stresses

X = + .

4 3

{ .A.9}

Differentiating Equation { .A.9} with respect to, , the engineering strain d X d 4 d = + . d d 3 d

{ .A.10}

The engineering strain is also related to the ratio of the initial density, O, and the shocked density, [Asay and Chhabildas, 1981]

= 1 O .

{ .A.11}

At the yield point during the reshock or unload, the shear stress, , is maximized and is equal to, C, the critical shear strength. After the initial stress pulse in the experiment, the material has a stress level, O, which is beyond the HEL. This stress has a factor of, (4/3) O that separates it from the mean stress,

O = + O .

4 3

{ .A.12}

In a material that follows an ideal elastic-plastic model, the shear stress, O, at the end of the initial stress pulse, which is past the yielding point is equal to, C,

212

but in actuality these two shear stresses may be different [Reinhart and Chhabildas, 2003]. During the second shock, the material unloads elastically at first and yields again at Point 1 or reloads in the same process to Point 2 (Figure 6.1), where inelastic behavior begins. In either case, once the stresses and strains from the initial impact up to Points 1 and 2 have been determined numerically from the velocity plot, Equation { .A.9} can be integrated also numerically over the rising stress pulse from the initial Hugoniot state to the final state where yielding occurs. The resulting equations relating the maximum shear strength, C, to the shear stress, O, are as follows [Asay and Chhabildas, 1981]. For the unloading experiment
1 3 2 C + O = O ( CL 2 CB )d . 4 O

{ .A.13}

For the re-loading experiment


2 3 2 C O = O ( CL 2 CB )d . 4 O

{ .A.14}

In Equations { .A.13} and { .A.14}, the Lagrangian wave velocity, CL, varies instantaneously and is related to the stress and the strain as:

CL 2 =

1 d X . O d

{ .A.15}

This wave velocity is measured experimentally. The bulk wave velocity, CB, is assumed to be constant and represents the wave velocity after the transition to plastic behavior [Asay and Chhabildas, 1981]. It can be written as:

2 CB =

1 d . O d

{ .A.16}

213

The factor, C + O, is detrmined numerically using shock velocity information from the release experiment. From the reloading experiment, C O, is determined. The points can also be determined graphically on the stress versus strain plot (Figure 6.1) [Reinhart and Chhabildas, 2003]. These two quantities are then related to the maximum and minimum longitudinal stress, max , and,

min , as follows

min = O
and,

4 ( C + O ) , 3

{ .A.17}

max = O +

4 ( C O ) . 3

{ .A.18}

These values can be related to, YC, the critical dynamic yield strength of the material:
YC = 2 C = ( max min ) .

{ .A.19}

The series of experiments were all performed above the HEL of the alumina, which was found to be between 6.7 and 7.9 GPa. Through multiple experiments, the dynamic yield strength was determined to increase with increasing pressure [Reinhart and Chhabildas, 2003]. A.2. Reverberation Experiments The second method for determining the state of the shocked material employs the use of successive unloading. This method uses a thick plate of test material as the flyer and a very thin plate of a material with known properties including higher impedance as the target. The reverberations of the stress in the target causes a series of unloading waves to propagate in the specimen. The calculation method by Hall et al., [Preliminary] examines the successive unloading states.

214

Dandekar examined the shock and release of glass-fiber-reinforced polyester (GRP) using this plate reverberation method [Dandekar et. al., 2003]. The material response of GRP was characterized by means of VISAR free surface velocity measurements and by x-cut quartz strain gages. These measured the free surface particle velocity and the internal flyer stress respectively. In this experiment series, copper and tantalum, which have higher impedances than the GRP, were employed as the target material. The experiments covered the range of Hugoniot state stresses between 1.3 and 20.3 GPa. The initial compressive shock wave causes the flyer material and the target to deform, reaching the Hugoniot stress state. However, because the target is much thinner than the flyer, by the time the wave has propagated to the rear surface of the specimen, the waves have reflected off both surfaces of the target several times (Figure 6.2). Each reverberation from the impacted surface results in partial wave propagation causing a partial unloading in the specimen. The weaker shock in the target continues to reflect. Each reflection from the free surface causes a drop in the stress in the specimen and the target and a corresponding rise in the free surface particle velocity. This velocity rise is measured by the VISAR. As a result of using the shock relations, the stress states in the specimen can be determined from the free surface particle velocity [Dandekar et al., 2003]. The Hugoniot state for the specimen can be found from the Rankine-Hugoniot jump conditions [Asay and Chhabildas, 1993]. The stress, H, and strain, H, at this state are given by

H = OUu , H = u U .

{ .A.20} { .A.21}

215

In Equations { .A.20} and { .A.21} the initial density is, O, the shock velocity is, U, and the particle velocity is, u. Often, a linear relationship between the shock velocity and the particle velocity (Equation of State) is used, i.e.

U = CO + Su .

{ .A.22}

In Equation { .A.22}, CO, and, S, are parameters. By combining equations { .A.20} and { .A.21} for stress and strain, with the linear relationship between, U, and, u, Equation { .A.22}: the stress as a function of the strain can be written as follows [Dandekar et al., 2003]
2 O CO H H = . 2 (1 S H )

{ .A.23}

This equation was employed to plot versus the stress versus strain data for Dandekar et al.s [2003] experiments. The following analysis is applicable if there is equivalent loading and release behavior in the flyer and target [Hall et al., Preliminary]. In determining the stress and strain in the flyer at the Hugoniot state in the case where the impedances of the flyer and target differ, the stress in the target, T, is found from Equation { .A.20},

T = OTU T ( u2 / 2 ) .

{ .A.24}

The unshocked density of the target is, OT. The particle velocity of the target, uT, is one half of the recorded velocity on the free surface, u2, after the first stress wave has passed,
uT = u2 2 .

{ .A.25}

216

This stress state on the free surface is State 2. The shock velocity in the target is, UT. The shock velocity of the target plate is determined from Equation { .A.22}. The stress in the flyer and the target are identical, because the stress is continuous over the plane of impact. Thus, the stress in the flyer, S, is also given by

S = T = OTU T ( u2 / 2 ) .

{ .A.26}

The subscript, S, refers to the specimen being the flyer, not the target, which is indicated with the subscript, T. The particle velocity of the flyer is the difference between the flyer impact velocity, VI, and the targets velocity at the Hugoniot state, uT, which was found in Equation { .A.25}:

uS = VI uT .

{ .A.27}

The relation in Equation { .A.27} enables the determination of the shock velocity of the flyer, US, from Equation { .A.20}:
US = S

( OS uS ) .

{ .A.28}

In Equation { .A.28}, the initial density of the flyer is, OS, and the particle velocity is, uS. The strain in the specimen, S, can be found by using the shock velocity from Equation { .A.28} and Equation { .A.21}:

S = uS U S .

{ .A.29}

Equations { .A.24} through { .A.29} determine the Hugoniot state stress strain, shock velocity, and particle velocity in the flyer and the target [Hall et al., Preliminary]. For each successive release state, incremental forms of these equations are used to determine the stress, shock velocity, and strain in the flyer. The change in the particle velocity on the free surface of the target, u X + 2 u X , is employed to determine the change in the particle velocity in the specimen, uS .

217

Here, X, represents the states on the free surface. For example, State 2 is, X=2, and State 4 is, X=4. First, from the stress velocity diagram in Halls paper [Preliminary], the velocity difference between States 1 and 3 is half the velocity difference of States 2 and 4, assuming equivalent shock and release behavior as above. following expression: This leads to the

uT (31) = u3 u1 =

1 ( u4 u 2 ) . 2

{ .A.30}

Generalizing this expression for any two points:

uT =

1 ( uX +2 uX ) . 2

{ .A.31}

The particle velocity jump in the flyer is the same as that for the target [Hall et. al., Preliminary], so the relation can be rewritten:

uS = uT =

1 (uX +2 uX ) . 2

{ .A.32}

The stress increment is found by using the incremental form of Equation { .A.26} :

S = T = OT CT uT .

{ .A.33}

This can be rewritten in terms of the free surface particle velocity jump from Equation { .A.31}:

S =

1 OT CT ( u X + 2 u X ) . 2

{ .A.34}

218

This is the equation used in [Hall et. al., Preliminary]. However, during the course of the experiments in the current study, a different relationship between the stress increment in the flyer and the free surface velocities was derived:
S = OS CT ( X 2) ( uT ( X 2) uT ( X 3) ) OS CT ( X ) ( uT ( X ) uT ( X 1) ) .

{ .A.35}

This modification is computed directly from the stress-velocity diagram for one of the reverberation experiments on silicon nitride (Figure 6.3). The quantity, CT, represents the average Lagrangian wave speed in the target during unloading,

CT = 2h t .

{ .A.36}

The target thickness is, h, and the time it takes the wave to transit this distance is, t . The thickness is measured prior to the experiment, and the time is determined from the distance between the two velocity jump midpoints on the free surface time history [Hall et. al., Preliminary]. With the changes in stress and particle velocity, the longitudinal unloading average wave velocity, CS, in the specimen can be determined by employing the incremental form of Equation { .A.28} [Hall et. al., Preliminary], even though the release velocity is not technically a shock [Dandekar et al., 2003]:
C S = S

( OS uS ) .

{ .A.37}

Then, the strain increment in the flyer, S , can be calculated from Equation { .A.37} and the incremental form of equation { .A.29} [Hall et. al., Preliminary]:

S = uS CS .

{ .A.38}

Finally, the increments in stress and strain are combined with the stress and strain levels at the previous state ( X , X ) to determine the absolute stress and strain for each new state ( X + 2 , X + 2 ):

219

X + 2 = X , X + 2 = X .

{ .A.39} { .A.40}

The stress versus strain and the stress versus particle velocity relationships for these experiments based on the method just described were presented in Hall et al.s [Preliminary]. Also in this paper, a plane acceleration method was employed to determine the same relations. That method used Isaac Newtons 2nd Law of Motion. Both methods agreed well in experiments with Hugoniot pressures up to 10 kbars prompting the suggestion that permanent compacting of the flyer was occurring in those experiments [Hall et al., Preliminary]. The dynamic yield strength in the material can also be estimated from Hall et al.s method when the HEL of the specimen material is exceeded. If the material is considered elastic-perfectly plastic, then the shear stress at the Hugoniot state,

H, is equivalent to the critical shear stress, C. This results in the dynamic yield
stress upon unloading being twice,C, at the point where the deformation along the unloading slope changes from elastic to elastic-plastic [Reinhart and Chhabildas, 2003]. The graphical method described in this paper can be applied using the curve connecting the unloading states on the stress-strain plot. Or, numerically, the method from Asay and Chhabildas [1981] can be used. At the point where the deformation changes from elastic to elastic-plastic, the shear stress in the specimen, CS, is measured by modifying Equation { .A.13}. Since, the dynamic yield stress, YCS, is twice the critical shear stress:
1S 3 2 2 = OS ( CS CBS )d S . 4 OS

YCS = 2 CS

{ .A.41}

In Equation { .A.41} the elastic-plastic (bulk) wave speed in the flyer is given as, CBS. Integrating Equation { .A.41} numerically results in an expression for the

220

dynamic yield stress of the material. This can be compared with the other method of calculating dynamic yield stress outlined in Section A.1.

221

B. Design of Shock Re-shock, Shock Unload, and Reverberation Experiments The experiments in this study were performed in the highest impact velocity range, i.e. 500 m/s, of the 3.25 inch diameter single stage gas gun at CWRU. The specifications of this gas gun and of the experimental procedures were discussed in Chapter IV. The high HEL of silicon nitride, determined in Chapter V to be 12 GPa, requires a relatively high impact velocity in order to study inelasticity in Si3N4. The specific design of each of these experiments are discussed below. As mentioned in Section A.1 dual shocks are created by using a dual component flyer. For the shock-reshock experiment, SR-1, the flyer was composed of silicon nitride and backed by a higher shock impedance tungsten carbide plate (Figure 6.4). The HEL of the silicon nitride was exceeded in this experiment. This is due to the higher impedance of the tungsten carbide. Silica glass, which has an impedance of 12.8 MPa/(m/s) is used as a window behind the Si3N4 specimen. in order to prevent spallation of Si3N4 during the experiment. The unloading experiment, SU-1 was backed with a Al 6061 plate, which has a relatively lower shock impedance of 16.6 MPa/(m/s) when compared to Si3N4, to create unloading. Again, a silica glass plate was used as a window material behind the Si3N4 target. In this way, the experimental configuration requires two plates for the flyer and two for the target (Figure 6.4). The reverberation experiments, patterned after those discussed in Section A.2, involve a thick flyer of silicon nitride impacting a thin plate of tungsten carbide. In experiment RB-1 the flyer was 4.5 mm thick while the target is only 1 mm thick (Figure 6.5). This configuration allows two reverberations in the WC target before the free surface boundary wave from the flyer arrives at the targets free surface. The second experiment, RB-2, (Figure 6.2) involved a 5.5 mm thick flyer, which allowed for three reverberations to occur before the arrival of the boundary wave.

222

For these four experiments, predictions were determined by using elastic and elastic-plastic line equations, such as those used in Chapter V to approximate one-dimensional hyperbolic wave theory. These equations and their results are discussed in Section C. As in Chapter V, the VISAR interferometer system was employed to measure the in-material free surface particle velocity histories at the interface of the target and the window to provide input for the equations. The results are compared to the elastic and elastic-plastic predictions from the equations in Section A.

223

C. Results and Analyses of Multiple Shock Experiments The four experiments were carried out according to the design in Section B. From the VISAR measurements of particle velocities, the internal stress behaviors are determined and related to the elastic-plastic predictions. The particle velocities indicate behavior which is close to that predicted by the elastic-plastic estimates in Section A and by the line equations derived in this section. The VISAR measurements and the internal stress levels are shown in Tables 6.1 and 6.2. Any changes from the behavior of unshocked material are documented in Sections C.1 and C.2. C.1. Shock Re-shock and Shock Unload Experiments In the dual shock experiments, the materials were subjected to either a shock followed by a re-shock from a higher impedance material or an unload from a lower impedance material. This resulted in an initial Hugoniot state followed by a velocity rise or drop to a second level. Two experiments were conducted with the impact velocity of 479 4 m/s, one shock re-shock and one shock unload. At this velocity, the HEL of silicon nitride is not exceeded in the first shock. In the experiment with the re-shock, however the HEL is exceeded during the second shock. In the unload experiment, the HEL is not exceeded. In Section C.1.1, the elastic and elastic-plastic computations for these experiments are given, and then in Section C.1.2, the free surface velocity profiles for the re-shock and unload experiments are discussed. C.1.1. Computations for Shock Re-Shock and Shock Unload Experiments The flyer is accelerated to an initial velocity and has zero stress, State 1 (Figure 6.6). The initial shock comes from the impact of the two silicon nitride plates, which in this study are 4.5 mm thick for the target and 3 mm thick for the flyer. Before this, the target is in a state of zero stress and zero particle velocity, known as State 2. Because both the flyer and the specimen are silicon nitride, the impact is initially symmetric and produces a particle velocity, V3, in State 3 that is one-half the free surface particle velocity, VFS.

224

V3 = 1 VFS . 2

{ .C.1}

The stress for this state is calculated by characteristic line equations that approximate 1-D hyperbolic wave theory, the use of which is described in Chapter V. Because the HEL is not exceeded, the stress in State 3 is found,

3 = ZTV3 .

{ .C.2}

Equation { .C.2} contains the shock impedance of the silicon nitride, ZT. Because a second flyer material is employed with a different impedance, a second stress pulse is generated from the interaction of the wave with the boundary between the two pieces. This stress lasts until the reflection of the pulse from the free surface of the backing plate. The stress state caused by this second shock is represented by State 4. In the loading experiment, SR-1, State 4 has greater stress and particle velocity than State 3 (Figure 6.6). Since the HEL is exceeded by this stress level, elasticplastic waves are generated and the stresses are obtained by employing the equations of state of the tungsten carbide and the silicon nitride. The tungsten carbide HEL, HEL WC , is only 7.2 GPa, whereas the HEL of silicon nitride is 12 GPa, HEL SiN . The velocity at the HEL, VHEL WC , of the tungsten carbide backing plate is found:

VHEL WC = V1 HEL WC / Z B .

{ .C.3}

The shock impedance here, ZB, is that of the backing plate, which for tungsten carbide is 106 MPa/(m/s). Using the bulk modulus of tungsten carbide, the elastic-plastic shock impedance, ZBP, can be approximated as 76.6 MPa/(m/s). The solution to Equation { .C.3} is 407 m/s. From this, a line can be drawn from the HEL stress and velocity of the backing plate to State 4 (Figure 6.6). This line has equation:

225

4 = Z BP (VHEL WC V4 ) + HEL WC .

{ .C.4}

Note that, the HEL stress of silicon nitride is 12 GPa and the HEL velocity is the HEL stress divided by the elastic impedance, ZT, which is 35.5 MPa/(m/s).
VHEL SiN = HEL SiN ZT .

{ .C.5}

The HEL velocity is thus 338 m/s. Using the elastic-plastic wave speed of silicon nitride, 8.27 km/s, the plastic impedance, ZTP, can be approximated as 27.1 MPa/(m/s). Because the equation of state is unknown, this estimate is used as the slope:

4 = ZTP (V4 VHEL SiN ) + HEL SiN .

{ .C.6}

Knowing this, Equations { .C.4} and { .C.6} can be rearranged to provide estimates of the particle velocity of State 4,

V4 =

Z BPVHEL WC + HEL WC + ZTPVHEL SiN HEL SiN . ZTP + Z BP

{ .C.7}

The stress at State 4 is then found using either Equation { .C.4} or Equation { .C.6}. The stress and particle velocity equations for State 4 are used to determine further relationships between the measured velocity and the internal stress behavior.

For the shock-unloading experiment, SU-1, the calculation of State 4 is simpler (Figure 6.7). The material remains elastic, so the elastic wave speeds and impedances suffice for the calculations. Thus, State 4 can be found as the intersection of the lines on the stress-velocity diagram connecting States 2 and 3 the line from State 1 that has the negative of the impedance of the aluminum

226

backing plate, ZB, as its slope. The stress can be found from the following equations:

4 = ZT (V4 V3 ) + 3 , 4 = Z B (V4 V1 ) .

{ .C.8} { .C.9}

Rearranging Equations { .C.8} and { .C.9}, the stress and particle velocity at State 4 are found:
V4 = ( Z BV1 + ZTV3 3 ) ( Z B + ZT ) ,

{ .C.10} { .C.11}

4 = ZT ( Z BV1 + ZT V3 3 ) ( Z B + ZT ) ZT V3 + 3 .

The silica glass window material causes a state of compressive stress at the back of the specimen following the passage of the initial wave (State 5). This is similar to the spall pressure shear experiments, where the window was used to prevent failure due to shear stress. In this case, the addition of compressive stress prevents spallation in State 6. To find the stress at State 6, first the stress and particle velocity for State 5 need to be found. Using the lines in Figure 6.7, from States 2 and 3, State 5 is generated:

5 = ZW V5 ,
5 = ZT (V5 V3 ) + 3 .

{ .C.12} { .C.13}

In Equation { .C.12}, ZW, represents the window impedance. Rearranging Equations { .C.12} and { .C.13} leads to:
V5 = ( 2 3 ) ( ZW + ZT ) .

{ .C.14} { .C.15}

5 = ( 2ZW 3 ) ( ZW + ZT ) .

227

The in-material particle velocity at the interface between the window and the specimen in State 5 is the first measured state. Using this velocity enables the determination of the state of stress in State 3 by reconfiguring Equation { .C.14}:

3 = V5 ( ZW + ZT ) 2 .

{ .C.16}

The state of stress when the two rarefaction waves meet is given by State 6 in both experiments (Figures 6.6, 6.7). The equations for the determination of the stress and particle velocity of this state are the same for both cases. Obviously, in the shock-unload experiment, the HEL is not exceeded, so the unloading is elastic. In the shock-re-shock experiment, the unloading is also elastic. The equations to proceed from State 4 to State 6 and State 5 to State 6 are as follows:

6 = ZT (V6 V4 ) + 4 . 6 = ZT (V6 V5 ) + 5 .

{ .C.17} { .C.18}

By rearranging Equations { .C.17} and { .C.18}, the stress and particle velocity can be found for State 6:
V6 = ( ZT V5 5 + ZT V4 + 4 ) ( 2 ZT ) .

{ .C.19} { .C.20}

6 = ( ZT V5 + 3 5 + ZT V4 + 4 ) ( 2 ) .

State 7 is a measured state on the interface between the specimen and the window. Its stress and particle velocity can be determined from States 5 and 6:

7 = ZW V7 .
7 = ZT (V7 V6 ) + 6 .

{ .C.21} { .C.22}

228

Because the particle velocity in State 7 is directly measured, the stress of State 6 can be found by combining equations { .C.18}, { .C.21}, and { .C.22}:

6 = ( ZW + ZT ) V7 + ( ZT ZW ) V5 .

{ .C.23}

Thus, the elastic and elastic-plastic predictions for dual shock experiments are established. The results of these experiments are detailed in Section C.1.2. C.1.2. Observations and Analysis of Shock Re-shock and Shock Unload Experiments The shock re-shock and shock unload experiments conformed well to the elastic and elastic-plastic predictions. This is expected because the HEL of the silicon nitride is barely exceeded during the re-shock. The measured velocity profiles show this agreement. Both Experiments SR-1 and SU-1 are discussed in this section. Some inelasticity was observed in Experiment SR-1, while Experiment SU-1 conformed entirely to the elastic prediction. The shock re-shock experiment, SR-1, shows good agreement with the elastic prediction in the Hugoniot state (Figure 6.8). The impact speed is 475 m/s. The Hugoniot state velocity is around 358 m/s, while the elastic value is 348 m/s, which is only 3% away. The material begins its increase to the off-Hugoniot state at 0.55 microseconds. Once the off-Hugoniot state is reached, it remains at about 510 m/s, which is within 1% of the elastic-plastic prediction of 506 m/s until boundary waves arrive at about 1 microsecond. Thus, the second shocked state also conforms to the predictions. However, during the re-loading to the second shocked state, a slope decrease is seen at 467 m/s. This is indicative of the HEL limit being exceeded. An estimate for the HEL velocity as measured on the surface between the flyer and the window can be obtained from Figure 6.8. This is found by the following equation:

229

VHEL measured =

2 HEl SiN . ZW + Z SiN

{ .C.24}

This value is 501 m/s. Thus the initial shock, although elastic and below the HEL, has reduced the velocity where the onset of inelastic behavior occurs. The corresponding HEL stress determined from the second shock is found by reversing Formula { .C.24}:

HEL SiN =

VHEL measured (ZW + Z SiN ) . 2

{ .C.25}

The resulting HEL stress is 11.3 GPa, which is reduced by 7% from the unshocked value of 12 GPa. This suggests that the material is permanently deformed and weakened by the initial compressive pulse, but only slightly. The shock unload experiment, SU-1, has an impact velocity of 483 m/s. The free surface particle velocity ascends initially to the Hugoniot state sharply to 372 m/s, which is within 5% of the elastic-plastic prediction of 355 m/s (Figure 6.8). The velocity remains constant at this level until the unloading wave arrives at 0.53 microseconds, at which time it drops to 230 m/s, which is about 1% away from 233 m/s, which is the predicted value. This time, the velocity drop is quite sharp, indicating elastic behavior. Again, the velocity remains constant until the arrival of the boundary wave at 1 microsecond. Thus, the times and magnitudes of the rises and drops in Experiment SU-1 correspond well to the predictions. Using the methods of Asay, Chhabildas, and Reinhart [Asay and Chhabildas, 1981; Reinhart and Chhabildas, 2003], the elastic-plastic deformation of the material can be examined. Equations { .A.3} through { .A.6} were used to numerically evaluate the stress and strain from the time and free surface particle velocity history. The resulting stress vs. strain plot shows that the material behavior is elastic under unloading (Figure 6.9). As can be seen, the off-Hugoniot stress and strain caused by the reloading in Experiment SR-1 follows the

230

hydrodynamic Hugoniot curve. This confirms that little permanent deformation has occurred. Also, the residual shear stress is computed using Equations { .A.13} through { .A.19} and numerically integrating. The resulting residual shear stress is 1.21 GPa. This confirms that little permanent deformation has occurred. The unloading in Experiment SU-1 proceeds in a linear fashion, which indicates elastic unloading. In summary, these impacts do not see large changes in the material behavior or the HEL due to in a material deformed by dual shocks. No visible behavior changes are noted in Figure 6.9. However, because the HEL is barely exceeded, no large variation is expected from the elastic behavior. The only indication of damage is that the HEL as calculated using the second shock wave in Experiment SR-1 is 6% lower than that calculated in Chapter V.

C.2. Shock Reverberation Experiments The reverberation experiments have multiple shocks that unload the material in increments. The impact velocities of experiments RB-1 and RB-2 were also around 500 m/s. In this case, because of the tungsten carbide target plates the HEL of the silicon nitride is exceeded at this impact velocity. Elastic and linearly elastic-plastic predictions are used to estimate the behavior of the material. These are discussed in the Section C.2.1 below. Then, these values are compared with the experimental VISAR free surface velocity data in Section C.2.2. The close correlation between the estimates and the actual velocities indicate that the material behavior is well predicted by the equations in Sections A.2 and C.2.1. C.2.1. Computations for Shock Reverberation Experiments In the reverberation experiments, the initial loading is expected to exceed the HEL of both the silicon nitride and the tungsten carbide. The unloading from this state is also elastic-plastic. However, the reverberations are all elastic. The

231

stress velocity and time distance diagrams for the target are used to determine the internal stress states from the free surface velocities (Figure 6.3). First, the initial states of the flyer and target are specified. For the flyer, State A has velocity, VI, which is the initial impact velocity, and zero stress. The target has State B which is a state of zero velocity and stress (Figure 6.3). The impact of the flyer and target produces State 1, which occurs in the plastic regimes of both materials. As above, the tungsten carbide HEL, HEL WC , is 7.2 GPa and the HEL of silicon nitride, HEL SiN , is 12 GPa. The velocity of the flyer at the HEL of silicon nitride, VHEL SiN , is found:

VHEL SiN = VI HEL SiN / Z SiN .

{ .C.26}

In Equation { .C.26}, ZSiN, represents the elastic impedance of silicon nitride. For the tungsten carbide target, the velocity is found,

VHEL WC = HEL WC / ZWC .

{ .C.27}

In Equation { .C.27}, ZWC, is the elastic impedance of tungsten carbide. From the HEL points of both materials, the elastic-plastic assumption of linear behavior is used. Thus, the two equations that determine State 1 are:

1 = ZWC P (V1 VHEL WC ) + HEL WC ,


and,

{ .C.28}

1 = Z SiN P (VHEL SiN V1 ) + HEL SiN .

{ .C.29}

In Equations { .C.28} and { .C.29} the additional subscript, P, represents estimates for the plastic impedance from the elastic-plastic wave velocities of the silicon nitride and tungsten carbide. Combining these equations to get the velocity in State 1:

232

V1 =

Z SiN PVHEL SiN + ZWC PVHEL WC + HEL SiN HEL WC . ZWC P + Z SiN P

{ .C.30}

The stress at State 1 in the target is then found by substituting equation { .C.30} into either equation { .C.28} or { .C.29}. From this point unloading in the target proceeds in an elastic-plastic fashion. Thus, State 2 has zero stress and exactly double the velocity of State 1 (Figure 6.3).

V2 = 2V1 .

{ .C.31}

State 2 is measured by the VISAR as the first velocity rise on the free surface. All of the reverberations are assumed to occur elastically, since they are below the HEL stress of either material. State 3 is the intersection of the line defined by the elastic equation from State A to the HEL of silicon nitride (Figure 6.3):

3 = Z SiN (VI V3 ) ,

{ .C.32}

and a line rising from State 2 in the tungsten carbide, with slope, ZWC:

3 = ZWC (V3 V2 ) .

{ .C.33}

Equations { .C.32} and { .C.33} are then combined to determine the particle velocity of State 3: V3 = Z SiNVI + ZWCV2 . ZWC + Z SiN

{ .C.34}

233

State 4, which corresponds to the first reverberation on the free surface of the target, is another state of zero stress (Figure 6.3). Its velocity can be found simply by connecting a line with slope, -ZWC, from State 3 to the zero stress axis: 0 = ZWC (V4 V3 ) + 3 . Rearranging Equation { .C.35}: V4 = V3 + 3 / ZWC . { .C.36} { .C.35}

The second and third reverberation free surface velocities and internal stresses in the target can be determined by successively applying Equations { .C.32} through { .C.36} to find States 5, 6, 7 and 8 (Figure 6.3). These formulae enable the determination of the elastic and elastic-plastic prediction of the reverberation experiments.

C.2.2. Observations and Analysis of the Shock Reverberation Experiments

The impact velocities were 471 m/s and 511 m/s respectively for experiments RB-1 and RB-2. Other experimental values are located in Table 6.2, including the free surface particle velocities and internal stresses. First, the free surface velocity results are examined and the relationship between them and the elasticplastic predictions as detailed in Section C.2.1 are examined. The analysis discussed in Section A.2 is then used to determine the material stress vs. strain behavior. Experiment RB-1 has an impact velocity of 471 m/s. The free surface particle velocity shows an initial jump to 240 m/s, which is the first shocked state. After 0.21 microseconds, the velocity rises to 247 m/s (Figure 6.10). This is off of the elastic-plastic prediction of 257 m/s for the Hugoniot by 4%. Then, the first reverberation causes the velocity to jump to 356 m/s. Another slow rise occurs for 0.12 microseconds to 363 m/s. This is the same as the prediction of 364 m/s

234

for the first reverberation to within experimental error of 2%. Following this, the second reverberation occurs, which brings the velocity to 417 m/s. The elastic prediction for this velocity is 417 m/s. The velocity remains at about this level, rising only to 420 m/s until the arrival of the boundary wave at 0.975 microseconds. In experiment RB-2, the impact velocity is 512 m/s. The initial free surface particle velocity rise is to 263 m/s due to the higher impact velocity (Figure 6.10). The elastic-plastic prediction is 277 m/s, from which the actual velocity is off by 5%. Then the velocity slowly rises for 0.17 microseconds to 267 m/s before jumping to 384 m/s. The predicted velocity level for the first reverberation is 395 m/s. This first reverberation rises to 390 m/s over 0.16 microseconds before jumping to the second reverberation level of 448 m/s. The velocity level rises to 455 m/s, 0.23 microseconds after the beginning of the reverberation. This compares well to the elastic-plastic prediction of the velocity level: 453 m/s. The final reverberation begins at 478 m/s and ends 0.1 microseconds later at 483 m/s. Its predicted level is 483 m/s. From these experimental plots (Figure 6.10) the materials can be seen to follow the elastic-plastic predictions quite well. The level of error of the VISAR is 2%. The Hugoniot and the reverberation free surface particle velocities are predicted to within 5% by the elastic-plastic theory derived in Section C.2.1. Additionally, the rise times of the prediction agree closely with the times of the jumps in the particle velocities between the states. The extended nature of these rise times indicates elastic-plastic behavior in the tungsten carbide specimen. By use of the equations provided by Hall [Hall et. al., Preliminary] and Dandekar [Dandekar et al., 2003], the stress and strains of each reverberation state can be observed. The midpoints of the velocities in each state are employed for this purpose. The stress and strains in the silicon nitride are plotted versus the elastic Hugoniot curve established in Chapter V (Figure 6.11). As can be seen from the figure, the stresses and strains cluster along the elastic predicted line,

235

within the experimental error, indicating that the behavior of the silicon nitride is primarily elastic. This, combined with the elastic-plastic behavior of the tungsten carbide, means that the material behavior during the reverberation experiment is well predicted by the current methods.

236

D. Conclusions and Discussion

Two methods for evaluating the elastic and elastic-plastic material behavior and dynamic yield strength subjected to successive shocks from the literature are presented in Section A. Using the single stage gas gun and appropriate choices of flyer, target, backing, and window materials, experiments using these methods were conducted at impact velocities around 500 m/s. However, the high HEL of silicon nitride prevented the experimental Hugoniot stress from greatly exceeding the HEL in the reverberation experiments. In the shock re-shock and shock unload pair of experiments, the HEL was only exceeded in the second shock of Experiment SR-1. An estimate of the residual shear stress is 1.21 GPa. Within the obtained range of stresses, however the material behavior in the four experiments is well predicted by the equations describing elastic-plastic motion presented in Sections A and C of this chapter. Shock re-shock Experiment SR-1 indicates that the silicon nitride deforms mainly in a hydrodynamic fashion, but reaches the HEL near the end of the re-shock (Figure 6.9). Shock unloading Experiment SU-1 initially follows the hydrodynamic curve as well, but unloads in an elastic manner, (Figure 6.9) as is expected below the HEL. In the two reverberation experiments, the tungsten carbide behaves in a predictable elasticplastic manner, while the silicon nitride exceeds its HEL during the initial shock, but the reverberations remain elastic. Because of this, the stresses and strains of the reverberation states conform to the elastic relationship established in Chapter V (Figure 6.11). As a result of the high HEL of silicon nitride, the experiments were unable to probe deeply into the nature of inelastic multiple shock behavior. However, one indication of a small amount of inelastic deformation from the first shock is that a simple calculation of the HEL in experiment SR-1 shows the HEL to be reduced 6% from 12 GPa to 11.3 GPa. The equations in Section A [Asay and Chhabildas, 1981; Reinhart and Chhabildas, 2003] indicate a residual shear stress of 1.21 GPa. The material behavior does not exhibit a significant change from

237

the hydrodynamic prediction along the lines of those described in the literature. Similarly, the reverberation experiments indicate elastic behavior. Because the initial Hugoniot state in the re-shock/unload experiments is below the HEL, the effect of subsequent shocks on the dynamic strength of the material above the HEL remains largely unexplored. A gas gun with the capability of a greater maximum impact velocity is necessary to further investigate this area.

238

References
Asay, J. R. and L. C. Chhabildas, 1981. Determination of the Shear Strength of Shock Compressed 6061-T6 Aluminum. In: Shock Waves and High-Strain Rate Phenomena in Metals. Plenum Publishing Corporation. New York, New York. 417-431. Dandekar, D. P., C. A. Hall, L. C. Chhabildas, W. D. Reinhart. 2003. Shock Response of a Glass Fiber-Reinforced Polymer Composite. Composite Structures. 61 [1-2] 51-59. Hall, C. A., L. C. Chhabildas, and W. D. Reinhart., Preliminary Report. Shock Hugoniot and Release States in GRP Composite From 3 to 20 GPa. Sandia National Laboratories. Reinhart, W. D.; L. C. Chhabildas. 2002. Strength Properties of Coors AD995 Alumina in the Shocked State. International Journal of Impact Engineering. 29 [1-10] 601-619.

239

Tables
Table 6.1: Dynamic Properties for Experiments SR-1 and SU-1. The abbreviation Hydro. stands for hydrodynamic. Free Surface State 5 Exp # SR-1 SU-1 Impact Velocity 475 m/s 483 m/s Backing Plate WC Al Measured Velocity 358 m/s 372 m/s Elastic Velocity 348 m/s 355 m/s Hugoniot State Hydro. Stress 8.0 GPa 8.6 GPs Hydro. Strain 2.6E-2 2.5E-2 Free Surface State 7 Measured Velocity 510 m/s 230 m/s Elastic Velocity 506 m/s 233 m/s

Table 6.2: Dynamic Properties for Experiments RB-1 and RB-2. Note, stress and strain in States 1, 3, 5, and 7 are computed respectively from measured velocity in States 2, 4, 6, and 8. Exp # Impact Velocity Flyer Thickness State 1: Actual Stress State 1: Actual Stain State 2: Measured Velocity State 2: Predicted Velocity State 3: Actual Stress State 3: Actual Stain State 4: Measured Velocity State 4: Predicted Velocity State 5: Actual Stress State 5: Actual Stain State 6: Measured Velocity State 6: Predicted Velocity State 7: Actual Stress State 7: Actual Stain State 8: Measured Velocity State 8: Predicted Velocity RB-1 471 m/s 4.5 mm 12.4 GPa 3.19E-2 245 m/s 257 m/s 6.2 GPa 1.48E-2 361 m/s 365 m/s 3.16 GPa 6.63E-3 420 m/s 418 m/s n/a n/a n/a n/a RB-2 512 m/s 5.5 mm 13.4 GPa 3.48E-2 265 m/s 277 m/s 7.01 GPa 1.56E-2 389 m/s 395 m/s 3.43 GPa 0.763E-3 452 m/s 454 m/s 1.59 GPa 3.82E-3 482 m/s 483 m/s

240

Figures

Figure 6.1: Theoretical stress vs. strain diagram showing graphical method for computing dynamic yield strength using the technique from Reinhart and Chhabildas [2003]. The initial loading occurs along the hydrodynamic Hugoniot curve from zero to the Hugoniot state. In the re-shock experiment, the wave velocity transitions from elastic to bulk (elastic-plastic) at Point 2. Point 1 is the location where the same occurs in the unloading experiment. By using the stress values at these points, the dynamic yield stress can be found.

241

Figure 6.2: Reverberation experiment design showing silicon nitride flyer and tungsten carbide target. The VISAR probe reflects off of a thin aluminum layer on the surface of the target. The thin target and thicker flyer are visible.

Figure 6.3: (a) Time vs. distance and (b) stress vs. velocity diagrams for reverberation experiment RB-2 showing computation of S . The change in stress between States 3 and 7 can be found using the velocities measured on the free surface, States 2 through 8. Knowing the slope of each purple line, which is the targets impedance, CT, and these velocities, the stress at each state can be found. Then, by subtracting two stresses, Equation { .A.35} is generated.

242

Figure 6.4: Specimen and flyer configuration for shock re-shock and shock release experiments. The backing plate on the flyer is shown, as is the silica glass window.

Figure 6.5: (a) Time vs. distance and (b) stress vs. velocity diagrams for reverberation experiment RB-1. The method of determining the stress jump is the same as in Figure 6.3. One fewer stress reverberation occurs in this experiment than in experiment RB-2, because of the reduced thickness of the flyer.

243

Silica Glass

Silica Glass

Figure 6.6: (a) Time vs. distance diagram, (b) stress vs. velocity diagram for shock re-shock experiment. The HEL of both tungsten carbide and of silicon nitride are exceeded in the re-shock state.

Silica Glass

Silica Glass

Figure 6.7: (a) Time vs. distance diagram, (b) stress vs. velocity diagram for shock unloading experiment. The HEL of silicon nitride is not exceeded in either the initial shock or the unloading states.

244

0.6

0.5

HEL

Re-shock

Velocity (km/s)

0.4

0.3

Hugoniot

0.2

Unload

0.1

0 -5E-07

5E-07

1E-06

1.5E-06

Time (seconds)
Figure 6.8: Free surface particle velocity vs. time profiles for shock re-shock and shock unload experiments SR-1 and SU-1. The initial rise corresponds well with the predictions. The shock reshock experiment shows an HEL at a velocity below that which occurred without a previous shock. The shock unload experiment appears elastic upon unloading.

245

1.4E+10 1.3E+10 1.2E+10 1.1E+10 1E+10


Hugoniot Curve Re-shock Unload
1

Stress (Pa)

9E+09 8E+09 7E+09 6E+09 5E+09 4E+09 3E+09 2E+09 1E+09 0 0 0.01

Elastic Line Elastic Unloading


2

Hydrodynamic Loading

0.02

0.03

0.04

0.05

Strain
Figure 6.9: The stress vs. strain plot for reloading and unloading experiments that shows the material behavior is elastic under unloading and hydrodynamic under re-shock. The initial shock loading is hydrodynamic. As expected from the equations, the loading under shock re-shock also follows the hydrodynamic predictions from Chapter V. The unloading is linear and elastic.

246

0.5 RB-2 RB-1 0.4

Velocity (km/s)

0.3

0.2

0.1

2.5E-07

5E-07

7.5E-07

1E-06

Time (sec)
Figure 6.10: Free surface particle velocity vs. time profiles for reverberation experiments RB-1 and RB-2. The experiential velocities correspond well with the elastic and elastic-plastic predictions.

247

16 15 14 13 12 11
Elastic Hugoniot RB-1 RB-2

Stress (GPa)

10 9 8 7 6 5 4 3 2 1 0 0 0.01 0.02 0.03 0.04 0.05

Strain
Figure 6.11: Stress and strain levels for the two reverberation experiments plotted against the elastic Hugoniot curve. The stresses and strains of the unloading states are elastic to within the experimental error.

248

Chapter VII Investigation of Shock Induced Failure Wave Propagation in Soda Lime Glass
In the present study, in order to examine the shock wave response of soda lime glass, plate impact experiments were performed using a single stage gas gun. In addition to the compression and shear waves that are observed in ceramics and metals, a second type of wave, known as a failure wave is present in glasses. This failure wave, which has been described in several studies, is a result of damage accumulation in glasses during the initial stages of planar shock compression. It is generally agreed that the spall strength as well as the shear impedance behind the failure wave in glasses are negligibly small [Kanel et al., 1992]. In order to examine the suitability of soda lime glass for use in applications where impacts are likely, a series of experiments were performed in the present study to characterize the failure wave and its effects. The goal of the experiments was to measure the effects of the failure wave on spall strength, dynamic shear strength, and the longitudinal and shear impedance under combined compression and shear shock-wave loading. Specifically, in the study one of the experiments was designed to evaluate the spall strength in soda-lime glass under uniaxial shock compression. A companion experiment examined the degradation in the spall strength with the addition of a shear stress due to a skew angle of 18 degrees. The results of these two experiments were used to establish the spall strength of soda lime glass. The remaining six experiments, discussed in this chapter, were designed to study of the failure wave and its effects on the impedance, the spall strength and the shear strength of the soda lime glass at the eighteen degree skew angle. In order to understand the nature of the failure wave, studies by several authors describing previous experimentation on glasses are discussed in Section A. This discussion is followed in Section B by a description of the experiments performed in the current study. The newly developed method for using the VISAR

249

interferometer to measure the shear velocity is also described in Section B. Then, in Section C, the experimental results are analyzed and compared with the previous results from the literature. A. Description of the Nature of the Failure Wave In order to understand the nature of the current experiments, an explanation of the nature of the failure wave and its effects on glasses is necessary. In Section A.1, several previous experimental studies are reviewed. Each of these studies provides an insight into the physical nature of the failure wave, or the change in properties of glass such as impedance, shear strength, or spall strength behind the failure wave. In Section A.2, three models are suggested to explain the failure wave and its observed features. These models are compared with experiments and numerical computations. A.1. Experimental Determination of the Properties and Effects of the Failure Front Several types of dynamic loading experiments have been employed in order to examine the failure wave in soda lime glass over the past fifteen years. These experiments have included gas gun shock compression experiments on plates such as those performed by Kanel et al. [1992]. These plate impact experiments have included sandwiched configurations [Espinosa et al., 1997] as well as flyer and single target experiments [Clifton et al., 1998]. Each of the experiments described in this section describes an aspect of the failure wave. Kanel et al. [1992] conducted an experimental and numerical work on the shock compression of glass and fused quartz. In this work, a shock loading of 4.5 GPa was induced in K19 glass plate specimens by means of high speed planar impacts using copper flyers. Shock loading of fused quartz with aluminum flyers was carried out by similar means. Free surface velocities were measured using the VISAR system. The experiments were conduced close to but not exceeding the HEL for the two materials.

250

A series of normal plate impact experiments was conducted on K19 glass to confirm the existence of failure waves in glass. In these experiments a release wave that originates at the back of the target plate interacts with the advancing failure wave and a small recompression was noted in the longitudinal particle velocity history [Kanel et al., 1992]. This anomalous recompression was explained as a wave reflection from the failure wave front due to a reduction in the mechanical impedance of the material behind the failure wave. Varying the thickness of the specimen showed that the depth of the failed region from the impact face increases with time. The authors interpreted this as a failure wave moving through the K19 glass at a slower speed than the elastic wave velocity. They also noticed that the failure wave front velocity decreases with increasing depth [Kanel et al., 1992]. Simulations of the experiments utilizing a onedimensional elastic-plastic model were able to match the anomalous recompression in the longitudinal stress history by means of a reduction in the shear modulus at the higher stress levels. This model indicated that the shear stress and the spall strength of the shocked glass dropped to zero behind the failure front. The decreasing velocity of the failure front was also accounted for in the calculations. In this way, the model predicted the form of the stress and free surface velocities for the propagation of failure waves in the K19 glass [Kanel et al., 1992]. Brar et al. [1991 A], were also among the first to publish on the topic of failure waves. They defined the failure wave front as the boundary between the intact and the comminuted material. The definition of a comminuted material is one that is broken into pieces that are fully compacted although heterogeneously deformed [Feng, 2000]. In their study, Brar et al. [1991 A] conducted impact experiments on both glass bars and plates. They employed 12.7 mm by 150 mm long diameter pyrex glass bars and impacted them with either steel or pyrex striker bars accelerated using a gas gun. Observations of the failure wave were conducted using high speed cameras. Soda lime glass plates were also impacted with flyers of copper and aluminum by means of a gas gun in a plate impact

251

configuration. Transverse strain gages embedded within the soda lime glass specimens were used to measure the strain. High speed camera images of the bar experiments in Brar et al.s [1991 A] study indicated that the failure wave propagated at differing velocities depending on the velocity of impact. The comminution of the material was observed directly. At an impact speed of 125 m/s, the failure wave speed was 2.3 km/s and at an impact speed of 330 m/s the failure wave front propagated at 5.2 km/s. The velocity of the failure wave was also greater when a steel striker bar rather than the pyrex glass was used. This suggests that the velocity of the failure wave is also dependent on the impact stress. Additionally, behind the failure wave, the bar was observed to expand in the radial direction in an explosive fashion [Brar et al., 1991 A]. On the other hand, the plate impact experiments on glass were designed to observe any changes in the shear strength of the shocked glass behind the failure wave front. A two wave structure was noticed in the gage record indicating the propagation of both the longitudinal and failure waves. The second recorded wave was not the failure wave itself, but rather the longitudinal release wave that is reflected first off the free surface of the flyer and then off the failed material boundary. The existence of this second wave indicated a change in material property in the shocked region behind the failure wave. From these measurements, the failure wave speed was estimated to be 2.2 0.2 km/s. Using the impedance matching method, the critical longitudinal stress to initiate the failure wave was found to be 3.8 GPa (38 kbar). Moreover, below a critical longitudinal stress, before the arrival of the failure wave, the measured shear strength of the shocked glass was observed to increase linearly with longitudinal stress. However, above this critical stress, the shear strength is observed to level out at about 1.1 GPa. The comminuting of the glass is assumed to be responsible for this loss of shear strength and a corresponding increase in the mean stress [Brar et al., 1991 A].

252

Kanel et al. [2002] in a more recent paper studied the failure wave more closely. In these experiments soda-lime glass were utilized instead of the K19 glass utilized used in their previous study [Kanel et al., 1992]. Two types of experiments were performed. In the first type, plates of glass sandwiched between copper were impacted by aluminum flyer plates to determine the HEL of soda lime glass. The HEL of the soda lime glass was measured to be 8 GPa. In the second set of experiments, aluminum flyer plates were used to impact single soda-lime glass plates. This configuration was used to study the effect of the failure wave on the measured free surface velocity histories and to examine the spall strength of the material ahead of the failure wave front. The spall strength of the intact glass ahead of the failure wave was determined to be greater than 3 GPa [Kanel et al., 2002]. Moreover, like in their earlier study, a recompression wave was observed in the measured particle velocity history at the rear surface of the glass target plate. From these measurements Kanel determined the propagation speed of the failure wave front to be 1.55 0.06 km/s [Kanel et al., 2002]. The Hugoniot Elastic Limit and the spall strength of glass were examined in a study performed by Rosenberg et al. [1985]. In Rosenberg et al.s [1985] release wave experiments, an increase in stress was recorded upon the arrival of the recompression wave. This indicated the presence of a failure wave. The HEL of soda lime glass was found to be 6.4 GPa. In the spall strength experiments above the HEL no spall strength was recorded [Rosenberg, et al., 1985]. Below the HEL, Brar et al. [1991 B] found that at shock stresses of 4.9 GPa through 5.7 GPa behind the failure wave there is no spall strength. In front of the failure wave, there is no spall recorded, and thus the spall strength is higher than 3 GPa [Brar et al., 1991 B].
[VP2]

In a paper by Ginzburg and Rosenberg [1998] three plate impact experiments were described on soda lime glass using brass flyers. These experiments used a PMMA backing plate and an internal strain gage. The impact velocities produced compressive stresses of about 2 GPa, 4.5 GPa and 6 GPa [Ginzburg

253

and Rosenberg, 1998]. A summary of the three commonly accepted conditions of the failure waves are given in this paper [Ginzburg and Rosenberg, 1998]. First, the failure waves are present at impact stresses that are greater than half of the HEL of the glass, but lower than the HEL. Second, the failure front propagates at a speed between 1.5 km/s and 2.5 km/s. Third, the shear strength decreases and the spall strength disappears entirely behind the failure front. In Ginzburg and Rosenbergs [1998] 6 GPa experiment, the failure wave was detected by a recompression in the velocity at the time of reflection from the rarefaction wave/failure front interface. This anomalous recompression is not seen in the lower stress experiments (4.5 GPa), suggesting that the threshold stress is between these two levels [Ginzburg and Rosenberg, 1998]. This is consistent with the previous work in the literature. Moreover, in these experiments a second recompression wave was not detected, which according to Ginzburg and Rosenberg [1998] implies no change in the impedance of the material. The time for this second wave to arrive was computed by assuming the failure front stops moving after the intersection of the unloading wave with the propagating failure front. Also, there were no differences in the wave velocities of the longitudinal compression and the rarefaction waves. Ginzburg and Rosenberg [1998] stipulate that the density along the shock direction must remain unchanged, while the material dilates in the lateral direction. A set of experiments conducted by Clifton et al. [1998] employed Hampden steel plates to impact soda lime glass at an oblique angle of eighteen degrees. These experiments included three experiments in which the steel was the flyer and three experiments where steel was the target. The set of experiments where steel was the flyer and glass was the target was used to examine the spall strength and residual shear strength of glass following shock compression. The objective of the other set of experiments, where steel was the target and the glass was the flyer, was to examine the state of stress behind the failure front. VISAR and normal and transverse displacement interferometers were employed to examine the free surface particle velocity histories [Clifton et al., 1998].

254

The first group of Clifton et al.s [1996] experiments was conducted using 5.6 mm thick soda lime glass targets and 4.4 mm thick steel flyer plates. These thickness choices resulted in a spall plane that is formed behind the failure front in the comminuted material. Two impact velocities were examined. At 302 m/s, or 3.3 GPa compressive stress, the failure wave did not occur and a spall strength was recorded. In the two experiments at about 392 m/s, or 4.3 GPa compressive stress, no spall signal was observed. A recompression signal indicated the arrival of the longitudinal wave reflected from the failure front. Free surface transverse particle velocity signals indicate an initial rise to near the elastically predicted transverse velocity followed by a continuous decrease in the particle velocity [Clifton et al., 1998]. Clifton et al.s [1996] experiments with the glass flyers and the steel targets were conducted at speeds such that the steel remained elastic. In the low velocity experiment, 310 m/s (3.4 GPa), the elastic prediction corresponds well to the measured normal and transverse particle velocities. It is noted that a long rise time exists for the last 10% of the transverse signal. This is markedly different from the higher velocity experiment at 397 m/s (4.4 GPa). In this experiment, the normal velocity is smaller than the elastic prediction, although the predicted velocity jump corresponding to the arrival of the unloading wave from the flyer is of the same magnitude as the elastically predicted rise. The consistency in the velocity jump magnitude indicates that the unloading response is unchanged from the elastic behavior. The signal of the transverse velocity indicates an initial rise to a level below the elastic prediction, then a drop to a lower constant level. This indicates an instability in the shear stress at high levels of shear strains or stresses [Clifton et al., 1998].

A.2. Proposed Mechanisms for the Propagation of the Failure Front Several models have been presented in the past to explain the phenomena of failure waves in glass. Espinosa et al. [1997] suggested that upon impact a deviatoric stress component drives the creation of a failure front consisting of

255

shear induced flow planes. Micro-cracking was assumed to occur along these planes, leading to microscopic fracture and the observed changes in glass properties. Feng [2000] argued that a diffusive wave was initiated at the impact surface due to the high transient stresses, and was propagated by deviatoric stresses ahead of the micro-cracks. Sundaram [1998] described a atomic bond switching model to explain the formation of the failure wave. using numeric computations [Sundaram, 1998]. Espinosa, Xu, and Brar [Espinosa et al., 1997] performed a study on glass plates and bars in which they examined the spall strength, shear resistance, and failure wave propagation. In these experiments the glasses were: soda lime glass, aluminosilicate glass, and pyrex glass. The plate experiments used both longitudinal and lateral imbedded manganin gages for normal impacts as well as normal and transverse displacement interferometers for pressure-shear impacts [Espinosa et al., 1997]. In Espinosa et al.s [1997] experiments below a compressive stress of 4 GPa, no failure wave was observed. The HEL of soda lime glass is quoted to be 6.4 GPa. Also, slightly above the HEL, at compressive stress of 7.6 GPa, ahead of the failure wave front no spall was observed and the particle velocity was observed to unload to its elastic no-spall prediction. An aluminum flyer was used in this experiment. For almost the same conditions, at a compressive stress of 7.5 GPa, and behind the failure wave front, a spall strength of 0.4 GPa was seen. The same pattern was seen at 5.7 GPa stress, which is below the HEL. Before the failure front full unloading occurs with no-spall, while behind the failure front, there is a reduced spall strength [Espinosa et al., 1997]. This paper thus suggests that the failure wave is present at stresses near the HEL. Also, the gage experiments show several details about the failure front. First the propagation speed is between 1.5 km/s and 2 km/s. Second, the shear resistance of the material is pressure dependent and is reduced behind the failure front. The longitudinal stress also decays following the failure front. The atomic structure of soda lime glass is reviewed the model is tested against experiments

256

Espinosa et al. [1997] proposed a mechanism for the failure front of shear induced flow planes. The driving force of these shear flow planes is the deviatoric stress component. Microscopic cracks, voids or other defects located at the intersection of these planes may propagate reducing the spall strength of the material. Without the micro-cracks the flow surfaces retain their cohesive strength. This model was found to be consistent with the shear and spall strength observations, and with characteristic surface features on recovered surfaces [Espinosa et al., 1997]. In a study conducted by Feng, a model is proposed which explains the failure front propagating as a wave. His model is based on three assumptions: (a) the material has a nonlinear elastic response in front of the failure wave; (b) the longitudinal stress, strain and particle velocity remain constant through the failure front; and (c) the thermal effects are insignificant [Feng, 2000]. The failure wave model proposed by Feng [2000] models the failure wave as a thin layer that propagates through the material wherein microscopic failure processes occur above a threshold stress level. Due to the lateral confinement inherent in shock compression experiments, shear dilatancy does not cause lateral expansion. Instead, voids collapse at the same time as shear dilatancy occurs. The result is that the material is comminuted, which implies fragmentation into pieces that are both fully compacted and heterogeneously deformed. The void collapsed state and the shear dilatancy, both cause a larger compressive stress in the comminuted material than would be present had the material remained intact. With the assumption of constant longitudinal stress, the lateral stress must increase resulting in a decrease in shear stress [Feng, 2000]. This phenomenon is seen in the experiments. The initiation of the failure front occurs at the impact surface. Feng suggests that initial damage occurs at the impact surface due to a combination of transient loading conditions. These transient loading conditions are caused by

257

tilt differences between the flyer and target plates, and by existing surface roughness. The resulting micro-cracking produces deviatoric strains ahead of the failure front which, in combination with existing microscopic flaws cause propagation of the failure wave. The wave is therefore diffusive on the macroscopic scale. Additionally, the modeling uses a two-dimensional simplification using an average of the percolation of micro fissures over the lateral direction. The percolation in the lateral direction is much faster than the percolation in the longitudinal direction due to the presence of more sites that meet the initiation conditions along the plane of the failure front [Feng, 2000]. The numerical model results are compared with Kanels [Kanel et al., 1992] experimental results. As indicated by Fengs theory, an increase in mean stress occurs in the comminuted material. Additionally, the two wave structure of the computed lateral stress profiles match the stress rise times and relative velocity rises from the experiment. Feng [2000] argues that the transient response of the gage may not clearly reflect the velocity of the failure front. This argument, as well as differences in the variation of the computed failure wave velocity with gage thickness suggested to Feng [2000] that this method of finding the wave velocity is not reliable. With this assumption, Feng [2000] used a measure of 5% of the second wave to estimate a wave speed of 3.32 km/s at 4.5 mm and 3.37 km/s at 6.5 mm into the material. These propagation speeds agree with measurements taken from Kanel et al.s [1992] experiments. However, Feng did not find evidence to support Kanel et al.s [1992] idea that the failure wave velocity is distance dependent. In a comparison with a different set of experimental data [Bourne and Rosenberg, 1996], Fengs [2000] calculations did show a decreasing failure front velocity with increasing depth although not as large a decrease as in the data [Bourne and Rosenberg, 1996]. Another issue discussed by Feng [2000] is that in Bourne and Rosenbergs [1996] experiments on soda lime glass a foil gage in the longitudinal direction indicates a

258

small strain increase that corresponds to the failure wave. Feng raises the question as to if this is really a decrease in impedance or just the reflection of an elastic recompression wave from the failure front. Fengs [2000] model is valid for experiments where little or no change in the longitudinal properties are recorded. He also notes that while the model predicts stress and particle velocity in soda lime glass, the inconsistency above remains to be explained [Feng, 2000]. In Sundarams [1998] experiments, a tungsten carbide flyer plate is used to impact a sandwich configuration. The front plate of the sandwich was tungsten carbide. The specimen, a 5 m layer of soda lime glass was vapor deposited on the Hampden steel rear plate. These experiments were conducted at a skew angle of 22 degrees and had velocities ranging from 118 m/s to 198 m/s. The resulting compressive stresses were 2.5 GPa, 3.5 GPa, and 5.7 GPa [Sundaram, 1998]. In Sundarams [1998] thesis, a model is proposed that links the failure wave to the shear strain and to atomistic bond rearrangements. The atomic bond changes cause the loss of shear and spall strength. The basic structure of soda lime glass is amorphous with a network of tetrahedra with one silicon atom surrounded by four oxygen atoms. A bond switching model is suggested. It is suggested that below the critical stress level, modifier ion movement at defects is responsible for causing the inelastic deformation. Other low activation energy mechanisms such as distortion of the covalent matrix also occur causing progressive hardening of the glass and the ramping of the stress. At large strains a bond switching occurs in the covalent silicon-oxygen bonds and causes stress relaxation. This switching can occur due to the decrease of potential energy barriers [Sundaram, 1998]. Sundaram [1998] states that the deviatoric and hydrostatic shock compression stress components combine to produce a phase transformation in the shocked glass. The transformation then proceeds into the material as a diffusive process because of the relaxation of the shear stresses caused by the distortion of the network of covalent bonds. Because this process is constant volume, there is no change in the longitudinal strain across the front. However, there is a distinct

259

shear stress change, which leads to an increase in the transverse stress. Additionally, this model predicts a decrease in velocity of the front with increasing depth into the material. Finally, the passage of a tensile wave through the transformed material would, Sundaram [1998] states, cause coalesce of nanocracks which would cause the material to have essentially no spall strength. The results of a numerical computation used to test this model indicate a clear decrease in shear stress and strain with increasing compressive stress. In the lowest velocity experiment (2.5 GPa), no loss in shear strength is observed. A shear stress level of 510 MPa is recorded and the end of the record corresponds to a total shear strain of 0.5. In the 3.5 GPa experiment, the shear stress increases to 580 MPa in approximately 250 ns, which is below the elastic prediction of 780 MPa. The total shear strain to this point is 2.05. After about 300 ns, the shear stress falls off to a level between 180 MPa and 100 MPa, where it remains steady. In the highest velocity shot (5.7 GPa), above the 4 GPa critical limit for failure waves, the shear stress rises to about 480 MPa before dropping off after only 150 ns to a level near 60 to 80 MPa. The shear strain to the peak is 1.97. From these experimental results, it is suggested that a critical shear strain is responsible for triggering the loss of shear strength [Sundaram, 1998]. The results for shear stress are similar to Cliftons findings [Clifton et al., 1998]. Sundarams experimental results fit well with this model for all three compressive stress experiments. The ramping initial rise and plateau in the higher two experiments is matched, as is the lower near constant behavior. Only the peak stress in the highest velocity experiment (the experiment with an impact stress of 5.7 GPa) does not quite match. Therefore, Sundaram [1998] concludes that this model explains the glass behavior under shock wave loading conditions. These three papers offer substantial insight into the mechanism of the failure wave. The models suggest micro-structural and atomic explanations for the initiation and propagation of the failure front. However, these models and experiments have not yet covered all of the regions where the failure wave occurs,

260

nor provided a single explanation that explains the entire process. The current research aims to increase the experimental knowledge base of failure waves.

261

B. Design of Experiments to Investigate the Failure Wave in Soda Lime Glass In order to build upon the research described above, a set of plate impact experiments was conducted in the present study to understand the behavior of glass under combined compression and shear impact loading. The present study, in particular, builds on Clifton et al.s [1998] work. The flyer and single target plate configuration of Cliftons study was chosen for the current experiments. In order to determine the spall strength of soda lime glass, aluminum flyers were used in both the shock compression and the combined shock compression and shear configurations. These two experiments compared the spall strengths of glass with and without the presence of shear. Next, four experiments were conducted using glass plates that were impacted with tungsten carbide flyers. Using glass specimens of two different thicknesses, changes in the normal and shear velocities and the spall strength in the undamaged and comminuted glass were examined. Also, in two experiments using a glass flyer and a tungsten carbide target, the impedance change in the glass due to the failure wave was measured. The VISAR interferometer system was used to measure the normal and transverse components of the particle velocity histories at the free surfaces of the target plates. B.1. Description of the Experimental Configurations

In Clifton et al.s [1998] paper, the single flyer and target configuration was employed. Hampden steel was used as the material for the flyer plates while the target plate was the glass specimens. The choice of steel limited the maximum stresses that could be employed as the steel was required to remain elastic. The maximum compressive stress in the experiments was 4.3 GPa, produced at an impact speed of 407 m/s [Clifton et al., 1998]. In the current study, tungsten carbide was chosen because it remains elastic when compared to steel under the high velocity impact loading. In fact, the HEL of tungsten carbide is about 7.2 0.8 GPa [Dandekar, 2004]. This, plus the greater impedance of tungsten carbide allows for the higher compressive stress of 5 GPa to be reached. A lower level of

262

3 GPa was also chosen in order to compare with the elastic results in Cliftons paper [Clifton et al., 1998]. Two experiments were conducted to obtain the spall strength, one experiment at the impact angle of eighteen degrees and the other under pure compression. Experiment Al/G1 was the normal shock compression (Figure 7.1) experiment, while Experiment Al/G2 was conducted under combined shock compression and shear loading at an angle of 18 degrees (Figure 7.2). These experiments employed aluminum 6061-T6 flyer plates of 5.9 mm thickness that impacted 12.5 mm thick glass specimens (Figure 7.3). The aluminums lower impedance of 16.6 MPa/(m/s), as opposed to tungsten carbides impedance of 106 MPa/(m/s), allowed for a larger tensile stress to be developed for the same impact velocity. By using the high normal velocity of 485 m/s, a tensile stress of around 3.5 GPa was reached, which is in excess of the recorded data which state that the spall strength is greater than 3 GPa [Kanel et al., 2002]. The tungsten carbide plates used for the other experiments only produced a maximum tensile stress of 1.2 GPa. Using 6 mm thick tungsten carbide flyers and soda lime glass targets, Experiments WC/G1 through WC/G4 examined both shock compression and shear loading. A shearing angle of 18 degrees was chosen, as in Cliftons paper [Clifton et al., 1998] to enable the study of normal and shear effects generated by the impact. The normal compressive stresses were ~ 3 GPa and 5 GPa (Figure 7.4). At 3 GPa, according to Clifton et al. [1998] the failure wave should not be present and the material should behave like an elastic brittle material. The 5 GPa experiments were performed to observe the fully developed failure wave. Both of these experiments were conduced twice, using two glass targets with different thickness of 6.5 mm and 12.5 mm. These plates were chosen so that in the experiments with the 12.5 mm thick specimen, the spall plane would occur in the target ahead of the failure wave front (Figure 7.5). In the other experiments, with the 6.5 mm thick specimen, the spall occurred behind the failure front, i.e. in the comminuted glass, following the passage of the failure wave (Figure 7.6).

263

The free surface particle velocity histories in these experiments were measured by using the VISAR. A second group of experiments were performed using soda lime glass as the flyer material and 4 mm thick tungsten carbide as the target (Figure 7.7). Experiments G/WC1 and G/WC2 were configured similar to the reverberation experiments described in Chapter VI (Figure 7.8). The 12 mm glass flyers were thick when compared with the WC target plate; three wave reverberations were possible in the target before the return of the release wave from the rear surface of the flyer plate. The key objective of these reverberation experiments was to examine the longitudinal and shear impedance of the comminuted glass near the impact surface. B.2. Modifications to the VISAR System to Enable the Simultaneous Measurement of the Normal and Transverse Components of the Particle Velocity Histories In the normal impact experiment the VISAR probe was used in the same configuration as the previous chapters. However, in all of the experiments with skew angles of eighteen degrees, both the normal and shear velocities of the specimen surface had to be examined. In order to measure these velocities, the multi-beam VISAR interferometer system with three probes was used. This is because at the high velocities generated by the impact, the conventional NDI and TDI techniques are not very suitable. The VISAR system with its variable etalons can be configured so as to resolve the expected particle velocity with enough accuracy. The mechanism to capture the shear and normal velocities was modified from the standard single probe holder discussed in Chapter IV. Issues that were considered included how to properly align multiple probes to light reflected from a single point on target surface and how to optimize the intensity of laser light to the probes. The complete laser diagnostic setup is shown in Figure 7.9. The Coherent laser was beamed into the impact chamber through a 1000 mm focal length lens and a

264

series of mirrors. This lens focused the laser light upon a grating, which was glued onto the back surface of each specimen. The cross-line gratings were from Photomechanics Inc. and had 1200 lines per millimeter. The laser beam was moved 8 mm horizontally from the center of the lens in order to allow the reflected zero beam to be separate from the input beam. The incident laser light beam was diffracted by the grating on the target plate; the zeroth order beam was reflected back normal to the target surface, while the plus 1 and the minus 1 beams were created at an angle of 39.7 degrees from the zeroth beam. The three output beams were collected by mirrors and directed out of the impact chamber. The beams were then picked up by three 120 mm focal length VISAR probes. The fiber optics then conveyed the beams to the interferometer as described in Chapter IV. In this way the interferometer recorded three particle velocity signals [Barker et al., 2000]. The measured components of the particle velocity were used to determine both the normal and the transverse components of the particle velocity during the combined pressure and shear impacts.

265

C. Experimental Results and Analysis of Failure Waves in Soda Lime Glass Each experiment exhibited unique features, which explored the effects of the failure wave on soda lime glass. The aluminum on glass spall experiments investigated the difference in the spall strength of the glass in pure shock compression (Al/G1) from the spall strength under both compression and shear loading (Al/G2). These experiments are discussed in Section C.1. The tungsten carbide impacting glass experiments (WC/G1 WC/G4) investigated the effect of the failure wave on the spall strength and also show direct evidence of the failure wave. The results are detailed in Section C.2. From the glass on tungsten carbide experiments (G/WC1, G/WC2), the impedance of the failed glass can be obtained. These results are discussed in Section C.3. C.1. Spall Strength of Glass under Shock Compression and PressureShear Loading The spall strength indicates the resistance of soda lime glass to failure in dynamic tension. Analysis of the relative impedances of various flyers showed that impacting tungsten carbide flyers against glass target plates does not produce sufficient tensile stress to cause spall. However, using Al6061-T6 flyer plates produces a dynamic tension of 3.5 GPa at a normal impact velocity of 485 m/s. In view of this, in the present study, two experiments were carried out using Al6061-T6 flyer plates against glass target plates, one with no skew angle and another at a skew angle of eighteen degrees. The normal velocity of 485 m/s was aimed for in both experiments. C.1.1. Normal Particle Velocity Calculations The 6061-T6 Al alloy has an longitudinal impedance of 16.6 GPa. The impedance is determined from its density, 2700 kg/m3 and its longitudinal wave speed, 6.15 km/s. In the present study two experiments were conducted in order to investigate the spall strength of soda-lime glass with and without the presence

266

of shear. For both experiments line equations were used to calculate the stress and particle velocity histories in the various states (Figure 7.3). The particle velocity in State 3 is calculated from the intersection of the lines drawn from State 1 in the flyer, and State 2 in the target, both prior to impact: V3 =V1ZF /(ZT +ZF ) . {VII.C.1}

Here, the soda lime glass target impedance, ZT, is 14.5 MPa/(m/s) and the flyer impedance, ZF, is 16.6 MPa/(m/s). In Experiment Al/G2, V1, is the normal component of the impact velocity, VI, and can be calculated using

V1 = VI cos .

{VII.C.2}

Here, , is the skew angle of impact and is equal to 18 degrees in the present study. In Experiment Al/G1, is zero, and the impact velocity is equal to V1.

The velocity, V3, which is also the particle velocity of the Hugoniot state can be related to the particle velocity of the rear surface of the glass in State 6, by
V6 2 =V3 .

{VII.C.3}

The stress in State 3 is simply the impedance of the glass times the particle velocity, i.e.

3 = ZT V3 .

{VII.C.4}

Calculations for the additional states are carried out in a similar fashion. A general formula is employed for each new state. This formula shows how to go from any two states to a third state. If State A is to the left and below State C on the t-X diagram and State B is the state to the right and below, then State C can be determined as follows:

267

VC =

VAZ1 + VB Z 2 + A B , Z1 + Z 2

{VII.C.5}

C = Z 2 (VC VB ) + B .

{VII.C.6}

Here, Z1, and, Z2, are the impedances of the left and right materials respectively. Using Equations {VII.C.5} and {VII.C.6}, the stresses and particle velocities can be computed for all the other states in the t-X diagrams. The equations for the important states are given below: State 4, which represents the state in the flyer after the first release: V4 = Z FV3 3 , ZF

{VII.C.7}

4 = 0 .

{VII.C.8}

State 5, which represents the stress and the particle velocity that is developed in the flyer and the target after the release wave from the free surface of the flyer plate reaches the flyer/target interface, i.e. the impact face V5 = Z FV4 , ZT + Z F

{VII.C.9}

5 = ZTV5 .

{VII.C.10}

State 6, which is the stress and the particle velocity at the free surface of the target following the reflection of the compressive wave back into the target V6 = 2V3 , {VII.C.11} {VII.C.12}

6 = 0 .

268

Stress and particle velocity in State 7, which also represents the spall plane ZT (V5 + V6 ) + 5 6 , 2ZT

V7 =

{VII.C.13}

7 = ZT (V7 V6 ) + 6 .

{VII.C.14}

The stress and particle velocity in the various states for Experiments Al/G1 and Al/G2, and the tungsten carbide on glass Experiments WC/G1, WC/G2, WC/G3 and WC/G4, are determined from these relations. The computed stress and particle velocity levels are shown in the figures for the particle velocity versus time histories as dotted lines.
C.1.2. Transverse Particle Velocity Calculations

In the combined pressure and shear experiments, the shear velocity can be calculated by simply replacing the normal component of the projectile velocity with the transverse component of the projectile velocity, that is using sine instead of cosine in Equation {VII.C.2}, and then using the shear wave speed, CS, instead of the longitudinal wave speed to calculate the shear impedance, i.e. Z S = C S . {VII.C.15}

Using the shear impedances and the transverse component of the projectile velocity instead of the normal impedances and normal impact velocity, Equations {VII.C.1} through {VII.C.14} are then used to compute the shear velocities and shear stresses in States 1 through 7. The elastic shear predictions are then compared with the actual experimental free surface particle velocities. These predictions are also shown by dotted lines on the figures, which show the experimental free surface particle velocity versus time histories.

269

The three VISAR probes record beams reflecting from the target surface at zero, +39.7 and -39.7 degrees. The later two are referred to as plus 1 and minus 1 diffracted beams, respectively. In order to determine the transverse (or shear) component of the particle velocity from the measured particle velocity histories along the directions of the plus and the minus diffracted beams, the following relations are employed: V +1 = 1 1 (1 + cos )V N + V S sin , 2 2 1 1 (1 + cos )V N V S sin . 2 2

{VII.C.16}

V 1 =

{VII.C.17}

In these equations V+1, represents the plus one order diffracted beam;V-1 represents the minus one order diffracted beam; and VN and VS represent the normal and transverse components of the particle velocity. The diffraction angle , is 39.7 degrees. By combining these relations, several combinations can be used for calculating the shear velocity. Two are used in this chapter. When all three beams are of good quality, an equation using the plus 1 and minus 1 velocities can be used, i.e.
V S = (V +1 V 1 ) ( sin ) . {VII.C.18}

This expression generates the least amount of noise in the transverse component of the particle velocity. When the plus 1 beam is of questionable quality an alternate expression can be employed
2 1 1 N V (1 + cos )V . sin 2

VS =

{VII.C.19}

270

C.1.3. Spall Strength and Particle Velocity Observations for Aluminum Impacting Glass

The spall strength experiments were conducted at a projectile velocity of approximately 500 m/s. In Experiment Al/G1, the glass specimen was impacted by an Al6061-T6 flyer at an impact velocity of 544 m/s; the resulting maximum tensile stress on the spall plane was 3.89 GPa. However, no spall was observed in this experiment (Figure 7.10). Experiment Al/G2 was conducted under combined pressure and shear loading. In this experiment the glass target was impacted at a speed of 497 m/s; as a result the glass was shocked in State 7 to a maximum tensile stress of 3.59 GPa, assuming elastic conditions. However, this stress level is not reached in the experiment, because the glass undergoes spall during the tensile loading process (Figure 7.3). As in Chapters V and VI, the spall dip and pull-back signals can be used to evaluate the spall strength under the various experimental conditions. It is to be noted that the dip in particle velocity at the time of spall was observed most clearly in particle velocity derived from the minus 1 order diffracted beam (Figure 7.11). The spall strength of the glass is determined in a similar manner as in Chapters V and VI. By examining the particle velocity dip in the particle velocity of the rear surface of the glass target in State 8, and employing the following relations, the spall strength can be calculated 1 Spall = (ZT )(Vo Vmin ) . 2

{VII.C.20}

This formula can be used rather than the formula corresponding to the pull-back height because the spall is elastic, as evidenced by the fact that the maximum predicted compressive stress at 544 m/s and 0 is 4.21 GPa, which is lower than the HEL of 6.4 GPa for glass [Rosenberg et al., 1985]. Important experimental parameters such as impact velocity, compressive stress, and spall strength are listed in Table 7.1.

271

Experiment Al/G1 was performed to examine the spall strength of glass under shock compression. In this experiment no spall signal was observed and the free surface particle velocity profile is observed to completely unload to its elastic nospall prediction (Figure 7.10). An Al 6061-T6 flyer plate was utilized in the experiment. The impact velocity was 544 m/s; this results in a maximum compressive stress of 4.21 GPa in glass, which is considerably below its HEL. From the results of these experiments it can be inferred that the lower bound for the spall strength of glass is at least 3.89 GPa. The second experiment, Al/G2, (Figure 7.11) was performed at an impact velocity of 523 m/s under a combined compression and shear loading at a skew angle of 18 degrees. The corresponding normal velocity was 497 m/s, which resulted in a compressive stress of 3.85 GPa, which is also below the HEL. In this experiment a spall strength of 3.49 GPa was measured. The free surface particle velocity profile clearly shows the shear wave arrival at 3.73 microseconds after impact. The transverse velocity is determined by combining the normal and minus one velocity signals using Equation {VII.C.19}. The peak transverse velocity is 78 m/s, which is lower than the elastic shear prediction of 159 m/s. The smaller transverse velocity is expected because similar traits occurred in the in the literature [Clifton et al., 1998]. The behavior of the transverse velocity as the impact progresses can not be determined because the spall wave arrives only 0.21 microseconds after the shear wave. These two experiments determined the difference in the spall strength of the intact soda lime glass between pure compression and combined compression and shear loading at a skew angle of 18 degrees and impact velocities around 500 m/s. The spall strength of the glass exceeds 3.89 GPa at 0 degrees and 4.21 GPa compressive stress while dropping to 3.49 GPa at 18 degrees and 3.85 GPa compressive stress. The drop in the spall strength due to a skew angle of 18 degrees shows that like ceramics, the spall strength of glass has a dependency on shear stress. The effects of the passage failure wave on the spall strength and the transverse velocity will be discussed in the Section C.2.

272

C.2.

The Effects of the Failure Wave on the Properties of Soda Lime Glass

Four combined pressure-shear plate impact experiments were performed with tungsten carbide flyer plates and soda-lime glass targets. Two of these experiments occurred at a maximum compressive stress of 3 GPa (WC/G1, WC/G2). At this stress level, the glass is expected to remain elastic and no failure wave should be present. If there is sufficient tensile stress, spallation should occur at the intersection of the two release waves which create State 7. The longitudinal wave generated by the spall event propagates to the rear surface and is measured at State 8 (Figures 7.5, 7.6). Experiments WC/G1 and WC/G2 were conducted to confirm the lack of a failure wave at this level of compressive stress. The glass specimen thickness in Experiment WC/G1 was 12.5 mm, while that in Experiment WC/G2 was 6.5 mm. With this variation in thickness, the location of the spall plane was altered from in front of the hypothetical location of the failure wave to behind it. The presence of a failure wave would cause differences in the spall strength and the transverse particle velocity in the two experiments. The other two experiments in this series (WC/G3, WC/G4) were conducted at stress levels below the HEL, but above the accepted critical stress limit for the formation of the failure wave. This critical stress has a lower limit of approximately one-half of the HEL [Ginzburg and Rosenberg, 1998]. Therefore, the critical stress could be as low as 3.2 GPa, which is one-half of the HEL of 6.4 [Rosenberg et al., 1998]. The chosen maximum compressive stress was 5 GPa. Experiment WC/G4 was designed so that the failure wave would propagate through the region of the glass where the spall should occur, prior to the spall event. The thickness of 6.5 mm was chosen for the glass specimen in order to place the spall behind the failure wave front (Figure 7.6). This thickness was the same as experiment WC/G2. Experiment WC/G3 employed the same impact velocity as experiment WC/G4, but altered the specimen thickness to 12.5 mm. This changed the spall position (Figure 7.5) so that the glass spalled first and

273

then the failure wave passed through the spalled region. However, due to insufficient tensile stress, the glass did not reach its spall strength, and thus the specimen behaved elastically.
C.2.1. Computations for the Elastic Predictions The elastic predictions for this set of experiments are found by using the line

equations as indicated on the stress versus particle velocity diagrams that are shown in Figures 7.5 and 7.6. Because the impedance of tungsten carbide, ZF, is 106 MPa/(m/s), the impacts generate a high particle velocity as well as a high compressive stress at State 3. The corresponding time versus distance diagrams show the wave propagation in the target and flyer (Figures 7.5, 7.6). As in the experiments Al/G1 and Al/G2, the impacts in experiments WC/G1 through WC/G4 are elastic, and the normal and shear waves propagate at their wave velocities, CL, and, CS, respectively. In soda lime glass, these velocities are, CL, 5.74 mm/s and, CS, 3.40 mm/s. The failure wave propagates at a slower velocity than the shear stress. The failure wave velocity is calculated from the longitudinal wave speed, the thickness of the specimen, , and the time at which the anomalous recompression is observed in the free surface velocity profile, tR.
CF = CL

t R CL 2 + t R CL 2

{VII.C.21}

Equation {VII.C.21} is based on formulae presented in Kanel et al. [2002]. Despite the larger impedance of tungsten carbide, its wave speeds are only slightly larger than those of soda lime glass; CL, is 6.99 mm/s and, CS, is 4.19 mm/s [Dandekar, 2004]. The inverse of these wave velocities are plotted as before in the t-X diagrams in Figures 7.5 and 7.6 to show the intersections of these waves during the impacts. The free surface particle velocities of experiments WC/G1 through WC/G4 are shown in Figures 7.12 through 7.15, respectively. In all of these experiments, the normal signals and the minus one signals are displaced in time. In the experiments where the signal was of sufficient strength for the plus one velocity

274

to be measured, it is also displayed. The shear signal is computed using either equation {VII.C.18} or {VII.C.19}. In each of these figures, the elastic predictions using the line equations are included for comparison.
C.2.2. Experimental Observations of Spall Strength and Transverse Particle Velocity In order to fully examine the effects of the failure wave on the soda lime glass

under shock compression and shear, the state of stress as evidenced by the free surface time histories in the material must be examined. It is evident from the literature that a failure wave in glass results in the deterioration of the shear strength, but the normal stress remains nearly constant. The deterioration in shear stress is reflected in the level of the transverse component of the particle velocity. The spall strength of the intact glass was also seen to disappear in the comminuted material behind the failure wave in the literature. In Experiment WC/G1, the specimen and target were impacted at 220 m/s, which causes a compressive stress of 2.67 GPa. This experiment was designed to be in the elastic region, with the longitudinal wave produced by the spall (the spall signal) reaching the free surface before the rarefaction wave reflected from the failure wave (the failure wave signal). No failure wave was observed in the signals in Experiment WC/G1. The free surface particle velocity signals from the normal and minus 1 beams are shown in Figure 7.12 along with the transverse velocity, which is computed in the manner discussed above from Equation {VII.C.19}. In Figure 7.12, the predicted arrival of the longitudinal compressive wave corresponds to the observed signal. The normal component of the particle velocity reaches a level of about 371 m/s after the initial shock. This velocity corresponds well to the elastic prediction of 368 m/s. The normal compression is also observed in the minus one velocity signal, with the minus one particle velocity at the longitudinal wave arrival time being reduced from the normal particle velocity signal. The difference is calculated as:

275

V 1 =

1 (1 + cos )V N . 2

{VII.C.22}

The signal, V-1, is the component of the normal particle velocity measured by the minus 1 beam before the arrival of the shear wave at the free surface. In Equation {VII.C.22}, the angle is 39.7 degrees. The shear wave also arrives at the indicated time and reduces the minus 1 particle velocity by about 30 m/s, which is in agreement with the elastic prediction of a 38 m/s drop. The transverse velocity is calculated from Equation {VII.C.19} to be on average 104 m/s, which is within 14% of the elastic value of 120 m/s. The transverse velocity remains constant for 0.17 microseconds until the rarefaction wave arrives. After this time, the transverse velocity can not be determined from the normal and the minus one beams because of the arrival of the release waves from the lateral boundary of the target. The interaction of the rarefaction wave from the rear surface of the flyer plate and the rarefaction wave from the free surface of the target plate is observed in State 8. As mentioned in Chapter V, if the resulting tensile stress is high enough, spall is produced. However, as evidenced by the lack of immediate reacceleration of the velocity following the velocity drop, the material has not spalled. The minus one beam clearly shows that the velocity corresponds to the elastic prediction (Figure 7.12). The minus one velocity drops to 76 m/s. The corresponding normal velocity drop, by reversing formula {VII.C.22}, is 86 m/s. The elastic prediction for the drop in normal velocity is 89 m/s. The close agreement between the elastic prediction and the actual velocity drop indicates that no spall has occurred. This drop is therefore distinct from the lack of reacceleration of the spall signal in silicon nitride in Experiment SC-13 from Chapter V, which does not correspond to the elastic prediction for no spall. In that experiment, the spall strength is zero. In Experiment WC/G2, the impact velocity was also 220 m/s. This resulted in a compressive stress level of 2.67 GPa. The normal, minus one and plus one beams

276

were all recorded with sufficient intensity for the velocity to be measured. The resulting free surface velocities (Figure 7.13) show an initial rise in the normal component of the free surface particle velocity to about 370 m/s, which is close to the elastic prediction of 368 m/s. At the time of the shear arrival, the plus one and minus one signals diverge in opposite directions, from the initial value, which represents the component due to the normal wave. This results in a transverse velocity wave with a peak of 109 m/s, which again is close to the elastic prediction of 120 m/s. The transverse velocity then drops off and reaches a near zero value by 0.70 microseconds. Although the material is supposed to be elastic, there is clearly a deflection visible in the normal and the minus one and plus one beams corresponding to a reflection of the compressive wave from the failed region. The signal occurs at 2.33 s. This raises the normal velocity by about 13 m/s, to around 383 m/s. Using Equation {VII.C.21} with 2.33 s as, tR, the velocity of the failure wave is computed to be 1.82 km/s. This value was used to predict the failure wave arrival time in all the other experiments. In Experiment WC/G2 the spall wave is clearly visible and has a velocity dip of 94 m/s. This corresponds well to the elastic prediction of 88 m/s. The normal velocity does not immediately reaccelerate, but rather stays the same for 0.53 microseconds. The reacceleration occurs at the elastically predicted time of the next wave reflection. This pattern corresponds to the material remaining intact and strong enough to prevent spallation. Therefore, the small change in the particle velocity recorded at the failure time of 2.33 s does not reflect the total failure of the material, but rather some non-critical damage. In Experiment WC/G3, the impact velocity is 411 m/s and the spall is designed to occur in the intact glass before the failure wave arrives. The impact produces a compressive stress level of 4.99 GPa. The normal and minus one beams are strong enough for the velocity to be determined (Figure 7.14). In the normal velocity signal, there is a slightly lower velocity in State 6 than the elastic normal

277

prediction, about 667 m/s. The elastic prediction is 687 m/s. These values are only off by 3% and is almost within the VISAR error tolerances of 2%. The velocity of the minus one beam indicates that the transverse velocity reaches a level below that of the elastic prediction of 224 m/s. The transverse velocity records a transient pulse that has a maximum value of 147 m/s. The transverse velocity is no longer accurately measured by the normal, plus one, and minus one beams after the rarefaction wave reaches the free surface, as was also the case in experiment WC/G1. The normal velocity signals dip to near the elastic level at the time when the spall wave should occur and the long time before reacceleration indicate that insufficient tensile stress has been generated to cause the material to spall. The velocity dip is 145 m/s this time. The elastic prediction is 165 m/s. There are oscillations in the measured signals at the time of the arrival of the longitudinal wave at the rear surface of the target plate after reflecting from the failure front, i.e. at time 4.36 microseconds, as estimated using the failure wave velocity calculated in Experiment WC/G2. The presence of an oscillatory signal at this time indicates that failure has occurred. These oscillations are particularly evident in the minus one beam where the velocity rises to about 30 m/s. In the normal beam, the oscillations only change the particle velocity by approximately 5 m/s. Thus, this failure wave is evident in the transverse motion of the plate. Strangely, the failure wave did not greatly affect the normal velocity signal at this time. Experiment WC/G4, with its thinner target plate exhibits a failure wave that clearly affects the spall strength. In this experiment the impact velocity is 408 m/s, which corresponds to a compressive stress of 4.96 GPa. The normal signal corresponds well to the elastic prediction (Figure 7.15). The measured particle velocity is 686 m/s and the elastic prediction is 683 m/s. The minus one signal shows a larger discrepancy between 632 m/s, the measured value and 603 m/s, which is the elastic prediction. However, this is still only a 5% difference. The

278

transverse velocity elastic prediction is 222 m/s. The actual transverse velocity only reaches 148 m/s and it rises and falls over a period of 0.28 microseconds. The failure wave occurs at, 2.34 microseconds, and is apparent as an increase in the particle velocity of both the normal and the minus one signals. The time of this increase approximately corresponds to the time of arrival of the failure wave using the failure wave velocity estimate of 1.82 km/s, as obtained from experiment WC/G2. This increase is of the order of 45 m/s for the normal component of the particle velocity. This increase is indicative of a reduction in impedance of the material behind the failure wave. Using Equation {VII.C.21}, and the actual time at which the velocity rise begins, the velocity of this failure wave is calculated to be 1.63 km/s for this experiment. No spall signal is observed indicating that the failed material has no resistance to the generated tensile stress, and thus zero spall strength. The resulting velocity signals show no wave reflections following the spall arrival, which confirms that the glass has separated at the spall plane.
C.3. Observation of the Impedance in the Comminuted Material The glass on tungsten carbide experiments (G/WC1, G/WC2) were performed

using the reverberation method discussed in Chapter VI to observe the change in the impedance of glass in to the presence failure wave (Figure 7.7). The glass flyers were three times as thick as the target plates in these experiments. The Hugoniot state and two reverberations were measured before the arrival of the flyer unloading wave. The experiments were conduced using 3.04 GPa and 5.51 GPa impact stresses. Experiment G/WC1, which was conducted at 3.04 GPa of impact stress, did not show any failure wave and the normal velocity profile corresponded well to the elastic prediction (Figure 7.16). The 5.51 GPa experiment (G/WC2) had a failure wave, and the resulting change in the normal velocity of the first reverberation is visible in Figure 7.17. Also, the normal and transverse velocities do not conform to the elastic predictions.

279

C.3.1. Elastic Computations for the Internal Stresses and the Acoustic Impedance in Experiments G/WC1 and G/WC2

The reverberation experiments can be analyzed in the same manner as the reverberation experiments in Chapter VI. In Figure 7.8, the stress versus particle velocity diagrams and time vs. distance diagrams for Experiments G/WC1 and G/WC2 are given. The only new equations are the equations for determining the stresses and the acoustic impedance from the free surface particle velocities. In these equations, the flyer impedance, ZF, refers to the soda lime glass and the target impedance, ZT, refers to the tungsten carbide. For the elastic calculations, State 1 is the flyers normal impact velocity, as found in Equation {VII.C.2}. State 2 is a state of zero stress and zero particle velocity in the target, prior to impact. This means that to reach the same stress level in the Hugoniot state as the tungsten carbide on glass experiments (WC/G), the material must be impacted at the same velocity. To prove this, the general line Equations {VII.C.5} and {VII.C.6} can be used to calculate the stress and particle velocity in State 3: V3 = Z FV1 , ZT + Z F

{VII.C.23}

3 =V1 ( ZT ZF ) /(ZT +ZF ) .

{VII.C.24}

This is the same as substituting Equation {VII.C.1} into Equation {VII.C.4}. The other states leading up to the velocity equation for State 6, where the impedance is measured are also determined by the application of the line equations. State 4, which follows the reflection of the compressive wave from the free surface of the specimen is V4 = 2V3 , {VII.C.25} {VII.C.26}

4 = 0 .

280

State 5 is computed from States 3 and 4, V5 = ZTV4 + Z FV1 , ZT + Z F

{VII.C.27}

5 = Z F (V1 V5 ) .

{VII.C.28}

And State 6, which is the state on the free surface of the specimen following the first reverberation is: V6 =

5 + ZTV5
ZT

{VII.C.29}

6 = 0 .

{VII.C.30}

These equations are then used to determine the change in the impedance of the material due to the passage of the failure wave. The values for the normal impact velocity, V1, and the two measured free surface velocities, V4, and, V6, are used. Rearranging and combining Equations {VII.C.27} through {VII.C.29} to solve for, ZF, yields:
ZT (V6 V4 ) . ( 2V1 V4 V6 )

ZF =

{VII.C.31}

The shear impedance of the flyer is computed in the same manner as Equation {VII.C.31}, except that the shear impedance of the target and the shear velocities are used in place of the longitudinal values.
C.3.2. Measurements of the Normal and Transverse Particle Velocities and the Impedance Change Due to the Failure Wave

281

Experiment G/WC1 was conducted at an impact speed of 250 m/s. This resulted in a maximum compressive stress of 3.04 GPa. The normal components of the particle velocity profiles are closely matched with the elastic predictions. In fact, the estimated velocity in the Hugoniot state of the tungsten carbide is 57 m/s, which is the same as the elastic prediction. The first reverberation has a velocity of 101 m/s and the elastic prediction is 102 m/s. The particle velocity in the second reverberation is 134 m/s, which is also the elastic prediction. Unlike the normal particle velocity, the velocities of the different states of the plus one and minus one beams do not agree with the elastic predictions for these velocities. In fact, the velocities of these two signals are closer to the average of their elastic predictions. This is because, the plus one and minus one beams deflect less in the presence of shear than is elastically predicted in all the cases. As a result of the plus one and minus one beams not matching the elastic predictions, the transverse velocity only reaches 10 m/s at about 1 microsecond, which is lower than the elastic prediction of 18.4 m/s. This velocity slowly ramps downwards and reaches 2.5 m/s at 1.7 microseconds. In the 454 m/s (5.51 GPa) experiment, G/WC2, the Hugoniot velocity in the tungsten carbide occurs at the predicted time of 0.57 microseconds, but only reaches 95 m/s, instead of the elastic prediction of 104 m/s. The normal component of the particle velocity in the reverberations also exhibit similar behavior. The first reverberation only exhibits a velocity of 167 m/s instead of the elastic prediction of 185 m/s. The second reverberation is 234 m/s, which is lower than the elastic prediction of 244 m/s. The third reverberation is 274 m/s instead of the elastic prediction of 289 m/s. Again, the plus one and minus one signals are less than their elastic predictions . The transverse component of the particle velocity peaks at 5.9 m/s instead of 33.42 m/s, which is the elastic prediction. As in experiment G/WC1 there is a slow drop off from this peak. The transverse velocity takes 0.63 microseconds to return to the zero level.

282

In Figures 7.16 and 7.17, the first reverberation is clearly visible. In experiment G/WC1, using Equation {VII.C.31} results in a normal impedance of 13.6 MPa/(m/s), which is within 0.9 MPa/(m/s) of the elastic predicted value. The shear impedance is 5.94 MPa/(m/s), which is 2.66 MPa/(m/s) lower than the elastic value of 8.6 MPa/(m/s). In experiment G/WC2, once the reflection of stress waves from the failed material reaches the rear surface, there is a change in the particle velocity of 45 m/s. If the velocity for State 6 before this velocity change is used, the normal impedance found from Equation {VII.C.31} is 13.0 MPa/(m/s). This is slightly lower than that observed in experiment G/WC1. However, the shear impedance is 4.24 MPa/(m/s). This is 49% of the elastic shear impedance. These results suggest that the failure wave causes only a small change in the impedance and the longitudinal wave speed in the failed glass, but a substantial change in the shear impedance.

283

D.

Conclusions and Discussion

Based on the eight experiments conducted in this chapter, several observations can be made that extend the existing knowledge of failure waves. These conclusions cover the effect of the failure wave on the shear strength, spall strength, and the impedance of soda lime glass. Also, the spall strength sensitivity to change in the impact skew angle of soda lime glass is discussed. The spall strength of the material has been determined from Experiment Al/G1 to be larger than 3.89 GPa at 544 m/s impact velocity in the absence of shear stress. This is in agreement with previous results, such as Kanel [Kanel et al., 2002] and Brar [Brar et al., 1991 B], where the spall strength was found to be in excess of 3 GPa. The spall strength at a normal velocity of 497 m/s in the presence of shear due to a skew angle of 18 degrees was found to be 3.49 GPa in experiment Al/G2. This result indicates that spall strength is lost with the application of shear stresses, as in ceramic materials like AS800 silicon nitride. Recent unpublished work by Dandekar [2005] indicates the spall strength of glass under shock compression is around 3.5 GPa. This suggests that the spall strength does not change with shear stress. Further research is necessary to clarify this point. The spall strength is clearly altered behind the failure front. In the tungsten carbide impacting glass experiments (WC/G), ahead of the failure wave the spall strength is not degraded. This is evident because insufficient tensile stress exists to cause spall and the elastic reflections are clearly visible in the free surface velocity profiles. In experiment WC/G1 (220 m/s impact velocity), the velocity drops 86 m/s at the time where a spall signal should occur, and in experiment, WC/G3 (411 m/s impact velocity), it drops to 145 m/s. In these experiments tensile stresses of 644 MPa and 1.20 GPa, were generated. Clearly neither experiment has a tensile stress as large as the 3.5 GPa spall strength. Thus, the material behaves elastically, and no spall occurs. The lack of spall is confirmed by the observation that the material ahead of the failure wave is intact for

284

signals to propagate through. In Experiment WC/G3, a failure wave occurs and is measured by oscillations in the minus one velocity signal. As predicted, the spall strength of soda lime glass disappears behind the failure wave. In experiment WC/G2, instead of a spall signal elastic wave reflection occurs similar to that observed in the two experiments WC/G1, WC/G3. The velocity drops by 94 m/s, which is close to the elastic prediction of 88 m/s. However, some damage to the material occurs, as indicated by the small rise in the particle velocity seen at the approximated failure wave arrival time. In experiment WC/G4, at 408 m/s impact velocity, the material failed completely as zero spall strength is recorded and a larger rise in the particle velocity of 45 m/s is observed at the failure wave arrival time. The average failure wave velocity in these experiments is 1.7 0.1 km/s. The transverse particle velocity profiles are summarized in Figure 7.18. In the experiment WC/G1 the transverse velocities are close to the elastic predictions. The transverse free surface particle velocity is on an average 104 m/s which is lower than the elastic prediction of 120 m/s by 14%. The value is steady until the arrival of the rarefaction wave 0.39 microseconds later. Experiment WC/G2 has smaller normal velocities than the elastic predictions. The peak transverse velocity is 109 m/s, which is within 9.2% of 120 m/s, which is the elastic prediction. The transverse velocity decreases to zero after 0.7 microseconds. There are no major differences between the transverse velocities in the two experiments. This is also consistent with the results of Clifton et al. [1998] for the transverse component of the particle velocities. In measurements of transverse velocity, at higher impact velocities, the glass material behaves differently from previous results at high velocities [Clifton et al., 1998]. In the current experiments, the transverse velocity is about 30% smaller than the elastic predictions. Experiment WC/G3 has a transverse velocity that exhibits a transient pulse with a peak at 147 m/s, which is 34% smaller than the elastic prediction of 224 m/s. The transverse particle velocity peaks just before

285

the arrival of the rarefaction wave. In experiment WC/G4, the peak in the transverse particle velocity is 148 m/s, which is 33% less than the elastic prediction of 222 m/s. This particle velocity pulse lasts 0.28 microseconds before dropping to a level of about 67 m/s. The transverse component of the particle velocity is less than the elastic prediction in both cases. In Cliftons experiments, at 4.3 GPa compressive stress [Clifton et al, 1998], the transverse velocities reached a level near the elastic prediction and then fell off continuously. This is different from the behavior seen in Figure 7.18, where the peak value is 30% smaller than the elastic estimate. Of course, the current high velocity experiments are conducted at a higher stress level than conducted by Clifton. The experiments where the glass impacts the tungsten carbide also exhibit similar trends. The normal velocity of experiment G/WC1, is predicted well by the elastic equations. The transverse velocity reaches 10 m/s, which is down by 45% from 18.4 m/s, the elastic prediction, and then it falls slowly. In experiment G/WC2, on the other hand, the normal velocity is distinctly below that of the elastic prediction. In the Hugoniot state, the normal velocity is 95 m/s instead of the elastic value of 104 m/s. The peak transverse velocity is 5.9 m/s, which 82% lower than the elastic value of 33.42 m/s. This inconsistency between the experimental and elastic values is much greater than the mismatch in the WC/G experiments. This large transverse velocity is also different from the research done by Clifton et al. [1998]. For these types of experiments, the transverse velocity was the same as the elastic prediction for the experiment at 3.4 GPa compressive stress and decreased from a peak level to a lower plateau at 4.4 GPa [Clifton et al., 1998]. Neither behavior is exhibited in the current experiments. The impedance of the material decreases slightly from Experiment G/WC1 to Experiment G/WC2. The elastic material impedance is estimated to be 14.5 MPa/(m/s). The calculated impedance in Experiment G/WC1 is 13.6 MPa/(m/s), which is 0.9 MPa/(m/s) lower than the elastic value. The normal particle velocity profile fits the elastic prediction in experiment G/WC1. The difference in the calculated impedance from experiment G/WC1 to experiment

286

G/WC2 is only 0.6 MPa/(m/s). This difference, though present, and apparently dependent on impact velocity, is only 10% at 5.51 GPa. This suggests that the assumption of constant longitudinal properties across the failure front is nearly valid.

287

References
Barker L.M, Barker V.J., Barker Z.B., 2000. Valyn VISAR Users Handbook. Albuquerque, New Mexico, USA. Bourne, N. K. and Z. Rosenberg., 1996. Shock Compression of Condensed Matter 1995. American Institute of Physics. 567. Brar, N. S., S. J. Bless, and Z. Rosenberg., 1991 A. Impact-Induced Failure Waves in Glass Bars and Plates. Applied Physics Letters. Vol. 59. No. 26. 3396-3398. Brar, N. S., Z. Rosenberg, and S. J. Bless., 1991 B. Spall Strength and Failure Waves in Glass. Journal de Physique IV. Vol 1. C3. 639-644. Clifton, R. J., M. Mello, and N. S. Brar., 1998. Effect of Shear on Failure Waves in Soda Lime Glass. Shock Compression of Condensed Matter 1997. American Institute of Physics. 521-524. Dandekar, D. 2004. Spall Strength of Tungsten Carbide. U.S. Army Research Laboratory. ARL-TR-3335. Dandekar, D. 2005. Unpublished measurements of the spall strength of glass. Espinosa, H. D., Y. Xu, and N. S. Brar., 1997. Micromechanics of Failure Waves in Glass: I, Experiments. Journal of the American Ceramic Society. Vol. 80. No. 8 2061-2073. Feng, R., 2000. Formation and Propagation of Failure in Shocked Glasses. Journal of Applied Sciences. Vol. 87. No. 4. 1693-1700. Ginzburg, A. and Z. Rosenberg, 1998. Using Reverberation Techniques to Study the Properties of Shock Loaded Soda-Lime Glass. Shock Compression of Condensed Matter 1997. American Institute of Physics. 529-531. Kanel, G. I., S. V. Rasorenov, and V. E. Fortov., 1992. The Failure Waves and Spallations in Homogeneous Brittle Materials. Shock Compression of Condensed Matter 1991. Elsevier Science Publishers. 451-454. Kanel, G. I., A. A. Bogatch, S. V. Razorenov, and Z. Chen., 2002. Transformation of Shock Compression Pluses in Glass Due to the Failure Wave Phenomena. Journal of Applied Physics. Vol. 92. No. 9. 5045-5052.

288

Rosenberg, Z., S. J. Bless, and D. J. Yaziv., 1985. Journal of Applied Physics. Vol. 58. 3249. S. Sundaram, 1998. Pressure-Shear Plate Impact Studies of Alumina Ceramics and the Influence of an Intergranular Glassy Phase. Doctoral Thesis. Brown University.

Tables
Table 7.1: Experimental parameters for the experiments studying the spall strength of glass (Al/G) and those studying the effect of a failure wave on the spall strength and transverse (Trans.) velocity of glass (WC/G). This table includes both the experimental data and elastic estimates. Exp # Al/G1 Al/G2 WC/G1 WC/G2 WC/G3 WC/G4 Flyer Type Aluminum Aluminum WC WC WC WC Specimen Thickness 12.5 mm 12.5 mm 12.5 mm 6.5 mm 12.5 mm 6.5 mm Impact Angle 0 18 18 18 18 18 Impact Velocity 544 m/s 523 m/s 220 m/s 220 m/s 411 m/s 408 m/s Hugoniot Stress 4.21 GPa 3.85 GPa 2.67 GPa 2.67 GPa 4.99 GPa 4.96 GPa Peak Trans. Vel. n/a 78 m/s 104 m/s 109 m/s 147 m/s 148 m/s Elastic Trans. Vel. n/a 159 m/s 120 m/s 120 m/s 224 m/s 222 m/s Spall Strength >3.89 GPa 3.49 GPa n/a n/a n/a 0

Table 7.2: Experimental parameters for glass on tungsten carbide experiments. Note, all velocities and stresses are in the tungsten carbide. Only the Hugoniot velocity and stress are given.. Normal Velocity Exp Number G/WC1 G/WC2 Impact Angle 18 18 Impact Velocity 250 m/s 454 m/s Hugoniot Stress 3.04 GPa 5.51 GPa 13.6 MPa/(m/s) 13.0 MPa/(m/s) Impedance Exp. Hugoniot 57 m/s 95 m/s Elastic Hugoniot 57 m/s 104 m/s Transverse Velocity Exp. Hugoniot 10 m/s 5.9 m/s Elastic Hugoniot 18.4 m/s 33.4 m/s

289

Figures

Figure 7.1: Specimen configuration for normal spall strength experiment. The impact surfaces are normal to the direction of motion. The laser beam from the VISAR probe reflects off the aluminum coating on the free surface of the soda lime glass.

Figure 7.2: Specimen configuration for pressure-shear spall strength experiment. The impact surfaces are at 18 degrees to the direction of motion. A grating creates three output laser beams.

290

Figure 7.3: Spall strength Experiments Al/G1 and Al/G2. (a) Time vs. distance diagram including pressure and shear wave propagation. (b) Stress vs. velocity diagram including the effects of spall strength.

Figure 7.4: Specimen configuration for WC/Glass experiments. This configuration is similar to Experiment Al/G2, except that tungsten carbide replaces the aluminum as the flyer material. The result is a greater compressive stress in the Hugoniot state, but a smaller tensile stress in State 7.

291

Figure 7.5: (a) Time vs. distance diagram and (b) Stress vs. velocity diagram for spall ahead of the failure front in experiments WC/G1 and WC/G3. In these experiments, the geometry is such that any spallation occurs in the intact material.

Figure 7.6: (a) Time vs. distance diagram and(b) Stress vs. velocity diagram for spall behind the failure front in experiments WC/G2 and WC/G4. The thickness of the specimen is smaller than in Figure 7.5, so that any spall takes place in the comminuted glass.

292

Figure 7.7: Specimen configuration for the glass flyer and WC target experiments. In these experiments, the velocity measurements on the tungsten carbide are used to calculate the stresses inside of the soda lime glass.

Figure 7.8: (a) Time vs. distance diagram showing longitudinal and shear waves and (b) Stress vs. velocity diagram for glass flyer and tungsten carbide specimen experiments. Three reverberations are indicated on the S-V diagram.

293

Figure 7.9: Experimental measurement technique using three VISAR probes for normal and transverse velocity measurements and velocity measurement system. The green (535 nm wavelength) laser beam passes through a 1000 mm focal length lens and reflects off of a mirror (M1) before reaching the specimen. The grating produces three beams, one normal and two at 39.7. Because the input laser beam passes through the lens at 8 mm from the center of the lens, the zero order reflected beam from the grating actually is reflected at a slight angle. Mirrors M2, M3, and M4 reflect the output laser beams to points outside of the impact chamber. Three VISAR probes collect these beams. The velocity system uses three red laser beams perpendicular to the path of the projectile. These beams are cut in sequence and as they are, the measured intensity at the photodiode decreases. A time record of this intensity is used to determine the impact velocity.

294

0.6 Normal 0.5

Velocity (km/s)

0.4

0.3

0.2

0.1

0 1E-06

2E-06

3E-06

4E-06

5E-06

6E-06

Time (seconds)
Figure 7.10: Spall experiment Al/G1 at an impact velocity of 544 m/s and an impact angle of zero degrees. The actual velocity follows the elastic predictions for Hugoniot state and State 8.

295

0.6 Normal Minus 1 Transverse

0.5

Velocity (km/s)

0.4

0.3

0.2

0.1

0 1E-06

2E-06

3E-06

4E-06

5E-06

6E-06

Time (seconds)
Figure 7.11: Spall experiment Al/G2 at an impact angle of 18 degrees and an impact speed of 523 m/s. The normal beam signal is lost during the spall, but the rebound of the minus one beam indicates that spallation is occurring. The elastic levels, shown by the dotted lines agree closely with the actual results of the Hugoniot state. The difference between the minimum spall velocity and the elastic prediction for State 8 is another indication that spall is occurring.

296

0.75 0.7 0.65 0.6 0.55 0.5 0.45 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 1E-06 2E-06 3E-06 4E-06 5E-06 6E-06
Arrival of Normal Wave Arrival of Shear Wave Arrival of Spall Wave

Normal Minus 1 Transverse

Velocity (km/s)

Failure Wave

Time (seconds)
Figure 7.12: Experiment WC/G1 with an impact velocity of 220 m/s. The predicted arrival of the spall signal is before the time a failure wave could reach the surface. The elastic predictions agree well with the normal and transverse velocities. No failure wave or spall effects occur. Instead, the elastic wave interaction causes a drop in particle velocity. No velocity change is seen at the time of arrival of a longitudinal wave reflecting from the failure front.

297

0.75 0.7 0.65 0.6 0.55 0.5 0.45 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0 2E-06 4E-06
Shear Wave Failure Wave Spall Wave Normal Wave

Normal Minus 1 Plus 1 Transverse

Velocity (km/s)

Time (s)
Figure 7.13: Experiment WC/G2 has an impact speed of 220 m/s. The elastic predictions for the normal and transverse particle velocities predict the actual behavior well. In this experiment, a small upturn in the material upon is seen at the predicted time of arrival of the longitudinal rarefaction wave reflecting from the failure front (the failure wave). This wave arrival is before the predicted arrival time of the elastic interaction, which if the tensile stress exceeded the spall strength, would be the spall wave.

298

0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 1E-06 Normal Minus1 Transverse Shear
Wave Arrival Normal Wave Arrival Failure Wave Spall Wave Arrival

Velocity (km/s)

2E-06

3E-06

4E-06

5E-06

6E-06

Time (seconds)
Figure 7.14: Experiment WC/G3 has a velocity of 411 m/s, and the elastic wave arrival at the spall wave arrival time is before the failure wave arrival. As in the lower velocity experiments (WC/G1, WC/G2) the material has greater spall strength than the tensile stress that the experiment produces. Therefore, the material behaves elastically. Vibrations in the minus 1 beam appear at the time, which a longitudinal reflection from the failure wave should reach the free surface.

299

0.7 0.6 0.5 0.4 0.3 0.2 0.1 0


Shear Wave Arrival Normal Wave Arrival Approx. Failure Wave Arrival Spall Wave/ Reflected Wave Arrival

Velocity (km/s)

Normal Minus 1 Transverse

1E-06

2E-06

3E-06

4E-06

5E-06

Time (seconds)
Figure 7.15: Experiment WC/G4 shows the arrival of the failure wave, which is the rarefaction wave reflected from the failure front, and no spall signal at an impact speed of 408 m/s. The normal and minus 1 beams show a drop followed by a rise in particle velocity at the time of arrival of this failure wave.

300

200 175 150 Normal Plus 1 Minus 1 Transverse

Velocity (m/s)

125 100 75 50 25 0
Elastic Shear Prediction

Plus 1 Elastic

Elastic Normal Prediction

Minus 1 Elastic

1E-06

2E-06

3E-06

4E-06

Time (seconds)
Figure 7.16: Experiment G/WC1 shows the reverberations in the tungsten carbide at an impact velocity of 250 m/s. The elastic prediction for the normal velocity is followed precisely by the experimental data. However, the deflections in the plus 1 and minus 1 beams due to shear are less than the elastic predictions. As a result, the transverse velocity is smaller than predicted.

301

0.3 Normal Plus1 Minus 1 Shear 0.2


Plus 1 Elastic Minus 1 Elastic

Velocity (km/s)

Normal Elastic

Failure Wave From Back Of Failed Material

0.1

Shear Elastic

1E-06

2E-06

3E-06

4E-06

Time (seconds)
Figure 7.17: Experiment G/WC2 which was impacted at a speed of 454 m/s shows the reverberation in the tungsten carbide. The normal velocity from the experiment is lower than the elastic prediction. The transverse velocity is also much lower than elastically predicted.

302

150
WC/G3 WC/G4 WC/G1 WC/G2 G/WC1 G/WC2

Velocity (m/s)

100

50

2.5E-07

5E-07

7.5E-07

Time (seconds)
Figure 7.18: Transverse velocity measurements for tungsten carbide on glass and glass on tungsten carbide experiments. The transverse velocities in Experiments WC/G1 and WC/G2 are smaller than the elastic predictions by between 14% and 9.2% respectively. A larger discrepancy occurs in experiments WC/G3 and WC/G4, where the experimental velocities are off by 34% and 33% respectively. The glass impacting tungsten carbide experiments have the smallest transverse velocities compared to the elastic predictions. Experiment G/WC1s transverse velocity is off by 45%, while Experiment G/WC2s transverse velocity is off by 82%.

303

Chapter VIII Summary of Experimental Analyses


In this thesis, two materials were studied: soda lime glass and AS800 grade silicon nitride. Soda lime glass is often used in windows of military vehicles and aircraft where integrity in the event of shrapnel impacts is of vital concern. Soda lime glass is also commonly used as a model material, in order to study the behavior of glasses and glass ceramics under static and dynamic loading conditions. AS800 grade silicon nitride is considered one of the leading material candidates for the next generation of aircraft engine turbine blades because of its superior high temperature properties when compared with nickel based superalloys. Silicon nitride has a higher melting temperature [Brandes and Brook, 1999; Matweb, 2005] better creep resistance, at elevated temperatures [Lin et al., 2001; Brandes and Brook, 1999] than nickel based superalloys. Also, this grade of silicon nitride has superior particle impact resistance than other silicon nitride grades [Choi et al., 2002]. The environment of a turbine includes many unburned fuel particles and other small pieces of debris traveling at high velocity, so the resistance of these blades to high velocity impacts is of critical importance. In order to examine the impact of these brittle materials, both planar shock compression and particle impact studies were conducted. Understanding the full extent of stress wave propagation in three dimensions requires both experimental observations and numerical modeling. A projectile accelerator system was designed specifically for the particle impact experiments. The particle impact experiments on soda lime glass and the numerical study of these impacts [Nathenson et al., 2005] provide insight into the three dimensional features of the impacts. Planar shock compression and pressure-shear experiments enable the determination of quantifiable measures of dynamic strength such as the Hugoniot Elastic Limit and the spall strength. Many factors influence the material including impact velocity, skew angle, and previous shock damage. Shock compression experiments examine the properties of silicon nitride and glass under these conditions.

304 A. Particle Impact Experiments on Soda Lime Glass In order to conduct the particle impact experiments on soda lime glass, a projectile firing system was designed and constructed. The design is described in Chapter II. This system uses compressed air to accelerate a sabot and projectile arrangement down a 4.5 foot long, inch diameter barrel. The sabot and projectile are separated by means of a sabot stripper and the spherical 1/16th inch diameter projectile impacts the plates of soda lime glass at velocities varying from 150 m/s to 350 m/s. This impact velocity was determined using a laser velocity measurement. Four plate thicknesses were tested, 3 mm, 5 mm, 15 mm, and 25.4 mm. These experiments were monitored both with surface strain gages and with high speed camera imagery. The strain gages monitored both the radial and hoop strains on the impact surface at a predetermined point, which was approximately 10 mm from the impact location, on the specimen surface. The experiments and their results are detailed in Chapter III. Stress wave theory predicts three types of waves within a elastic solid during a typical three dimensional impact event. Two of these waves travel within the material, expanding in a hemispherical pattern; a longitudinal compressive wave and a shear wave. In soda lime glass, these waves travel at velocities of 5740 m/s and 3400 m/s. On the surface of the material, the Rayleigh wave travels at about 90% of the shear wave speed. The Rayleigh wave has a cylindrical wave front. These three waves create a particle velocity field and a stress field in the material. The effect of these fields can be seen on the surface from strain gage measurements and deduced throughout the interior of the material by means of the numerical simulations. The numerical work using LS-DYNA 3D and performed by Guodong Chen [Nathenson et al., 2005] built on the experiments performed in this thesis. The experimental impact velocities and contact times were used to calibrate the impact model. Hertzian elastic theoretical models for the peak force and contact time also confirmed the models accuracy. For example, the contact times of the

305 Hertzian theoretical equations, numerical modeling, and experimental observations all agreed to within 3%.

The LS-DYNA numerical model of an infinite half space indicates the arrival of a single strain pulse that first exhibits tension and then compression followed by smaller reverberations. The nature of this wave concurs with the form of the theoretical estimates of the surface strain wave [Mitra, 1964]. The numerical simulations of the experiments using LS-DYNA 3D, that take into account the finite thickness and infinite lateral boundaries of the specimens indicate an effect with decreasing thickness. Specifically, the tensile part of the strain pulse is intensified, the compressive part disappears, and the reverberations change form. The alterations in the strain pulse occur at or after the time of the wave arrival from the rear surface. The addition of lateral boundaries only changes the late time reverberations after the boundary waves have returned to the measurement location [Nathenson et al., 2005]. The elastic strains computed in LS-DYNA at the measurement location have features both similar to and distinct from the experimental data. The experimental strain gage signals record a single pulse in tension. This pulse is an order of magnitude smaller than the elastic simulations, reflecting the limiting fracture strength of the actual glass. The LS-DYNA numerical model does not incorporate cracking. The structure of this tensile pulse varies depending on the thickness of the specimen more than the velocity. The thicker specimens, 15 mm and 25.4 mm, exhibit a smaller magnitude tensile pulse than the thinner specimens, 3 mm and 5 mm. Additionally, the thin experiments have a longer pulse duration due to the crack propagation. With regards to the experiments undertaken to measure rear surface strains, the cracking patterns were observed to vibrate. The strain records in both of these experiments clearly indicated oscillations whose periods match the cracking pattern vibrations. The LS-DYNA simulations also show oscillations with this approximate period, especially in the

306 thinner specimens. This suggests that the internal stresses and surface strains are related to the oscillations in the cracking patterns. A kinetic analysis of the partitioning of energy of impact into various dissipative modes was performed using the coefficient of restitution calculated via stress wave theory, the LS-DYNA numerical model, and the experimental camera images. Stress waves and vibrations take up most of the elastic energy. The experimental coefficients of restitution are much smaller than those obtained from the LS-DYNA simulations or the theoretical modeling. Therefore, it is surmised that cracking is responsible for a larger fraction of energy loss when compared with the elastic processes. The amount of energy lost due to cracking and elastic modes increases with impact velocity, with the larger percentage of energy being dissipated in cracking. In the experiments four types of cracking patterns were observed in the material. Cone cracks were observed in the thin specimens at all impact velocities. They became more prominent with increasing impact velocity, ejecting from the rear surface in some cases. Radial cracking from both the front and rear surface was seen in the thinner specimens. These cracks also became more prominent with increasing impact velocity. Lateral cracks were not recorded in the thinnest specimens, but occurred in most of the other experiments. These cracks were seen to increase in size and fullness with impact velocity, but not thickness. In almost every experiment, surface chipping occurred. It also increased in diameter with impact velocity but not thickness. The initiation of splinter cracks appears to be suppressed in the thinner specimens at higher impact velocities. Of the four modes of cracking observed in the experiments and in previous experiments, radial cracks are the most critical because they propagate laterally through the specimen. The conical and lateral cracks both can cause local material ejection, so they are more deleterious to the material than the splinter cracks, which only propagate for a distance on the order of 1 mm.

307 It is evident from the above study that the LS-DYNA numerical elastic model predicts the experimental behavior well, except for the effects of cracking. This has allowed several observations to be made. The propagation of stress waves on the surface has been observed to be limited by the finite failure strength of the glass. The partitioning of energy during the impact indicates energy dissipation by vibrations as well as stress waves and cracking. The observed cracking patterns have also been shown to vary with thickness and with impact velocity and their relative damage potential has been discussed. These observations provide a guide to the behavior of soda lime glass, and by extension other glasses and glass ceramics of the same nature, under three dimensional high speed particle impact.

308 B. Planar Shock Compression and Pressure-Shear Experiments on AS800 Grade Silicon Nitride The planar nature of the stress wave propagation in shock compression experimentation allows for the use of one dimensional stress wave theory. The configuration of the single stage gas gun and the preparation of experimental specimens are discussed in Chapter IV. From these experiments, the free surface particle velocity is observed. This in turn allows the internal stress state, the Hugoniot Elastic Limit (HEL), and the spall strength of Si3N4 to be evaluated. In the experiments on AS800 grade silicon nitride, discussed in Chapter V, impacts were carried out under pure compression and combined pressure-shear loading conditions. The shock compression experiments occurred at impact speeds varying from 65 m/s to 546 m/s. The highest velocity experiments, combined with tungsten carbide flyer plates allowed the elastic-plastic regime of Si3N4 to be examined and the HEL to be found. Also, experiments were conducted at impact velocities between 115 m/s and 300 m/s at a skew angle of 12 degrees to examine the effects of combined a pressure and shear shock loading on the residual spall strength of Si3N4. The shock vs. particle velocity and stress vs. strain relationships for silicon nitride were determined. Below the HEL, the shock velocity in the material remains nearly constant. In fact, the average shock velocity of 10.7 km/s is within 1.8% of the elastic wave speed of 10.9 km/s, which is computed from the elastic modulus, Poissons ratio, and density. At the highest particle velocities, the shock speed is reduced indicating the onset of plasticity. However, insufficient data exists to generate a trend in this elastic-plastic region. In observing the stress-strain behavior, the elastic Hugoniot line and the hydrodynamic curve diverge near the onset of plasticity. The Hugoniot Elastic Limit for the material is calculated from the elastic equations using the highest impact velocity Experiment SC-6, and was found to be ~ 12 GPa, which agrees well with values found in the literature, such as 12.1 GPa [Nahme et al., 1994]. The HEL indicates the onset of inelastic behavior in the material under

309 conditions of uniaxial shock compression loading. Using the HEL, the dynamic yield strength under plane stress was estimated to be 7.6 GPa. The spall strength, which is the failure of the material in dynamic tension created by the intersection of two rarefaction waves, is found from each experiment. The elastic region under shock compression shows a steady, linear decrease in the residual spall strength with increasing compressive stress. The spall strength decreases from an average of 895 MPa at an impact velocity of 65 m/s (SC-5, SC-8) by 37% to 564 MPa at an impact velocity of 417 m/s (SC-9). This indicates inelastic microcracking even under the HEL. The microscopic modes of failure, as observed by the Scanning Electron Microscope, indicate increasing spall surface roughness and a transition from predominately intergranular fracture at 65 m/s impact speed (SC-8) to a more complex dual mode fracture at 356 m/s (SC-4). The two predominant regions on the spall surface in Experiment SC-4 are a brittle fracture area and a melted zone. In the higher impact velocity experiments, the onset of plastic deformation counteracts the decreasing trend in the spall strength, and the spall strength is nearly at the same level in the two elastic-plastic experiments. No fragments suitable for postimpact SEM analysis were recovered from these experiments. The addition of a pressure-shear loading causes a degradation of the spall strength of Si3N4. A twelve degree skew angle, producing a relatively moderate level of shear stress/strain in the Si3N4, causes the material spall strength to decrease five times as rapidly as in pure shock compression. This is due to additional damage caused by the shear. The spall strength drops by 69% of its level of 803 MPa for Experiment SC-11 at an impact velocity of 115 m/s, to 249 MPa at an impact speed of 233 m/s (SC-12). This decreased spall strength is not reflected in a change of damage mode. This is also observed in the SEM pictures of the material, where the same brittle damage mode that was seen at the 208 m/s (SC-2) impact speed under shock compression is present in the shear experiments conducted at an impact velocity of 233 m/s (Experiment SC-12). Under the combined pressure and shear loading at impact velocities of 233 m/s

310 (SC-12), silicon nitride is found to loose almost its spall strength. In fact in Experiment SC-13 at 299 m/s impact velocity the spall strength is essentially zero. The results of these experiments can be summarized as follows. The AS 8000 grade Si3N4 exhibits qualities that recommend it as a candidate for aircraft engine turbine blades. The HEL of 12 GPa is quite high, indicating that under impact conditions Si3N4 retains its elastic behavior. This is comparable to other silicon nitrides [Nahme et al., 1994; Mashimo, 1998] and to silicon carbide, another aerospace material [Bourne and Millet, 1997; Feng et al., 1998]. The spall strength of the material in the low speed elastic region is 895 MPa (SC-5, SC-8), which compares favorably with that of other silicon nitrides [Nahme et al., 1994]. The inelastic deformation that occurs above the HEL has the effect of reversing the decrease in spall strength. In contrast, the spall strength of silicon carbide disappears after the HEL [Bourne and Millet, 1997]. The spall strength for Si3N4 is ~ 249 MPa even under combined pressure-shear loading at an impact velocity of 233 m/s and at a 12 degrees skew angle (SC-12). Moreover, during dynamic shock compression of Si3N4, in all the experiments conducted in the present study, no failure waves were observed. This is in contrast to a dynamic shock compression of silicon carbide and glass under similar shock loading conditions, where failure wave front shave been consistently reported [Bourne and Millet, 1997]. These observations make AS800 silicon nitride more suitable for use in the environment of aircraft turbines than the other grades of silicon nitride and other materials such as silicon carbide.

311 C. The Effect of Dual Shock Loading and Shock Reverberation on the Strength of AS800 Grade Silicon Nitride Shock compressed materials that are still intact, but have deformed inelastically exhibit different dynamic strength than the same material that has not been shocked. Two types of multiple shock experiments were conducted in the present on silicon nitride in order to explore these effects. The first set of experiments involves the use of a two layer flyer where the second layer is of either higher (tungsten carbide) or lower (aluminum) impedance than the silicon nitride. This is modeled after the experiments performed on alumina [Reinhart and Chhabildas, 2003]. The second method involves creating reverberations in a thin target material, in this case tungsten carbide by impacting it with a thicker flyer of silicon nitride. This is modeled after experiments conducted on glass-fiberreinforced polyester [Dandekar et. al., 2003]. These experiments are detailed in Chapter VI. Experiments SR-1 and SU-1, using dual flyers, were unable to generate compressive stresses exceeding HEL in the first shock compressed Hugoniot State. This is due to the restriction of the maximum impact velocity of the single stage gas gun to around 500 m/s. Using both one dimensional stress wave theory and computations based on the methods detailed in Reinhart and Chhabildas [2003] the stress and the particle velocity states were predictable. Analysis of the stress and strain from the re-shock impact data of Experiment SR-1, indicates that the material follows the hydrodynamic Hugoniot curve during initial loading and during the re-shock. A basic computation of the HEL gives a value of 11.3 GPa, down by 6% from the value of 12 GPa for un-shocked specimens. The deformation during the re-shock followed the hydrodynamic Hugoniot curve, and the analysis method [Reinhart and Chhabildas, 2003] determined the residual shear strength to be 1.21 GPa. The shock unloading in Experiment SU-1 reveals that the unloading from the initial shocked state occurs elastically. An elastic unloading is consistent with theory, because this experiment occurs within the elastic range. Unlike the spall strength, these results do not indicate a decrease in longitudinal properties at high velocities.

312 Shock reverberation experiments were conducted at impact speeds around 500 m/s using a thin tungsten carbide target and a thick silicon nitride specimen. Because of the geometry, several unloading states were generated in the two materials. The free surface velocity records from the two reverberation experiments directly provided the target materials stress state and particle velocity. The target, which is tungsten carbide behaves elastic-linear plastically. By using the analysis method provided by Hall et al. [Preliminary], the stresses in the silicon nitride flyer were calculated. The silicon nitride remains elastic. The stresses and strains during each reverberation in both experiments, i.e. RB-1 and RB-2, follow the predictions established in Chapter V for the elastic Hugoniot curve at the stress levels in the reverberation experiments. As a result, the shocking and reverberations do not cause a measurable change in the dynamic response of AS800 grade silicon nitride at the impact velocities used in the present study. D. Shock Induced Failure Waves During Dynamic Compression of Soda Lime Glass The failure wave is a phenomena observed in glasses and ceramics with glassy phases. This failure wave causes the material to fail in such a way that the longitudinal properties such as impedance and wave speed remain largely unchanged. The shear and spall strength of the material, on the other hand are both adversely affected. The specific nature of this wave and its effects on glasses is discussed in several papers such as [Kanel et al., 2002]. In the present study, in order to understand the initiation of propagation of failure in soda lime glass, shock compression and pressure-shear experiments were conducted. These experiments are detailed in Chapter VII. Three types of experiments were conducted: spall strength experiments with and without the presence of shear, spall and shear strength measurement experiments, and impedance measurement experiments. One of the spall strength experiments (Al/G1) was conducted under pure compression. The remaining experiments were conducted under combined pressure and shear loading at a skew angle of 18

313 degrees. These experiments were designed to extend the findings of Clifton et al. [1998] to higher impact velocities. Aluminum flyers were used in the spall strength experiments because their low impedance enabled the creation of high tensile stresses. The spall strength of soda lime glass under shock compression was found to be greater than 3.89 GPa in Experiment Al/G1. This spall strength compares favorably with other investigators who found the spall strength to be in excess of 3 GPa [Kanel et al., 2002; Brar et al., 1991]. In Experiment Al/G2, with a pressure-shear loading at 18 degrees, the spall strength was found to be 3.49 GPa, which shows that the spall strength is degraded by the presence of shear. Thus, the soda lime glass behaves like ceramics such as silicon nitride where a skew angle of 12 degrees was observed to decrease the spall strength significantly. The second set of experiments used tungsten carbide flyers was to increase the compressive stress level. The failure wave effects of shear were determined by conducting experiments at two impact velocities, 220 m/s (WC/G1, WC/G2) and 410 m/s (WC/G3, WC/G4), and using two thicknesses of glass targets so that the spall plane would be either in front or behind the propagating failure front. The corresponding shock compression stress were 2.67 GPa and 4.98 GPa. Ahead of the failure wave, in both the low and high velocity shots, an elastic wave reflection occurs. Thus, in these experoments the glass remains intact because the spall strength is greater than the applied tensile stresses, which were 644 MPa (WC/G1) and 1.20 GPa (WC/G3). Experiment WC/G2 was designed such that the spall plane was formed behind the failure wave front. The impact velocity was 220 m/s. The measured free surface particle velocity profile showed a drop in particle velocity at the time corresponding to the arrival of the spall signal. However, like in experiment WC/G1, the particle velocity does not show the rapid pull-back consistent with the spall event and is consistent with a no-spall level prediction of the particle velocity. A small velocity rise of 13 m/s that was seen at the approximated

314 arrival time of the failure wave signal suggests that some damage occurred. However, this damage was not enough to effect the spall strength, which remained higher than the maximum tensile stress generated by the impact. Experiment WC/G2 is contrasted by Experiment WC/G4 at 410 m/s where no velocity drop is seen at the time of the spall signal. This lack of spall strength and the particle velocity rise of 45 m/s at the anticipated arrival time of the failure wave indicated that the glass had failed completely. From Experiments WC/G2 and WC/G4 an estimate for the failure wave velocity was made. This estimate is 1.7 0.1 km/s, with the slightly higher velocity occurring for experiment WC/G2. This finding is close to previous estimates for the failure wave velocity. The shear strength of the material has been observed to change with increasing impact velocity. This shear strength is reflected in the transverse particle velocity measured during impact. At the lower impact velocities, the level of the measured transverse velocity is near the predicted elastic level. Experiment WC/G1 shows the transverse velocity to be on average 104 m/s. This is 14% lower than the elastic value of 120 m/s. In Experiment WC/G2, the peak transverse velocity is 109 m/s. The elastic transverse velocity is still 120 m/s, so this value is 9.2% off. In the higher impact velocity experiments, the transverse velocity is much smaller a percentage of the elastic prediction. In Experiment WC/G3, in front of the failure wave a peak transverse velocity of 147 m/s is seen. This is 34% below the elastic value of 224 m/s. Experiment WC/G 4 exhibits a peak transverse velocity of 148 m/s. The elastic level is 222 m/s, so the actual value is 33% below it. This suggests that the initiation of the failure wave, which according to Feng [2000] occurs at the impact surface, may affect the level of the shear wave prior to its propagation through the material. Clifton et al. [1998] also found that the transverse velocity rose to a level at or just below the elastic prediction and then decreased. Cliftons experiments were at stress levels of 4.3 GPa, lower than the highest levels conducted in this study.

315 His results agree with the lower velocity shear results from the experiments in Chapter VII, but do not show the 30% drop in from the elastic velocities seen in the current studys the higher velocity experiments. Two experiments were conducted to measure the impedance of the damaged material. These experiments used tungsten carbide as targets and glass as the flyer plates, and were conducted at impact velocities of 250 m/s (G/WC1) and 454 m/s (G/WC2). The corresponding compressive stresses were 3.04 GPa and 5.51 GPa, respectively. The configuration created reverberations similar to those in the silicon nitride experiments RB-1 and RB-2, described in Chapter VI. In Chapter VII, the behavior of the tungsten carbide was analyzed. The normal material behavior is well predicted by elastic theory in the lower velocity case (G/WC1). The measured transverse velocity peaks at 10 m/s and then decreases to 5 m/s in about 0.6 microseconds. This is 45% lower than 18.4 m/s, the elastic prediction. Clifton et al. [1998] saw a near constant level of shear stress at his elastic prediction. In Experiment G/WC1, the longitudinal impedance of soda lime glass is measured to be 13.6 MPa/(m/s). This is within 7% of the elastically computed value of 14.5 MPa/(m/s). The shear impedance is 5.94 MPa/(m/s), which is 69% of the elastic shear impedance, 8.6 GPa. In the high speed experiment (G/WC2), the normal velocities are below the elastic prediction. For example, the first stressed state is 95 m/s, which is down 8.6% from the elastic value of 104 m/s. The transverse velocity is 82% lower than predicted, being only 5.9 m/s instead of 33.42 m/s. It decreases to zero after about 6 microseconds. This indicates that the material has failed and that the interaction of the shear waves is not predicted by elastic computations from the intact material parameters. Clifton et al.s [1998] experiments showed oscillatory transverse particle velocity profiles at the higher impact velocities, which do not match the near zero behavior seen in Experiment G/WC2. The normal impedance in this experiment is measured to decrease to 13.0 MPa/(m/s). This is only 10% lower than the elastic longitudinal impedance, suggesting that the assumption of constant longitudinal properties that several of the theories

316 depend upon is valid. The shear impedance is 4.24 MPa/(m/s) which is 49% of the elastic shear impedance. This is further evidence that the shear properties are effected by the failure wave, but that the longitudinal properties are nearly unchanged.

317 E. Discussion of Results This dissertation has explored the effects of impact on two brittle materials, AS800 grade silicon nitride and soda lime glass. In both cases, the material has been examined under planar shock compression. The spall strength of both materials was examined, as was the HEL of silicon nitride. Additional particle impact tests were carried out on soda lime glass. These tests examined the three dimensional nature of the impact process. Similar tests had already been carried out on AS800 grade silicon nitride by Choi et al. [2002]. The particle impact experiments and a numerical model for the internal stress propagation [Nathenson et al., 2005] resulted in the examination of the behavior of soda lime glass under three dimensional impact. The surface strains show an increase in tensile stress with decreasing thickness and a limit on the strain imposed by the finite failure strength of the material. The amount of energy consumed by cracking is larger than that dissipated by elastic processes. The material cracking patterns were also evaluated and radial cracking causes the greatest detriment to the integrity of the glass. In shock compression, the spall strength of intact soda lime glass decreases with the addition of shear stress. At high compressive stress levels, the region in the failure wave shows no spall strength and a reduced shear strength. Also, the normal impedance of the comminuted glass changes only 10% from the impedance of the intact glass, while the shear impedance changes by 51%. AS800 grade silicon nitride is a leading candidate for aircraft engine turbine blades due to its desirable high temperature properties and its strong resistance to damage from particle impact. In this work, AS800 grade silicon nitride was studied using both shock compression and pressure-shear at a skew angle of 12 degrees. In the shock compression experiments, the HEL of the material was found to be 12 GPa, with a corresponding dynamic yield strength of 7.6 GPa. The materials spall strength decreases linearly with increasing impact velocity, but to ceases decreasing upon the onset of plastic deformation. The addition of

318 shear caused the material to lose spall strength with increasing impact velocity at a rate five times as fast as under normal shock compression. Experiment SR-1 using dual shocks showed that the path between increasing stress states follows the hydrodynamic curve. Unloading after an initial shock proceeds elastically in Experiment SU-1, a fact which is confirmed by the unloading in reverberation Experiments RB-1 and RB-2. As a result of these experiments, the material is shown to have superior or comparable strength properties to other grades of silicon nitride, silicon carbide, and nickel based superalloys. These observations on the dynamic strength of AS800 grade silicon nitride make it a leading candidate for the next generation of turbine blades.

319 References
Bourne, N. and J. Millet 1997. Delayed Failure in Shocked Silicon Nitride. Journal of Applied Physics. 81 [9] 6019. Brandes E. A. and G. B. Brook ed. 1999. Smithells Metals Reference Book. Butterworth Heinemann. Brar, N. S., Z. Rosenberg, and S. J. Bless., 1991 B. Spall Strength and Failure Waves in Glass. Journal de Physique IV. Vol 1. C3. 639-644. Choi, S. R., J. M. Pereira, L. A. Janosik, and R. T. Bhatt., 2002. Foreign Object Damage Behavior of Two Gas-Turbine Grade Silicon Nitrides by Steel Ball Projectiles at Ambient Temperature. NASA/TM-2002-211821. Clifton, R. J., M. Mello, and N. S. Brar., 1998. Effect of Shear on Failure Waves in Soda Lime Glass. Shock Compression of Condensed Matter 1997. American Institute of Physics. 521-524. Dandekar, D. P., C. A. Hall, L. C. Chhabildas, W. D. Reinhart. 2003. Shock Response of a Glass Fiber-Reinforced Polymer Composite. Composite Structures. 61 [1-2] 51-59. Feng, R.; G. F. Frasier; and Y. M. Gupta. 1998. Material Strength and Inelastic Deformation of Silicon Carbide Under Shock Wave Compression. Journal of Applied Physics. 83 [1] 79. Hall, C. A., L. C. Chhabildas, and W. D. Reinhart., Preliminary Report. Shock Hugoniot and Release States in GRP Composite From 3 to 20 GPa. Sandia National Laboratories. Lin H. T. et al. 2001. Evaluation of Creep Property of AS800 Silicon Nitride from As-Processed Surface Regions. Ceramic Engineering and Science Proceedings. 22 [3] 175-182. Kanel, G. I., A. A. Bogatch, S. V. Razorenov, and Z. Chen., 2002. Transformation of Shock Compression Pluses in Glass Due to the Failure Wave Phenomena. Journal of Applied Physics. Vol. 92. No. 9. 5045-5052. Mitra, M. 1964. Disturbance Produced in an Elastic Half-Space by Impulsive Normal Pressure. Proceedings of the Cambridge Philosophical Society. 69: 683-696. Mashimo, T. 1998. In: High-pressure shock compression of solids III. L. Davison, M. Shahinpoor, editors New York : Springer 101-146.

320
Matweb. 2005. www.matweb.com Nahme, V. Hohler, A. Stilp. 1994. Determination of the Dynamic Material Properties of Shock Loaded Silicon-Nitride. American Institute of Physics. 765-768. Nathenson, D., G. Chen, and V. Prakash. 2005. Dynamic Response of Soda Lime Glass to Small Particle Impacts. Society for Experimental Mechanics Conference Proceedings. Reinhart, W. D.; L. C. Chhabildas. 2002. Strength Properties of Coors AD995 Alumina in the Shocked State. International Journal of Impact Engineering. 29 [1-10] 601-619.

321

Appendices
Appendix 1: Equations of Wave Propagation Resulting from a Point Load on a Elastic Half Space This discussion is taken from Lamb [1904]. In his paper, Lamb described the Rayleigh wave component of the radial surface displacement. The following manipulations are used to derive his result for the surface displacement. 1.1. Using the Equations of Motion to Generate the Radial and Vertical Surface Displacements The equations of motion for an isotropic elastic solid are: 2u = ( + ) + 2 u , 2 t x
2v = ( + ) + 2 v , 2 t y

{A.1.1}

{A.1.2}

and, 2w = ( + ) + 2 w . 2 t z

{A.1.3}

In Equations {A.1.1} through {A.1.3}, u v w + + . x y z

{A.1.4}

In Equations {A.1.1} through {A.1.4}, u, v, and, w, are the displacements in the x, y, and, z directions respectively; , is the density; and , and , are the Lame constants. For simple harmonic motion of the form eipt , where t is time, i is the square root of negative one, and p is the frequency, the solutions for the displacements in terms of a potential , can be written as:

322

u=

+u', x

{A.1.5}

v=

+ v', y

{A.1.6}

and,

w=

+ w' . z

{A.1.7}

Equations {A.1.5} through {A.1.7} explicitly state the homogenous component of the solution. The particular solutions are not explictly stated, but are identified as, u, v, and, w. These equations satisfy differential Equations {A.1.1} through {A.1.3} if the following conditions are met:

( ( ( (

+ k2 )u ' = 0 , + k2 )v' = 0, + k2 ) w' = 0 , + h2 ) = 0 ,

{A.1.8} {A.1.9} {A.1.10} {A.1.11}

and,

u ' v ' w ' + + = 0. x y z


In Equations {A.1.8} through {A.1.11},

{A.1.12}

323
p2 , + 2

h = pa 2 =

{A.1.13}

and,
p2

k = pb 2 =

{A.1.14}

The inverse of the wave velocities are defined for the longitudinal and shear waves, respectively, as follows:

a= 1

CL

{A.1.15}

and,

b= 1

CT

{A.1.16}

In the above equations, the longitudinal wave velocity is CL, and the shear wave velocity CS. In order to solve partial differential Equations {A.1.5} through {A.1.7}, particular solutions, u, v, and, w, must be specified. can then be written in terns of one variable, , as: These particular solutions

u'=

2 , xz
2 , yz

{A.1.17}

v' =

{A.1.18}

and,

324 2 + k 2 . 2 z

w' =

{A.1.19}

Equations {A.1.17} through {A.1.19} satisfy Equations {A.1.1} through {A.1.7} if the following condition is true

+ k2 ) = 0,

[VP1]

{A.1.20}

Substituting into Equations {A.1.5} through {A.1.7} yields:

u=

2 + , x xz

{A.1.21}

2 , v= + y yz

{A.1.22}

and,

w=

2 + 2 + k 2 . z z

{A.1.23}

Symmetry is then considered about the Z-axis. Defining the radial direction, ,

2 = x2 + y 2 ,
In terms of , the Laplacian can be written as

{A.1.24}

2 =

2 1 2 + + 2. 2 z

{A.1.25}

Generalizing the u and the v coordinates as q and w, and re-writing Equations {A.1.21} to {A.1.23}:

325

2 q= + , z w= 2 + + k 2 . z z 2

{A.1.27}

{A.1.28}

The particular solutions, , &, are now written explicitly:

= Ae z J o ( ) , = Be z J o ( ) .

{A.1.29} {A.1.30}

In Equations {A.1.29} and {A.1.30}, A, and, B, are constants, is a variable used in the integration, and[VP2] J0 is the Bessels function of the first kind. In the exponents of Equations {A.1.29} and {A.1.30}:

=
and,

h2 ) ,

{A.1.31}

k2 ) .

{A.1.32}

Plugging these back into Equations {A.1.27} and {A.1.28} for q & w, yields

q = Ae z J o ' ( ) +

Be z J o ' ( ) , z

{A.1.33}

q = Ae z J o ' ( ) + BJ o ' ( ) ( ) e z ,

{A.1.34} {A.1.35}

q = Ae z Be z J o ' ( ) ,

326
q = Ae z + Be z J1 ( ) . w = Ae z J o ( )( ) + B ( ) e z J o ( ) + k 2 Be z J o ( ) , z
2

{A.1.36}

{A.1.37}

w = A ( ) e z J o ( ) + B ( ) e z J o ( ) + k 2 Be z J o ( ) ,
w = Ae z J o ( ) + 2 Be z J o ( ) + k 2 Be z J o ( ) .

{A.1.38} {A.1.39}

In Equation {A.1.36} J1 is the Bessels function of order ??? Additionally,

2 + k2 = 2 .
Using Equation {A.1.40}, the following simplification can be made

{A.1.40}

w = ( Ae z + 2 Be z ) J o ( ) .

{A.1.41}

1.2. Calculating the Radial and Vertical Stresses on the Surface Starting from the differential Equation {A.1.42} for the radial stress z , and substituting from Section 1.1:

z =

q w , + z

{A.1.42}

Ae z 2 Be z J1 ( ) = , z z 2 Ae + Be J1 ( )

{A.1.43}

z = { Ae z 2 Be z + Ae z 3 Be z } J1 ( ) , z = 2 Ae z + 2 Be z 2 Be z J1 ( ) ,

{A.1.44} {A.1.45}

327

z = 2 Ae z + ( 2 2 ) Be z J1 ( ) ,

{A.1.46}

2 + 2 = ( 2 k 2 ) + 2 = 2 2 k 2 ,
z = 2 Ae z ( 2 2 k 2 ) Be z J1 ( ) .
But, z, is equal to 0 on the surface, so

{A.1.47}

{ {

{A.1.48}

z = 2 A B ( 2 2 k 2 ) J1 ( ) .

{A.1.49}

Next, calculating the vertical stress, zz , from the differential Equation {A.1.50}:

zz = + 2 , z zz = , + + 2 z z
w = 2 Ae z 2 Be z J o ( ) , z q w w

{A.1.50}

{A.1.51}

{A.1.52}

q w z z + = ( Ae + Be ) J1 ( ) . z

{A.1.53}

Using the properties of Bessels functions,


1 J o ( ) = J1 ( ) + J1 ( ) .

{A.1.54}
1 :[VP5]

When the radius from the impact point becomes large,


J o ( ) J1 ( ) ,

{A.1.55}

328

( Ae z + Be z ) J1 ( ) = ( Ae z + Be z ) J 0 ( ) ,
( Ae z + Be z ) zz = J o ( ) , + ( + 2 ) 2 Ae z 2 Be z

{A.1.56}

{A.1.57}

2 Ae z + ( + 2 ) 2 Ae z zz = + ( + 2 ) 2 Ae z + ( 2 Be z ) J o ( ) . 2 z + ( + 2 ) Be
Simplifying,

{A.1.58}

zz = { 2 Ae z + ( + 2 ) 2 Ae z ( 2 ) 2 Be z } J o ( ) .

{A.1.59}

Using Equations {A.1.13}, {A.1.14}, and {A.1.30}, the following relations can be derived:
h 2 ( + 2 ) = p 2 = k 2 .

{A.1.60}

Therefore, Equation {A.1.59} can be simplified further,

zz = 2 Ae z + ( + 2 ) ( 2 h 2 ) Ae z ( 2 ) 2 Be z J o ( ) ,

{ {

{A.1.61}

zz = 2 Ae z + ( ( + 2 ) 2 k 2 ) Ae z ( 2 ) 2 Be z J o ( ) , {A.1.62}
2 Ae z + 2 Ae z zz = + ( 2 2 k 2 ) Ae z J o ( ) , 2 z ( 2 ) Be

{A.1.63}

zz = ( 2 2 k 2 ) Ae z ( 2 ) 2 Be z J o ( ) ,

{A.1.64}

329

zz = ( 2 2 k 2 ) Ae z 2 2 Be z J o ( ) .
Setting, z, equal to 0 on the surface:

{A.1.65}

zz = ( 2 2 k 2 ) A 2 2 B J o ( ) .
1.3. Specific Solution for a Concentrated Vertical Pressure

{A.1.66}

Assuming a case of normal stress acting on the surface, i.e. z=0, the normal component of the stress can be written:

[ zz ]z =0 = P J o ( ) ,
and the radial stress:

{A.1.67}

[ z ]z =0 = 0 .

{A.1.68}

Then, the coefficients in Equations {A.1.29} and {A.1.31} can be found by Newtons second law. For the vertical stress, zz , this means:

zz = ( 2 2 k 2 ) A 2 2 B J o ( )

dy , dx

{A.1.69}

[ zz ]z =0 = P J o ( ) = {( 2 2 k 2 ) A 2 2 B} J o ( ) ,
P = ( 2 2 k 2 ) A 2 2 B , P = ( 2 2 k 2 ) A 2 2 B .

{A.1.70}

{A.1.71}

{A.1.72}

For the radial stress, z :

330

z = 2 A B ( 2 2 k 2 ) J1 ( ) ,

{A.1.73}

[ z ]z =0 = 0 = {2 A B ( 2 2 k 2 )} J1 ( ) ,
0 = 2 A B ( 2 2 k 2 ) J1 ( ) ,

{A.1.74}

{A.1.75}

0 = 2 A B ( 2 2 k 2 ) .
Combining Equations {A.1.72} and {A.1.76}:

{A.1.76}

A=

2 2 k 2 P , F ( )

{A.1.77}

and,
2 P . F ( )

B=

{A.1.78}

In Equations {A.1.77} and {A.1.78},

F ( ) = ( 2 2 k 2 ) 4 2 .
2

{A.1.79}

Thus, Equations {A.1.36} and {A.1.41} for the generalized coordinates q & w can be re-written as follows. For q:

q = Ae z + Be z J1 ( ) ,
2 2 k 2 P z 2 P z q = e + e J1 ( ) , F ( ) F ( )

{A.1.36}

{A.1.80}

331
and thus,

qz = 0 =

( 2 2 k 2 2 ) P J ( ) . 1 F ( )

{A.1.81}

For w:

w = ( Ae z + 2 Be z ) J o ( ) , 2 2 k 2 P z 2 P z w = e + 2 e J o ( ) , F ( ) F ( )
and thus,

{A.1.41}

{A.1.82}

wz =0 =

k 2 P J o ( ) . F ( )

{A.1.83}

Using a concentrated vertical pressure with an exponential loading function,

R eipt : P = R d / 2 .
{A.1.84}

In Equation {A.1.84}, R, represents the force. Integrating from zero to infinity gives the displacements due to the forcing function:

qo =

R 2

2 ( 2 2 k 2 2 )
F ( )

J1 ( ) d ,

{A.1.85}

wo =

R k 2 J 0 ( ) d . 2 F ( ) 0

{A.1.86}

Next, the Bessels functions are written in terms of hyperbolic cosines:

332

Jo =

(e
i
0

i cosh u

e i cosh u ) du ,

{A.1.87}

J1 =

(e
0

i cosh u

+ e i cosh u ) cosh udu .

{A.1.88}

Using Equations {A.1.87} and {A.1.88}, Formulas {A.1.85} and {A.1.86} for the surface displacements become:

R qo = 2 cosh udu 2 0

( 2

F ( )

k 2 2 )

ei cosh u d ,

{A.1.89}

wo =

iR 2 2

du
0

k 2 i cosh u e d . F ( )

{A.1.90}

Using the method of principle values, symbolized by, P :

2 ( 2 2 k 2 2 )
F ( )

2 H sin ( cosh u ) k ei cosh u d = 2 ( 2 2 k 2 ) i cosh u , {A.1.91} 2 e d +4k F ( ) f ( ) h

2 K sin ( cosh u ) k 2 i cosh u i cosh u P e d = 2k 2P e d . F ( ) F ( ) k 2 2 2 k 2k 2 ( 2 k ) e i cosh u d F ( ) f ( ) h

{A.1.92}

In Equations {A.1.91} and {A.1.92},

333 ( 2 2 k 2 21 1 ) F ( )

H=

{A.1.93}

k 21 , K= F ( ) F ( ) = ( 2 2 k 2 ) 4 2 ,
2

{A.1.94}

{A.1.95}

f ( ) = ( 2 2 k 2 ) + 4 2 ,
2

{A.1.96}

= kc

{A.1.97}

In Equation {A.1.97}, the inverse of the Rayleigh wave speed is denoted by, c. Further simplification yields:
2 2 2 k R ik 2 R ( 2 k ) HK1 ( ) + D1 ( )d , P qo = F ( ) f ( ) 2 h

{A.1.98}

i R KJ ( ) o 2 2 ik R Pwo = P D0 ( ) d . 2 k F ( ) 2 2 2 k 2 ik R ( 2 k ) D ( ) d 2 F ( ) f ( ) 0 h

{A.1.99}

In Equations {A.1.98} and {A.1.99}, K1 ( ) is a Bessels function of the second kind and,
D0 ( ) = K o ( ) iJ o ( ) .

{A.1.100}

334
Superimposing a system of free waves to the solutions in Equations {A.1.98} and {A.1.99}:

qo =

R HD1 ( ) , 2 i R y . KJ 0 ( ) 2 x

{A.1.101}

wo =

{A.1.102}

Or writing Equations {A.1.101} and {A.1.102} in a different way,


2 2 2 k R ik 2 R ( 2 k ) qo = HD1 ( ) + D1 ( )d , F ( ) f ( ) 2 h

{A.1.103}

2 2 k ik 2 R ik 2 R ( 2 k ) P wo = D0 ( ) d D0 ( ) d . 2 F ( ) 2 F ( ) f ( ) k h 2

{A.1.104}

Inserting the time factor, which is the exponential term, eipt , into Equations {A.1.103} and {A.1.104}:
2 2 2 k R ik 2 R ( 2 k ) ipt ipt qo = HD1 ( ) e + F ( ) f ( ) D1 ( )e d , h 2

{A.1.105}

ik 2 R P D0 ( ) eipt d 2 k F ( ) wo = . 2 2 2 k ik 2 R ( 2 k ) ipt 2 F ( ) f ( ) D0 ( ) e d h
Using only the most singular portion of the principle value integral,

{A.1.106}

335
k 2 k 2 e i d P D0 ( ) eipt d = 2i K cos ( cosh u ) + 2 F ( ) F ( ) 0
h k 2 ( h 2 + 2 )c d 1 1 2 i + k e d 2 2 2 2 2 h 0 ( 2 + k ) 4 F ( ) f ( ) ( h2 + 2 ) ( k 2 + 2 ) k

. {A.1.107}

This reduces equation {A.1.107} to:

k 2 P D0 ( ) eipt d = 2i K cos ( cosh u ) + H .O.T . F ( )

{A.1.108}

In Equation {A.1.108}, H.O.T., stands for higher order terms. specifying only the lower order terms

Therefore,

Pwo =

R KK 0 ( ) + H .O.T . 2

{A.1.109}

Simplifying Equations {A.1.105} and {A.1.106} to show only the most important terms:

qo =

R HD1 ( ) eipt + H .O.T . , 2

{A.1.110}

wo =

R KD0 ( ) eipt + H .O.T . 2

{A.1.111}

When the radial distance, , is very large, the asymptotic expansion of, Do ( ) , yields approximate exponential forms:

Do ( ) =

i ( + e 2

4)

12 12 32 + ... . 1 2 1!( 8i ) 2!( 8i )

{A.1.112}

Equations {A.1.110} and {A.1.111} are then further simplified,

336 qo i R H 2 e i( pt 4) ,

{A.1.113}

wo

R K 2

e i( pt

4)

{A.1.114}

1.4. Generalizing the Equations of Displacement for an Arbitrary Loading The Equations {A.1.110} and {A.1.111} can be written using the Bessels function:

Do ( ) =

e
0

i cosh u

du ,

{A.1.115}

as,

qo =

HR ip ( t c cosh u ) du + H .O.T . , e 0

{A.1.116}

and,

wo =

iKRc ip( t c cosh u ) e du + H .O.T . t 0

{A.1.117}

Equations {A.1.116} and {A.1.117} can be generalized by replacing the exponential inside the integral with an arbitrary forcing function, R(t).

H qo = R ( t c cosh u ) du + H .O.T . , 0 Kc R ( t c cosh u ) du + H .O.T . t 0

{A.1.118}

wo =

{A.1.119}

In Equation {A.1.119},

337

R ( t ) =

dp R ( ) sin p ( t ) d .
0

{A.1.120}

Using a specific forcing function, R(t):


R (t ) = R
2

t + 2
R

{A.1.121}

R ( t ) =

t . t + 2
2

{A.1.122}

Where t is the time; R is a constant force; and is related to t by:

t c = tan .

{A.1.123}

When the radius, , is much greater than a characteristic length defined as ,

/ c [VP6], and the variable, , is of moderate value Equations {A.1.118} and


{A.1.119} can be approximated for the general exponential loading, R eipt .

R ( t c cosh u ) du = 2
0

2 cos c

) )

cos ,

{A.1.124}

R ( t c cosh u ) du = 2
0

2 sin c

cos .

{A.1.125}

Ignoring the residual terms, yields the following approximations for the displacement:

qo = H

Rc 4
2

2 sin c

cos 2 , 2
3

{A.1.126}

338 Rc 4
2

wo = K

2 cos c

cos 2 . 2
3

{A.1.127}

These formulae describe two types of disturbances as shown in Figure A.1. Complete expressions for the disturbances of this loading are found by not ignoring or simplifying the higher order terms. A new variable related to the time t, and the radial distance , is defined in order to assist with this process:

{A.1.128}

This variable is then used in the solutions to Equations {A.1.118} and {A.1.119},

H R ( t c cosh u ) du 0 qo = , b 2 2 U ( ) R ( t cosh u ) du d b 0 a
1 P V ( ) R ( t cosh u ) du d . 2 b a t 0

{A.1.129}

wo =

{A.1.130}

In Equation {A.1.129},

U ( ) =

b3 ( 2 2 b 2 )

( 2 2 b2

( a ) ( b ) + 16 ( a )( b
2 2 4 4 2 2

2 ) 2 )

{A.1.131}

In Equation {A.1.130} when, a < < b :

V ( ) =

b3 ( 2 2 b 2 )

a2 )
{A.1.132}

( 2

b 2 ) + 16 4 ( 2 a 2 )( b 2 2 ) .
4

When, > b :

339

V ( ) =

b3

( 2

2 2

+ 4

( a ) ( a ) (
2 2 2 2

{A.1.133}

Equations {A.1.129} and {A.1.130} are extremely difficult to integrate, thus the approximations used for the general loading (Section 1.3), as well as the approximation in Equations {A.1.134} and {A.1.135} below are used.

R ( t cosh u ) du sin 0

32 cos 2 ,
3

{A.1.134}

R ( t cosh u ) du sin t 0

32 cos 2 ,
3

{A.1.135}

Using these approximations, Equations {A.1.129} and {A.1.130} become:

3 Rc 2 sin cos 2 H 4 2 2 4 c , qo 3 2 2 U ( ) sin 32 cos 2 d 4 b 0

{A.1.136}

1 sin wo 2 P V ( ) b a

3 32 cos 2 d .

{A.1.137}

The numerically computed displacement corresponding to the Rayleigh wave obtained by using Equations {A.1.136} and {A.1.137}, and estimates for the location and type of the longitudinal and shear waves using the methods suggested in Lambs [1904] paper, are shown in Figure A.2.

340
Appendix 2: Calculations for the Surface Wave Produced by the Loading of a Circular Area of a Half Space by an Impulse Load. 2.1. Solutions to the Equations of Motion Transformed using Laplace Transforms

The calculations in Appendix 2 use Cagniards Method as described by Mitra [1964]. The coordinate system is cylindrical and uses r and z as spatial dimensions and t as time. The half-space of the material is defined to be elastic and isotropic and is located in the half of the coordinate system where, z 0 . The surface traction for the impulsive loading of a circular area of radius a, is given as:

zz =

0 r a P ( t ) for , 0 a r

{A.2.1}

rz = 0 .

{A.2.2}

In Equations {A.2.1} and {A.2.2}, ( t ) is the Dirac delta function; zz is the vertical stress; and rz is the radial stress. Because of symmetry, there is no stress in the hoop direction. Other important conditions are that the half-space is at rest before the impulse and that the displacement is bounded as z . The equations of motion for a cylindrical coordinate system with no motion in the hoop direction are given as:
2u 1 u u 2u 2u 2w , = ( + 2 ) 2 + 2 + 2 + ( + ) r r r t 2 z r z r

{A.2.3}

2u 1 u 2 w 1 w 2w 2w 2 = ( + ) + + 2 + + ( + 2 ) 2 . r r t z r z r z r

{A.2.4}

Here, u, and, w, are the displacement components along the, r, and, z, axes respectively; , and, , are the lame constants; , is the density and; , and, ,

341
are the compression and shear wave velocities for the half-space respectively. These last two variables are related to the Lame constants. The definition of the wave velocities are:

=
and,

+ 2 ,

{A.2.5}

{A.2.6}

The solution to Equations {A.2.3} and {A.2.4} involves integral transforms and also Cagniards method. First, the Laplace transform of a function can be made for any, p, which is positive and real. The transform is defined:

f1 ( r , z , p ) = L f ( r , z , t ) = e pt f ( r , z , t ) dt .
0

{A.2.7}

The Laplace transforms of the equations of motion, with the initial conditions inserted are:
2u 1 u1 u1 2u 2 w1 , p 2u1 = 2 21 + 2 + 2 21 + ( 2 2 ) r r r z r z r
2 2 2u1 1 u1 1 w1 2 w1 2 w1 . p w1 = ( ) + + 2 + + r r z 2 r z r z r 2 2 2

{A.2.8}

{A.2.9}

The Henkel Transforms are defined:

u2 ( , z , p ) = rJ1 ( r ) u1 ( r , z , p ) dr ,
0

{A.2.10}

and,

342

w2 ( , z , p ) = rJ 0 ( r ) w1 ( r , z , p ) dr .
0

{A.2.11}

In Equations {A.2.10} and {A.2.11}, J1, and, J0, are Bessels functions. Multiplying the transformed Equations {A.2.8} and {A.2.9} by rJ1 ( r ) and
rJ 0 ( r ) , respectively, and integrating with respect to r, over an interval of zero

to gives:

2u1 1 u1 u1 p u1rJ1 ( r )dr = r 2 + r r r 2 rJ1 ( r )dr 0 0


2 2

2u 2 w1 rJ1 ( r )dr + 2 21 rJ1 ( r ) dr + ( 2 2 ) z r z 0 0

{A.2.12}

2u1 1 u1 p 2 w1rJ 0 ( r ) dr = ( 2 2 ) + rJ 0 ( r ) dr r z r z 0 0 . 2 w1 1 w1 2 w1 2 rJ 0 ( r )dr + 2 2 + rJ 0 ( r ) dr + r z 2 r r 0 0

{A.2.13}

Simplifying, Equations {A.2.12} and {A.2.13} take the form:

0= 2

2 u2 w 2 2 p u2 ( 2 2 ) 2 , 2 z z 2 w2 u 2s2 w2 + ( 2 2 ) 2 , 2 z z

{A.2.14}

0 =2

{A.2.15}

The definitions of p and s are:

p = +
2

p2

2
p2

{A.2.16}

s = 2 +

{A.2.17}

343
The variables p and s have real and positive components when the variable is real. The Equations {A.2.14} and {A.2.15} are subject to conditions for an infinite half space as r . Specifically, these equations should have a bounded solution as z + , i.e.
u2 = Ae
p z

+ Be s z ,

{A.2.18} {A.2.19}

w2 =

p z Ae + Be z . s
p s

In Equations {A.2.18} and {A.2.19} A and B are constants. The stresses from Equations {A.2.1} and {A.2.2} are defined using the well known stress-strain relationship:

zz = + + 2 , z r z
and,

{A.2.20}

zr = + . z r
Using the inverse of the Henkel transforms, these stresses become:

{A.2.21}

zz1 = J 0 ( r ) ( 2 2 2 ) u2 + 2
0

w2 d , z

{A.2.22}

and,

rz1 = J1 ( r )
0

u2 w2 d . z

{A.2.23}

The boundary conditions for Equations {A.2.22} and {A.2.23} are:

344

zz1 = Pa J 0 ( r )
0

J1 ( a )

d ,

{A.2.24}

and,

rz1 = 0 .
These boundary conditions are satisfied when,
p2 2 A ps + 2 + 2 2 B = 0 ,

{A.2.25}

{A.2.26}

and,
p2 Pa 2 2 J ( a ) . 2 + 2 A + 2 B = 1

{A.2.27}

In accordance to the Equations {A.2.26} and {A.2.27}, the constants A and B can be defined as:
PaJ1 ( a )

A=

+ 2 2

p2 2 2 + 2
4 ps

{A.2.28}

B=

PaJ1 ( a )

p22 + 2 2 ps 4
2

{A.2.29}

Using these values in Equations {A.2.18} and {A.2.19} for u2 and w2 , yields:

u2 =

PaJ1 ( a )

+ 2 2

4 p2 p s 2 2 + 2

p z

PaJ1 ( a )

p2 2 2 + 2

ps

e s z ,

{A.2.30}

345

w2 =

PaJ1 ( a )

p2

+ 2

p2 2 2 + 2
4 ps

p z

PaJ1 ( a )

2 + 2 2 ps 4
p2 2

e s z .

{A.2.31}

Using the inverse of the Henkel transforms on Equations {A.2.30} and {A.2.31} yields:

u1 ( r , z, p )
Pa
and,

=
0

J 1 ( r ) J 1 ( a )

p 2 2 + 2 e
2

p z

+ 2 ps e s z d ,

{A.2.32}

w1 ( r , z, p )
Pa

=
0

J 0 ( r ) J 1 ( a ) z p2 2 2 + 2 p e p + 2 p 2 e s z d .

{A.2.33}

In Equations {A.2.32} and {A.2.33},

p2

+ 2 2

4 2 ps .

{A.2.34}

In the analysis, since the desired goal is the surface displacement away from the impact region, it is only necessary to consider the case when the radius is greater than the contact area, r > a . For this condition, Watsons result as referenced by Mitra [1964] is as follows:

J1 ( r ) J1 ( a ) =

J ( R ) cos d ,
0 0

{A.2.35}

J 0 ( r ) J 1 ( a ) =

a r cos J ( R ) R d .
1 0

{A.2.36}

Where R is a function of the variable of integration ,

346
R = r 2 + a 2 2ar cos .
{A.2.37}

In accordance to Equations {A.2.35} through {A.2.37}, the displacements can be re-written as:

u1 ( r , z, p )
Pa
and,

= I1 cos d ,
0

{A.2.38}

w1 ( r , z, p )
Pa

= I2
0

a r cos d . R

{A.2.39}

In Equations {A.2.38} and {A.2.39},

I1 =
0

J 0 ( R )

p 2 2 + 2 e
2

p z

+ 2 ps e s z d ,

{A.2.40}

and,

I2 =
0

J 1 ( R )

p 2 2 + 2 p e
2

p z

+ 2 p 2 e s z d .

{A.2.41}

2.2. Cagniards Method for Obtaining the Inverse Laplace Transforms of, I1, and, I2 Mitra [1964] demonstrates Cagniards method using a typical term from Equations {A.2.40} and {A.2.41}. This term is known as, I, and is defined:
J 1 ( R ) J 0 ( R ) 1 p z z p2 p2 2 2 + 2 p e d = 2 2 + 2 p e p d .{A.2.42} I = R 0 0

The method for transforming Equation {A.2.42} uses the following substitutions:

347

= cos ( + ) , = sin ( + ) ,
X = R cos ,

{A.2.43} {A.2.44} {A.2.45}

and,
Y = R sin .

{A.2.46}

Simplifying Equation {A.2.42} using these substitutions and the integral representation of J 0 ( R ) :

I=

e i X + Y 1 p z p2 2 2 2 2 + 2 p e d d . R

{A.2.47}

The next step is substituting the following trigonometric functions into Equation {A.2.47}:

= cos + q sin ,
and,

{A.2.48}

= cos q sin ,

{A.2.49}

Substituting Equations {A.2.48} and {A.2.49} into Equation {A.2.47} and using a formulation developed by deHoop into {A.2.47} yields:

I=

e i R 1 p z p2 2 2 2 2 + 2 p e d dq . R

{A.2.50}

348
Removing the variable, 2 is the next step. 2 = 2 + q 2 . Equations {A.2.50} and {A.2.51} give: {A.2.51}

e i R z p + q = 2 2 2 + p e d dq , 00
2 2

{A.2.52}

Also, the variable q is replaced by qp, and w by wp:

I=

Jdq .
0

{A.2.53}

J = Im F ( , q )d
0

{A.2.54}

In Equation {A.2.54}, Im represents the imaginary part of the integral, and

F ( , q ) =

+ 2 + q2

)(

( 2 + q 2 )

+ 2 2 + 2q 2 exp p i R + z

+ 2 2 + 2q 2

4 ( 2 + q 2 )

+ 2 + q2

) (

+ 2 + q2
2
1

+ 2 + q2 {A.2.55}

Equation {A.2.55} has branch points at,

= i q 2 + (1 2 ) , i q 2 + (1 2 ) ,
and poles at,

{A.2.56}

= iq, i q 2 + (1 V 2 ) .

{A.2.57}

349
The Rayleigh wave velocity is V; it is smaller than the shear wave velocity , and the longitudinal wave velocity . The integral J, is then evaluated by contour integration, which has the result:
2 z ( t qR ) + F1 w1 ( t , q ) , q H t J = L 2

(z

R2 )

(q

+ 12 . {A.2.58}

In Equation {A.2.58} the , H,

[VP7]

followed by the statement in brackets

represents the Henkel transformation of that statement,

F1 [ , q ] = Im ( 2 + q 2 )

1
2

+ 2 2 + 2q 2

) (

1
2

+ 2 + q2

+ 2 2 + 2q 2

4 ( 2 + q 2 )

) t

+ 2 + q2

) (

. 2 2 + + q

{A.2.59} Then, the representative term I, becomes,


2 H t z ) 2 ( I = L + R
t 2

( R2 + z2 )(1 2 )

F1 w1 ( t , q ) , q dq .

{A.2.60}

The contribution of the first term in Equation {A.2.60} to the vertical displacement is proportional to
z H (t ) 0

a r cos d . R2

{A.2.61}

The inverse Laplace transforms for the remaining terms in Equations {A.2.40} and {A.2.41} are computed in a similar manner. From these two equations the following expressions are generated:

350
t 2 R 2 + z 2 1 2 ) ( ) ( R2 + z 2 H t F 1 w1 ( t , q ) , q dq 0 2 2 2 w ( r , z, t ) a r cos H t R + z d , = R 2 Pa qj 0 + F 2 w1 ( t , q ) , q dq R z 2 2 0 + (1 ) H t

{A.2.62} and,
t 2 z 2 + R 2 1 2 ) ( ) ( R2 + z 2 H t F 3 w1 ( t , q ) , q dq 0 2 2 2 u ( r , z, t ) R +z d = cos H t 2 Pa qj 0 + F 4 w1 ( t , q ) , q dq R z 2 2 0 + (1 ) H t

{A.2.63} In Equations {A.2.62} and {A.2.63}:

R> 0 = for 1 R<

2 2

, z 2 2

{A.2.64}

q j = q1 if = 1 or t > ( R 2 + z 2 )

2 ) z , 2 ) z .

{A.2.65}

q j = q2 if = 0 and t < ( R 2 + z 2 )

{A.2.66}

The functions F2, F3, & F4 are:

351 F2 [ , q ] = Im

+ 2 + q2

1
2

+ 2 2 + 2q 2

4 ( 2 + q 2 )

) t

+ 2 + q2

) (

, + 2 + q2 {A.2.67}

F3 [ , q ] = Re

2 2 + 2q 2 + 12

1
2

+ 2 2 + 2q 2

4 ( 2 + q 2 )

) t

+ 2 + q2

) (

, 2 2 + + q {A.2.68}

and,

F4 [ , q ] = Re

+ 2 2 + 2q 2

) ( ) 4 ( + q ) ( (
2
1

+ 2 + q2
2

2
1

+ 2 + q2 + 2

) t +q ) (
2

. 2 2 + + q {A.2.69}

2.3. Surface Displacements From Equations {A.2.62} and {A.2.63}, the surface displacements can be found by taking the limiting values of u and w as z 0 + . However, this process is complicated because the Rayleigh pole = i q 2 + (1 V 2 ) , lies on the real axis. The integrands of these two equations have poles at,

q = ( t 2 R 2 ) (1 V 2 ) ,
for t > R / V . As z 0 + :
t2 1 it q j = q2 2 2 , and w j . R R

{A.2.70}

{A.2.71}

352
The Rayleigh pole is located at t:

1 1 1 1 1 1 t = R q 2 + 2 iz 2 2 , R q 2 + 2 iz 2 2 . V V V V

{A.2.72}

Then, when t is real and z is positive and small, q is near the Rayleigh pole. Thus, the integral on the q-axis must bypass the pole when Im [ q ] < 0 :

q = ( t 2 R 2 ) (1 V 2 ) .
Using the sign convention:
R when, t > , q <

{A.2.73}

t2 R2

12 ,

+ w2 + q 2 i

t2 R2

12 q 2

{A.2.74}

R when, t > , q <

t2 R2

12

)
12 q 2

+ w2 + q 2 i

t2 R2

),

{A.2.75}

and,
when, or

R R R < t < ,0 < q

< t,

t2 R2

12 < q <

(
2

t2 R2

12

)
{A.2.76}.

+ w2 + q 2

t R2 + q 2 ,

353
Equations {A.2.62} and {A.2.63} for u and v can now be written in terms of real values:
t 2 R 2 1 2 R R H t H t F [t , q ] dq 0 2 u ( r , z , t ) cos t 2 R 2 1 2 = H ( t ( R V ) ) d . {A.2.77} 2 Pa R + H t R 0 F [t , q ] dq + 2 2 2 t R2 1 t 1 2 2 R V

and,
t 2 R 2 1 2 R R H t H t G [t , q ] dq 0 t 2 R 2 1 2 t a r cos ) R 2 ( +H t w ( r , 0, t ) 3 2 G [t , q ] dq d .{A.2.78} = R t 2 R2 1 2 Pa 0 2 2 2 t R 1 +P G [t , q ] dq 0

In Equations {A.2.77} and {A.2.78}:


2

F (t, q ) =

(
2

t2 R2

12 q 2
4

2t2 + 2q 2 R

) ( ) + 16 (

t R2 + q 2
2

) {2 ( q

t R2 + 12
2

} )

t2 R2

q2

)(
2

t2 R2

12 q 2

)(

t2 R2 + q 2

, {A.2.79}

G (t, q ) =

(q

t2 R2

)(

(
2
1

2t2 + 2q 2 R
2

2t 2 R2

+ 2q

) (
t2 R2

t2 R2

12 q 2
2

)
1 2

+ 16

)(
2

t2 R2

2 q

)(

t2 R2

, +q {A.2.80}
2

G (t, q ) =

t2 R2

12 q 2
t2 R2

)
t2 R2

(q

t2 R2

)(

2t 2 R2

+ 2q

) (
2

2 q
1

) (

t2 R2

+q
2

354
{A.2.81} and,

V2

4 4 3V 2 3V 2 2V 4 2 2 2 2

) (1 ) (1 ) . 2 ( 2 ) (1 ) (1 )
V2 V2

{A.2.82}

V2

V2

V2

These terms contain the longitudinal, shear, and Rayleigh wave contributions. The F terms give the displacements due to the longitudinal wave. The second term in the equation for w (Equation {A.2.78}), and the second and third terms in the equation for u (Equation {A.2.77}) give the contributions to the shear and longitudinal waves up to the arrival of the first Rayleigh wave. The term in {A.2.77} gives the Rayleigh wave contribution to u, while the G term in {A.2.78} gives the Rayleigh wave contribution to w. contribution in {A.2.77} can be written as, uo: The Rayleigh wave

uo ( r , 0, t )
2Pa

cos H t ( R V ) d . t2 1 R 0 2 2
R V

{A.2.83}

355
This Rayleigh wave contribution can be written in a more explicit form as
0 uo ( r , 0, t ) V = 2Pa ar V ar 2E ( k ) K ( k ) . {A.2.84} r 2 + a 2 V 2t 2 2 1 1 V 2t 2 ( r a ) E ( k ) K ( k ) V 2t 2 ( r a )2

when,
0 < t < ( r - a ) V ( r - a ) V < t < ( r + a ) V . ( r + a ) V < t

In Equation {A.2.84},

V 2t 2 ( r a )2 k= , 4ar
E ( ) =
2

{A.2.85}

1 2 sin 2 xdx ,

{A.2.86}

and,
2

K ( ) =

dx 1 2 sin 2 x

{A.2.87}

In the special case where = , Equations {A.2.77} and {A.2.78} for u and w, can be further simplified as:

356

P U ( ) cos u ( r , 0, t ) = u0 ( r , 0, t ) + 2 d , 0 R

{A.2.88}

w ( r , 0, t ) =

r cos a W ( ) d . R2

{A.2.89}

In Equations {A.2.88} and {A.2.89}: for,

0< <1 3 1 3 < < 1 : 1 <


0 6 K ( l ) 18 ( 8l 2 , l ) + 6 4 3 20 12 3 l 2 , l U ( ) = 1 3 8 2 , {A.2.90} + 6 + 4 3 20 + 12 3 l 2 , l 1 1 1 1 3K ( l ) 9 ( 8, l ) 2 3 3 20 12 3 , l 4 2 1 3 1 + 2 3 + 3 20 + 12 3 , l

) (

) (

) (

) (

and for,

0< <1 3 1 3 < < 1 : 1 1< < 2 3+ 3 1 2 3+ 3 <

357 0 W ( ) = 2 3 3 4 .

+ 43

21 4

12

2 3 3 3 3 5

2 34 3

+ 12

2 3 +3 3 3 +5

+ 6 3
3

2 3 +3 3 3 +5

{A.2.91}

3 3 2 4

In Equations {A.2.90} and {A.2.91},


3 2 1 l= , 2

{A.2.92}

t
R

{A.2.93}

2 3+ 3

{A.2.94}

and,
2

( n, k ) =

(1 + n sin )
2 0

d 1 k 2 sin 2

{A.2.95}

Equation {A.2.88} gives Mitras [1964] solution for the longitudinal, shear and Rayleigh waves contributions to the radial displacement. The numerical solution of this equation for the radial displacements is plotted in Figure A.3.

358
Appendix 3: Calculation of Energy Absorbed by Elastic Stress Waves During Impact In Chapter III, the impact energy that is imparted to the specimen are compared with the predictions obtained from the simulations using LS-DYNA and from an analytical estimate for the elastic energy. This analytical solution is designed to estimate the fraction of the impact energy that is carried away as stress waves. The derivation of this energy estimate is taken from calculations by Hunter [1957]. The rate of work W, done on the specimen is given by:

dW du = a2 P (t ) . dt dt

{A.3.1}

In Equation {A.3.1}, a is the contact radius; P(t) is the pressure on the surface; t is time; and, u is the Fourier synthesis of the mean surface displacement. Integration over time yields:

W=

dW du 2 dt dt = a P ( t ) dt dt .

{A.3.2}

Miller and Pursey [1954] provide an equation of the normal surface displacement

u in a semi-infinite solid, for a uniform pressure Pei t ( > 0 ) , over the contact
area 0 r a as:

2 u ( r , t , z = 0 ) = Peit k2 0

k12 ) J12 ( a ) d
1 2

C44 F ( )

{A.3.3}

The radial dimension is represented by r, and the vertical by z, and is the variable of integration. In Equation {A.3.3},
k1 = / ( C11 / ) 2 ,
1

{A.3.4}

359
k2 = / ( C44 / ) 2 ,
1

{A.3.5}

and,
2 F ( ) = ( 2 2 k22 ) 4 2 ( 2 k12 )( 2 k22 ) .

{A.3.6}

In Equations {A.3.3} through {A.3.6},C11 and C44 are the components of the stiffness matrix, and is the density of the specimen. The mean displacement within the contact radius a, is:

u=

2 urdr = Pei t k2 2 a 0 0

k12 ) J12 ( a ) d
1 2

C44 F ( )

{A.3.7}

This can also be re-written as,

2 2 a 2 Pei t u= C44

k1 ( 2 1) J12 ( k1a ) d
1 2

k12 a 2 Fo ( )

{A.3.8}

In Equation {A.3.8},
2 Fo ( ) = ( 2 2 2 ) 4 2 ( 2 1)( 2 2 ) ,

{A.3.9}

and,

= k2 k = ( C11 / C44 ) = {2 (1 ) / (1 2 )} .
1 2 1 2

{A.3.10}

In Equation {A.3.10}, is the Poissons ratio of the specimen. The integral in Equation {A.3.8} can be re-written for > 0 , as:

360 2 2 a 2 Peit ( A ( ) + iB ( ) ) . C44

u=

{A.3.11}

In Equation {A.3.11}, the variable , represents frequency, and i represents the square root of negative one. Also,

A + iB = k1
0

1) J12 ( k1a ) d
1 2

k12 a 2 Fo ( )

{A.3.12}

For < 0 ,

u=

2 2 a 2 Peit ( A ( ) + iB ( ) ) . C44

{A.3.13}

Where, A, and, B, represent the real and imaginary parts of the integral. This gives real values of displacement u, for pressures P cos (t ) , and P sin (t ) . Next, an arbitrary pulse shape, P(t), is considered:

P (t ) =

P ( ) e

i t

d .

{A.3.14}

The superposition principle is used to determine the mean normal displacement,

u , for this load:


P ( ) eit A ( ) + iB ( ) d 2 2 2 a 0 u= 0 . C44 it + P ( ) e A ( ) + iB ( ) d
Simplifying,

{A.3.15}

361
P ( ) eit ( A ( ) + iB ( ) ) 2 2 a 2 u= d + P ( ) eit ( A ( ) iB ( ) ) . C44 0

{A.3.16}

Substituting Equation {A.3.16} into the differential Equation {A.3.1}:

P ( ) ei ( '+ )t ( A + iB ) dW 2 2 a 4i = d d ' P ( ') P ( ) ei( ' )t ( A iB ) . dt C44 0

{A.3.17}

Integrating with respect to time, t:

e e

i ( ' + )t

dt = 2 ( '+ ) ,

{A.3.18}

i ( ' )t

dt = 2 ( ' ) .

{A.3.19}

The Dirac delta function is . Therefore, Equation {A.3.17} becomes:

4 2 2 a 4 W= i d P ( ) P ( ) ( A + iB ) ( A iB ) . C44 0

{A.3.20}

Using the fact that the function P(), is equal to its complex conjugate, P*():
P ( ) = P ( ) ,

{A.3.21}

Equation {A.3.21} can be written as:


2 8 2 2 a 4 W= B ( ) P ( ) d . C44 0

{A.3.22}

362
In simplifying, Equation {A.3.22} the frequencies, , are restricted to satisfy, the condition

a / ( C11 / )

1 2

1.

[VP8]

{A.3.23}

A new quantity is defined as being equal to the quantity on the left hand side of Equation {A.3.23}
k1a a / ( C11 / ) 2
1

{A.3.24}

For these frequencies:

A + iB =

a
4 ( C11 / ) 2

Fo ( )

1) 2 d
1

{A.3.25}

B ( ) =

4 ( C11 / ) 2
1

{A.3.26}

and,
1 2 1 2 d ( ) = ( 2 1) 2 d . = Im 0 Fo ( ) 2 2 2 2 4 2 2 1 2 2 1 2 ) )( ) 0( (
1

{A.3.27}

The notation, Im, represents the imaginary part of the integral. Beta, , in Equation {A.3.27} was estimated to be a function of Poissons ratio such that when = 1/ 4 , 0.5374 , and when = 1/ 3, 0.415 . Using the integral equation for W, i.e. {A.3.22}, and combining it with the equation for B, i.e. {A.3.26}, yields :

W =

2 2 2 a 4

1 2

C44 ( C11 / )

2 P ( ) d . 2 0

{A.3.28}

363

In Equation {A.3.28},

4 (1 + )

2
C44

{A.3.29}

Simplifying Equation {A.3.28} using:

f ( ) = ( 2 )

f (t ) e

i t

dt = a 2 P ( ) .

{A.3.30}

yields,
8 (1 + ) 1 2 2 2 2 W= f ( ) d . 3 Co 1 2 0
1

{A.3.31}

In Equation {A.3.31},
Co = ( E / ) 2 ,
1

{A.3.32}

In Equation {A.3.32}, the modulus of elasticity of the specimen is, E. Using the specific pulse shape for an elastic collision, which is defined:
t 2o f ( t ) = M o cos (ot ) = . t 2 f ( t ) = 0 o

{A.3.33}

In Equation {A.3.33}, Mo, is the pulse magnitude, and o the pulse frequency. Using Equation {A.3.33}, Equation {A.3.30} is redefined as:

364

M f ( ) = o 2

/ 2 o
/2

cos ( t ) e
o
o

i t

dt =

M o o cos ( / 2 o ) . o2 2

{A.3.34}

Plugging this into the Equation {A.3.31} for W:


1 8 (1 + ) 1 2 2 M o o W= 3 Co 1 2 2
1

z 2 cos 2 ( z / 2 )

(z

1)

dz .

{A.3.35}

Using residue calculus, the integral in Equation {A.3.35} is the real part of the integral I:
2 i z 1 z ( e + 1) I= dz . 4 C ( z 2 1)2

{A.3.36}

The integral in Equation {A.3.36} has simple poles at,

Z = 1 .
These poles have residues: Res ( 1) = i / 4 . Cauchys integral theorem states:

{A.3.37}

{A.3.38}

z 2 cos 2 ( z / 2 )

(z

1)

dz =

i i

i 2 . = 4 4 4 8

{A.3.39}

Therefore, Equation {A.3.35} for the energy produced by the impact, W, is given by:

365

(1 + ) 1 2 2 W= M . 3 C0 1 2 o o
An equation for the contact indentation distance, Z, is:
Z = Z o sin (1.068Vo t Z o ) .

{A.3.40}

{A.3.41}

The maximum indentation, Z0, can be determined as


15 m m1 5 1 4 Z0 = g R 5V0 5 . 16 m + m1
2

{A.3.42}

In Equation {A.3.42}, m1 is the projectile mass; R is the radius of the projectile; andV0 is the impact velocity. The constant g is defined as:
2 1 2 1 1 . + E E1

g=

{A.3.43}

The modulus of elasticity of the projectile is E1, and the Poissons ratio of the projectile is 1. The impact force is approximated as one half period of a sine function, and is related to the acceleration in the vertical direction Z ,
f ( t ) = m1Z .

{A.3.44}

For specific time during which the force acts, and the forcing function in Equation {A.3.44} can be written as:
0 t o f ( t ) = m1Z oo2 sin (ot ) = . t 0, t f ( t ) = 0 o

{A.3.45}

In Equation {A.3.45}:

366

o = 1.068 (Vo Z o ) .

{A.3.46}

Offsetting the zero time by / 2 o , such that the peak is at time zero, the Equation {A.3.45} can be written as:
t 2 2o f ( t ) = m1Z oo sin (o t ) = . t 2 f ( t ) = 0 o

{A.3.47}

If the magnitude of the pulse is specified as,


2 M o = m1Z o o ,

{A.3.48}

then the elastic energy in Equation {A.3.40} can be written,

(1.068) (1 + ) 1 2 W=
5

1 2

3 o

3 5 m1Z o Vo . 1 2

{A.3.49}

In the limit where m , Equation {A.3.42} becomes,


15 Z o = {16 m1 g} 5 R 5Vo4 / 5 .
2
1

{A.3.50}

Substituting Equation {A.3.50} for, Zo, and taking the limit gives,

m R g Vo W= 1 . 3 Co
4 5 3 5

13

{A.3.51}

In Equations {A.3.50} and {A.3.51},


1 2 16 5 = (1.068 ) (1 + ) . 1 2 15
5 1 6

{A.3.52}

367

Normalizing by the initial kinetic energy, KEi , which is

KEi =

1 m1Vo 2

{A.3.53}

yields,

2 m1 5 R 5 g 5Vo 5 = . 3 1 Co m1Vo2 2
W
1 3 6 3

{A.3.54}

If the impacting mass is a sphere, then the mass can be defined in terms of the radius R, and the density of the sphere 1 , as
m1 = ( 4 3) R 3 1 .

{A.3.55}

Substituting Equation {A.3.55} into Equation {A.3.54}, yields:

1 m1Vo2 2

2 ( 4 / 3)

1 5 g 5Vo 5

3 Co

{A.3.56}

This equation for the normalized impact energy is used in the energy analysis in Chapter III.

368
References
Hunter, S. G. 1957. Energy Absorbed by Elastic Waves During Impact. Journal of the Mechanics and Physics of Solids. 5 162-171. Lamb, H., 1904. On the Propagation of Tremors over the Surface of an Elastic Solid. Philosophical Transactions of the Royal Society. A 203 1-42. Miller, G. F. and H. Pursey, 1954. The Field and Radiation Impedance of Mechanical Radiators on the Free Surface of a Semi-Infinite Isotropic Solid. Proceedings of the Royal Society of London Series A, Mathematical and Physical Sciences. 233 [1155] 521-541. Mitra, M. 1964. Disturbance Produced in an Elastic Half-Space by Impulsive Normal Pressure. Proceedings of the Cambridge Philosophical Society. 69: 683-696.

369
Figures

1 0.75

Normalized Displacement

0.5 0.25 0 -0.25 -0.5 -0.75 -1 -1


Radial Displacement Vertical Displacement

-0.5

0.5

Non-Dimensional Time
Figure A.1: Disturbance in the surface displacement as described by Lambs equations [1904]. Numerical plots of Equations {A.1.126} and {A.1.127} for, q0, and, w0, are depicted.

370

(a)
0.75
Radial Displacement

Rayleigh Wave

Normalized Displacement

0.5 0.25 0 -0.25 -0.5

Shear Wave

Longitudinal Wave
-0.75 -1

0.5

1.5

2.5

Non-Dimensional Time

1
Vertical Displacement

(b)

Normalized Displacement

0.75

Rayleigh Wave
0.5

Shear Wave
0.25

Longitudinal Wave

-0.25

0.5

1.5

Non-Dimensional Time

Figure A.2: Schematic solutions of the entire displacement wave based on Lambs Equations {A.1.136} and {A.1.137}. Displacements caused by the longitudinal and shear wave pulses are estimated here only in form because they could not be determined with this method. The distance between the waves and the relative amplitude of the waves have been altered to show the profiles of the different waves. (a) Radial displacement. (b) Vertical displacement.

371

1 0.9 0.8

Longitudinal Wave

Normalized Displacement

0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

-0.1 -0.2 -0.3 -0.4 -0.5 0 5E-06

Rayleigh Wave Shear Wave

1E-05

1.5E-05

Time (s)
Figure A.3: Mitras displacement along the impact surface of the specimen in the radial direction. This solution was arrived at by numerically solving of Equation {A.2.88}. The longitudinal (PWave), Shear (S-Wave) and Rayleigh wave (R-Wave) components are all visible. The differentiation of this solution provides the strain in the radial direction.
[VP9]

You might also like