.uhl; - u.- ,.:.: .,.a...:. .113; 1 335... a mfi‘i‘ 5755‘; This is to certify that the thesis entitled THE HYDROTHERMAL BAROHYGROUS CORROSION OF ALUMINA CERAMICS presented by Bejeir Delayne Brooks has been accepted towards fulfillment of the requirements for M. S. degree in Chemical Engineering & Materials Science Date W01 MSU is an Affirmative Action/Equal Opportunity Institution 0-7639 LIBRARY Michigan State University PLACE IN REIURN Box to remove this checkout from your record. TO AVOID FINES retum on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cJCIRC/DateDm.p65-p.15 THE HYDROTHERMAL BAROHYGROUS CORROSION OF ALUMINA CERAMICS BY Bejeir Delayne Brooks A THESIS Submitted to Michigan State University In partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Chemical Engineering and Materials Science 2002 Abstract THE HYDROTHERMAL BAROHYGROUS CORROSION OF ALUMINA CERAMICS By Bejeir Delayne Brooks Hydrothermal corrosion is a process wherein both water and temperature are simultaneously involved in the corrosion of a material. Prior to this work, there has not been an extensive study of the hydrothermal corrosion phenomena in alumina. For ceramics, much of the hydrothermal corrosion literature deals with hydrothermal acidic or alkaline attack and not corrosion due to a nominally neutral solution of deionized water. Although, hydrothermal corrosion only occurs below the boiling or evaporation point of the liquid media, for the present work and most other work dealing with hydrothermal corrosion of any sort (alkaline thru acidic) the substrate to be corroded is placed within a closed pressure vessel adding the elements of pressure (baro-) and steam (hygrous) to - the hydrothermal corrosion environment. This thesis involves hydrothermal barohygrous corrosion of two different grades of alumina (94% and 99.5% purity) Alzoa. Specimens were treated over the first 50 hours of corrosion at 250°C, with a majority of the AD-94 specimens scrutinized for <75 hours and ADS-995 specimens scrutinized from O to 50 hours. The specimens were analyzed through atomic force microscopy with focused investigations performed through other methods as needed (scanning electron microscopy, electron dispersive analysis and x-ray diffraction). Copyright Bejeir Delayne Brooks August 2, 2002 For Reginald, Rebecca, Telaekah and my thirteen times removed grandparents on both sides, because without them I would not be here. To Doris without whom I would have been content with my Bachelors degree. ACKNOWLEDGEMENTS I would like to acknowledge the aid and assistance given me on my thesis from: Benjamin Simkin, Benjamin Tyszka, Boon-Chai Ng, Chee Kuang Kok, ' Craig Carrier, Dr. Darren Mason, Dr. David Grumman, Dr. Gary Cloud, Dr. Haiping Geng, Dr. John Heckman, Dr. K.N. Subramanian. Dr. Ki-Yong Lee, Dr. Martin Crimp, Dr. Patrick Kwon, Dr. Per Askeland, Dr. Richard Schalek, Dr. Ronald Averill, Dr. Stanley Flegler, Dr. Thomas Bieler, Ewa Danielewicz, John Hulbert, Jong-gi Lee, Kenneth Geelhood, Martin Traub, Raelynn Deller, Richard Stabley, Theodore Manko, Timmothy Hoepfner, Timmothy Monroe, Wangyang Ni. I would especially like to thank William Havens and David Enslin who wrote the programs I used to section the AFM micrographs and of course Dr. Eldon Case who helped make this thesis possible. TABLE OF CONTENTS LIST OF TABLES vi LIST OF FIGURES viii CHAPTER 1: INTRODUCTION 1 1.1. Grain Boundary Attack 1 1.2. Hydrothermal corrosion of alumina ceramics [5] 2 1.3. Delayed failure of a porous alumina [8] 6 1.4. Reaction of alumina ceramics with saturated steam [9] 10 1.5. Grain Boundary Segregation under Hydrothermal Corrosion 14 1.6. Grain boundary segregation on A|203 [15] 15 1.7. Surface segregation of calcium in dense alumina exposed to steam 17 and steam +CO [10] 1.8. Grain boundary segregation in MgO-doped Al203 [16] 20 1.9. The Effects of Hydrothermal Corrosion on Non-Oxide Ceramics 22 1.10. Corrosion behavior of silicon carbide in 290°C water [20] 22 1.11. Degradation of SiC fibres in high-temperature, high-pressure 28 water [21] 1.12. Hydrothermal oxidation of Si3N4 powder [26] 34 1.13. The Effects of Hydrothermal Corrosion on Glasses 37 1.14. Hydrothermal corrosion of Iithia disilicate glass and glass- 37 ceramics [28] 1.15. Summary of the Literature 42 CHAPTER 2: 44 EXPERIMENTAL PROCEDURES 2.1. Material Properties 44 2.2. Sectioning Procedure 44 2.3. Mounting Procedure 47 2.4. Polishing Procedure 48 2.5. Procedure for the Loading and Unloading of the Digestion Bomb 49 [32] 2.6. Digestion Bomb Furnace Procedure 56 2.7. Atomic Force Microscope Examination [33] 59 CHAPTER 3: RESULTS AND DISCUSSION 63 3.1. Hydrothermal Attack 63 3.2. Roughness Data 141 3.3. X-Ray Diffraction 147 3.4. Mass Changes 158 3.5. Sectioning of the Atomic Force Microscopy Micrographs 158 vi 3.6. Electron Dispersive Study 3.7. Scanning Electron Microscope Analysis 3.8. Surface Morphology 3.9. Summary and Conclusions 3.10. Future Work REFERENCES APPENDIX APPENDIX A: Slip Systems for Alumina APPENDIX B: Elastic Moduli of Alumina APPENDIX C: Sectioning Data vii 1 89 1 92 201 202 205 206 211 212 214 216 LIST OF TABLES Table 1.2.1: Characteristics of different purities of Alumina Ceramics [5] Table 1.3.1: Breaking strengths of Alumina bars [8] Table 1.3.2: Composition of 65.1% dense alumina [8] Table 1.3.3: Mean fatigue life for porous alumina bars [8] Table 1.3.4: Values of a0 and a1 [8] Table 1.4.1: Chemical Properties 96% alumina [9] Table 1.4.2: Modulus of Rupture data for a 96% pure alumina [9] Table 1.10.1: Compositions of testing solutions [20] Table 1.11.1: Properties of the Tyranno fibers [21] Table 2.1.1: Properties of Coors AD-94 and ADS 995 [31] Table 3.3.1: X-Ray Diffraction Analysis of Non-Hydrothermally Attacked Specimens [35] Table 3.3.2: X-Ray Diffraction Analysis of 10-hour Hydrothermally Attacked Specimens [35] Table 3.3.3: X-Ray Diffraction Analysis of 50-hour Hydrothermally Attacked Specimens [35] Table 3.3.4: X-Ray Diffraction Analysis of Aluminum Mounting Slide [35] Table 3.3.5: X-Ray Diffraction Standards for Diaspore and Boehmite [35] Table 3.3.6: X-Ray Diffraction Standard for Wairakite (Pseudo-Cubic, Monoclinic) [35] Table 3.3.7: Correlation between the d-spacings of the or-Alumina Standard and different stages of the Alumina Specimen Hydrothermal Attack [35] Table 3.3.8: Correlation between the d-spacings of the Boehmite Standard and different stages of the Alumina Specimen Hydrothermal Attack [35] Table 3.3.9: Correlation between the d-spacings of the Diaspore Standard and different stages of the Alumina Specimen Hydrothermal Attack [35] viii 149 150 151 152 153 154 155 156 157 Table 3.3.10: Correlation between the d-spacings of the Wairakite Standard and different stages of the Alumina Specimen Hydrothermal Attack [35] Table 3.5.1: Regressions from the grain boundary etching of ADS-995 attacked at 250°C for 10 to 15 hours as a function of the depth from the specimen surface Table 3.5.2: Regressions from the grain boundary etching of AD-94 attacked at 250°C for 1 to 2 hours as a function of the depth from the specimen surface Table 3.5.3: Regressions from the grain boundary etching of ADS-995 attacked at 250°C for 10 to 15 hours as a function of the depth from the specimen surface and the average grain size Table 3.5.4: Regressions from the grain boundary etching of AD-94 attacked at 250°C for 1 to 2 hours as a function of the depth from the specimen surface and the average grain size Table 3.5.5: Regressions from the grain boundary etching of ADS-995 attacked at 250°C for 10 to 15 hours Table 3.5.6: Regressions from the grain boundary etching of AD-94 attacked at 250°C for 1 to 2 hours Table 3.1: Single crystal elastic constants of alumina in the trigonal system [40] Table 8.2: The reduced trigonal elasticity matrix representing crystallographic classes 32, -3m and 3m, using compliance constants the matrix is symmetric about the main diagonal [39] Table C.1: Area present for the sectioning data of ADS-995 Table 0.2: Area present for the sectioning data of AD-94 ix 157 163 164 165 166 167 168 215 215 216 216 LIST OF FIGURES Figure 2.2.1: lsomet Low-speed diamond saw: A) Diamond blade, B) Rotating specimen holder, C) Lever Arm, D) U.S. Quarter Figure 2.5.1: Side view of the Parr Digestion Bomb inner part, a) Top View, b) Side View, 0) Bottom View: A) Stainless bomb body; B) Teflon Cup; C) Corrosion disc (0.05 mm thick); D) Rupture disc (0.254 mm thick); E) Teflon cover; G) U.S. quarter. Figure 2.5.2: Bottom view of the Parr Digestion Bomb outer parts, a) Bottom View, b) Side View, c) Top View: A) Bronze plated screw cap; B) Stainless steel pressure plate; C) Spring; D) Nickel bottom plug; E) Stainless steel top cap; G) U.S. quarter. Figure 2.5.3: A) Spanner-Jack, used to tighten down the top and bottom portions of the digestion bomb; B) U.S. quarter. Figure 2.6.1: Machined lower brick A) 6.4 cm diameter X 1.35 cm tall cylinder; B) 2.5 cm X 1.7 cm tall cylinder; C) 0.9cm X 3.23 cm tall cylinder; D) U.S. quarter. Figure 2.6.2: Bottom view of upper refractory brick A) 22.75 cm X 11.35 cm X 2.73 cm tall; B) Inner raised semi-cylinder 14.75 cm diameter X 3.7 cm tall; C) Through-hole for thermocouple D) U.S. quarter. Figure 2.6.3: Side view of upper refractory brick A) 22.75 cm X 11.35 cm X 2.73 cm tall; B) Inner raised semi-cylinder 14.75 cm diameter X 3.7 cm tall; C) U.S. quarter. Figure 2.6.4: Reverse side view of upper refractory brick A) 22.75 cm X 11.35 cm X 2.73 cm tall; B) Inner raised semi-cylinder 14.75 cm diameter X 3.7 cm tall; C) U.S. quarter. Figure 2.7.1: Silicon Nitride Probe Schematics [34] Figure 2.7.2: Micrographs of the silicon nitride probe a) 800x b) 5000X [34] Figure 3.1.1: AFM micrograph of Coors ADS-995, with no hydrothermal treatment, polished to a 1 um finish, 5.000 urn scan size, 4.069 H2 scan rate, 256 samples; a) surface view - z-range: 1 uni/division, b) top view. Figure 3.1.2: AFM micrograph of Coors AD-94, with no hydrothermal treatment, polished to a 1 pm finish, 9.667 um scan size, 15.26 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. 46 51 52 55 57 57 58 58 62 62 65 67 Figure 3.1.3: AFM micrograph of Coors ADS-995, hydrothermally attacked for 30 minutes at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.1 .4: AFM micrograph of Coors AD-94, hydrothermally attacked for 30 minutes at 250°C, 5.000um scan size, 6.104 H2 scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.1 .5: AFM micrograph of Coors ADS-995, hydrothermally attacked for 1 hour at 250°C, 10.00pm scan size, 16.28Hz scan rate, 512 samples; a) surface view - z-range: him/division, b) top view. Figure 3.1 .6: AFM micrograph of Coors AD-94, hydrothermally attacked for 1 hour at 250°C, 20.00pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: writ/division, b) top view. Figure 3.1.7: AFM micrograph of Coors AD-94, hydrothermally attacked for 1.5 hours at 250°C, 20.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1um ldivision, b) top view. Figure 3.1 .8: AFM micrograph of Coors AD-94, hydrothermally attacked for 2 hours at 250°C, 10.00pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: Ium Idivision, b) top view. Figure 3.1.9: AFM micrograph of Coors AD-94, hydrothermally attacked for 2.5 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 um Idivision, b) top view. Figure 3.1.10: AFM micrograph of Coors ADS-995, hydrothermally attacked for 3 hours at 250°C, 10.00 pm scan size, 9.766 Hz scan rate, 512 samples; a) surface view - z-range: 1 pm Idivision, b) top view. Figure 3.1 .11: AFM micrograph of Coors AD-94, hydrothermally attacked for 3 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 um /division, b) top view. Figure 3.1 .12: AFM micrograph of Coors AD-94, hydrothermally attacked for 3.5 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.1.13: AFM micrograph of Coors AD-94, hydrothermally attacked for 4 hours at 250°C, 10.00 um scan size, 16.28 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. xi 69 71 73 75 78 80 82 85 87 90 92 Figure 3.1.14: AFM micrograph of Coors AD-94, hydrothermally attacked for 4.5 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.1.15: AFM micrograph of Coors ADS-995, hydrothermally attacked for 5 hours at 250°C, 10.00 pm scan size, 4.883 H2 scan rate, 512 samples; a) surface view - z-range: 1 rim/division, b) top view. Figure 3.1.16: AFM micrograph of Coors AD-94, hydrothermally attacked for 5 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 um/division, b) top view. Figure 3.1.17: AFM micrograph of Coors AD-94, hydrothermally attacked for 5.5 hours at 250°C, 10.00 um scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 rim/division, b) top view. Figure 3.1.18: AFM micrograph of Coors AD-94, hydrothermally attacked for 6 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 uni/division, b) top view. Figure 3.1.19: AFM micrograph of Coors AD-94, hydrothermally attacked for 6.5 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 rim/division, b) top view. Figure 3.1.20: AFM micrograph of Coors AD-94, hydrothermally attacked for 7 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.1.21: AFM micrograph of Coors AD-94, hydrothermally attacked for 7.5 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.1.22: AFM micrograph of Coors ADS-995, hydrothermally attacked for 10 hours at 250°C, 10.00 pm scan size, 6.104 Hz scan rate, 256 samples; a) surface view - z-range: 1 rim/division, b) top view. Figure 3.1.23: AFM micrograph of Coors AD-94, hydrothermally attacked for 10 hours at 250°C, 10.00 pm scan size, 9.766 H2 scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.1.24: AFM micrograph of Coors ADS-995, hydrothermally attacked for 12.5 hours at 250°C, 10.00 pm scan size, 16.28 Hz scan rate, 512 samples; a) surface view - z-range: 1 rim/division, b) top view. xii 95 97 99 102 104 106 108 110 113 115 117 Figure 3.1.25: AFM micrograph of Coors ADS-995, hydrothermally attacked for 15 hours at 250°C, 10.00 pm scan size, 16.28 Hz scan rate, 512 samples; a) surface view — z-range: 1 urn/division, A) grain ~2p.m in size, B) grain larger than 2 pm, b) top view, c) top view without lines. Figure 3.1.26: AFM micrograph of Coors ADS-995, hydrothermally attacked for 20 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view- Figure 3.1.27: AFM micrograph of Coors ADS-995, hydrothermally attacked for 25 hours at 250°C, 10.00 um scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 uni/division, b) top view. Figure 3.1.28: AFM micrograph of Coors ADS-995, hydrothermally attacked for 30 hours at 250°C, 10.00 um scan size, 16.28 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.1.29: AFM micrograph of Coors ADS-995, hydrothermally attacked for 35 hours at 250°C, 10.00 pm scan size, 16.28 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.1.30: AFM micrograph of Coors ADS-995, hydrothermally attacked for 40 hours at 250°C, 10.00 pm scan size, 16.28 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.1.31: AFM micrograph of Coors ADS-995, hydrothermally attacked for 45 hours at 250°C, 10.00 pm scan size, 16.28 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.1.32: AFM micrograph of Coors ADS-995, hydrothermally attacked for 50 hours at 250°C, 10.00 pm scan size, 16.28 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.1 .3: AFM micrograph of Coors AD-94, hydrothermally attacked for 50 hours at 250°C, 10.00 pm scan size, 9.766 H2 scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.2.1: Normalized roughness analysis as a function of the topographic depth for Coors ADS-995, with no hydrothermal treatment, polished to a 1 pm finish. Figure 3.2.2: Normalized roughness analysis as a function of the topographic depth for Coors AD-94, with no hydrothermal treatment, polished to a 1 pm finish. xiii 120 123 125 128 130 132 134 137 139 143 143 Figure 3.2.3: Normalized roughness analysis as a function of the topographic depth for the of Coors ADS-995, hydrothermally attacked for 5 hours at 250°C, Figure 3.2.4: Normalized roughness analysis as a function of the topographic depth for Coors AD-94, hydrothermally attacked for 5 hours at 250°C. Figure 3.2.5: Normalized roughness analysis as a function of the topographic depth for Coors ADS-995, hydrothermally attacked for 10 hours at 250°C. Figure 3.2.6: Normalized roughness analysis as a function of the topographic depth for Coors AD-94, hydrothermally attacked for 10 hours at 250°C. Figure 3.2.7: Normalized roughness analysis as a function of the topographic depth for Coors ADS-995, hydrothermally attacked for 50 hours at 250°C. Figure 3.2.8: Normalized roughness analysis as a function of the topographic depth for Coors AD-94, hydrothermally attacked for 50 hours at 250°C. Figure 3.5.1: AD-94 hydrothermally attacked for 1.5 hours at 250°C at a depth of -5.4 to —25.4 nm. Figure 3.5.2: AD-94 hydrothermally attacked for 1.5 hours at 250°C at a depth of -25.4 to -45.4 nm. Figure 3.5.3: AD-94 hydrothermally attacked for 1.5 hours at 250°C at a depth of —45.4 to —65.4 nm. Figure 3.5.4: AD-94 hydrothermally attacked for 1.5 hours at 250°C at a depth of —65.4 to —85.4 nm. Figure 3.5.5: AD-94 hydrothermally attacked for 1.5 hours at 250°C at a depth of -85.4 to —105.4 nm. Figure 3.5.6: Grain boundary etching as a function of depth for ADS-995 hydrothermally attacked at 250°C for 10 hours Figure 3.5.7: Grain boundary etching as a function of depth for ADS-995 hydrothermally attacked at 250°C for 12.5 hours Figure 3.5.8: Grain boundary etching as a function of depth for ADS-995 hydrothermally attacked at 250°C for 15 hours xiv 144 144 145 145 146 146 169 169 170 170 171 171 172 172 Figure 3.5.9: Grain boundary etching as a function of depth for ADS-995 hydrothermally attacked at 250°C for 10-12.5 hours Figure 3.5.10: Grain boundary etching as a function of depth for ADS- 995 hydrothermally attacked at 250°C for 10-15 hours Figure 3.5.11: Grain boundary etching as a function of depth for AD-94 hydrothermally attacked at 250°C for 1 hour Figure 3.5.12: Grain boundary etching as a function of depth for AD-94 hydrothermally attacked at 250°C for 1.5 hours Figure 3.5.13: Grain boundary etching as a function of depth for AD-94 hydrothermally attacked at 250°C for 1.5 hours (-5 to —275 nm) Figure 3.5.14: Grain boundary etching as a function of depth for AD-94 hydrothermally attacked at 250°C for 2 hours Figure 3.5.15: Grain boundary etching as a function of depth for AD-94 hydrothermally attacked at 250°C for 1-1.5 hours Figure 3.5.16: Grain boundary etching as a function of depth for AD-94 hydrothermally attacked at 250°C for 1-2 hours Figure 3.5.17: Grain boundary etching as a function of depth and average grain size for ADS-995 hydrothermally attacked at 250°C for 10 hours Figure 3.5.18: Grain boundary etching as a function of depth and average grain size for ADS-995 hydrothermally attacked at 250°C for 12.5 hours Figure 3.5.19: Grain boundary etching as a function of depth and average grain size for ADS-995 hydrothermally attacked at 250°C for 15 hours Figure 3.5.20: Grain boundary etching as a function of depth and average grain size for ADS-995 hydrothermally attacked at 250°C for 10- 12.5 hours Figure 3.5.21: Grain boundary etching as a function of depth and average grain size for ADS-995 hydrothermally attacked at 250°C for 10- 15 hours Figure 3.5.2: Grain boundary etching as a function of depth and average grain size for AD-94 hydrothermally attacked at 250°C for 1 hour XV 173 173 174 174 175 175 176 176 177 177 178 178 179 179 Figure 3.5.23: Grain boundary etching as a function of depth and average grain size for AD-94 hydrothermally attacked at 250°C for 1.5 hours Figure 3.5.24: Grain boundary etching as a function of depth and average grain size for AD-94 hydrothermally attacked at 250°C for 1.5 hours (-5 to -275 nm) Figure 3.5.25: Grain boundary etching as a function of depth and average grain size for AD-94 hydrothermally attacked at 250°C for 2 hours Figure 3.5.26: Grain boundary etching as a function of depth and average grain size for AD-94 hydrothermally attacked at 250°C for 1-1.5 hours Figure 3.5.27: Grain boundary etching as a function of depth and average grain size for AD-94 hydrothermally attacked at 250°C for 1-2 hours Figure 3.5.28: Grain boundary etching as a function of depth and average grain size for ADS-995 (10-15 hours) and AD-94 (1-2 hours) hydrothermally attacked at 250°C Figure 3.5.29: Grain boundary etching as a function of root time for ADS- 995 hydrothermally attacked at 250°C for 10-15 hours (-10 to —30 nm) Figure 3.5.30: Grain boundary etching as a function of root time for ADS- 995 hydrothermally attacked at 250°C for 10-15 hours (-30 to —50 nm) Figure 3.5.31: Grain boundary etching as a function of root time for ADS- 995 hydrothermally attacked at 250°C for 10-15 hours (-50 to -70 nm) Figure 3.5.32: Grain boundary etching as a function of root time for ADS- 995 hydrothermally attacked at 250°C for 10-15 hours (-70 to -90 nm) Figure 3.5.33: Grain boundary etching as a function of root time for ADS- 995 hydrothermally attacked at 250°C for 10-15 hours (~90 to -110 nm) Figure 3.5.34: Grain boundary etching as a function of root time for ADS- 995 hydrothermally attacked at 250°C for 10-15 hours (-10 to -110 nm) Figure 3.5.35: Grain boundary etching as a function of root time for AD- 94 hydrothermally attacked at 250°C for 1-2 hours (-5 to —25 nm) Figure 3.5.36: Grain boundary etching as a function of root time for AD- 94 hydrothermally attacked at 250°C for 1-2 hours (-25 to —45 nm) xvi 180 180 181 181 182 182 183 183 184 184 185 185 186 186 Figure 3.5.37: Grain boundary etching as a function of root time for AD- 94 hydrothermally attacked at 250°C for 1-2 hours (-45 to —65 nm) Figure 3.5.38: Grain boundary etching as a function of root time for AD- 94 hydrothermally attacked at 250°C for 1-2 hours (-65 to —85 nm) Figure 3.5.39: Grain boundary etching as a function of root time for AD- 94 hydrothermally attacked at 250°C for 1-2 hours (-85 to —105 nm) Figure 3.5.40: Grain boundary etching as a function of root time for AD- 94 hydrothermally attacked at 250°C for 1-2 hours (-5 to 105 nm) Figure 3.6.1: EDS analysis of Coors ADS-995 with no hydrothermal anack. Figure 3.6.2: EDS analysis of Coors AD-94 with no hydrothermal attack. Figure 3.6.3: EDS analysis of Coors ADS-995 hydrothermally attacked for 5 hours at 250°C. Figure 3.6.4: EDS analysis of Coors AD-94 hydrothermally attacked for 5 hours at 250°C. Figure 3.7.1: Scanning electron image of Coors ADS-995, with no hydrothermal treatment, polished to a 1 pm finish. Figure 3.7.2: Scanning electron image of Coors AD-94, with no hydrothermal treatment, polished to a 1 um finish. Figure 3.7.3: Scanning electron image of Coors ADS-995, hydrothermally attacked for 5 hours at 250°C. Figure 3.7.4: Scanning electron image of Coors AD-94, hydrothermally attacked for 5 hours at 250°C. xvii 187 187 188 188 190 190 191 191 193 195 197 199 Chapter 1: Introduction The goals of the current research are the analysis of the first 2 days (50 hours) of alumina under hydrothermal corrosion. The aluminas of interest are ADS-995 and AD-94 (Coors Ceramic Company: Boulder, Colorado) as the effects of a corrosive atmosphere are believed to attack a material differently based on the purity of the material. The reason for this study is that alumina is often placed within a hydrothermal situation (as pump seals pumps in the nuclear industry for instance) without investigating the effects of the hydrothermal conditions the alumina experiences. 1.1 Grain Boundary Attack In general, impurities segregate to the grain boundary phase of a material [1]. This trend is more pronounced when the material has an exact chemical composition than when the material exists as a solution [1]. An example of this is steel, where there is a certain amount of alloying elements within the iron matrix residing in interstitial and substitutional sites, in this case the grain boundary phase of the iron is not necessarily high in alloying elements [1]. Conversely, for alumina (AI203) where there is a desired chemical composition the grain boundary phase will be high in impurities, leaving the rest of the matrix nominally pure [2]. Additionally the grain boundary phases of ceramics are often amorphous [3], as the grain boundary lacks any long-range order the grain boundary will melt first and solidify last. However, the grain boundaries are locations of grain mismatch and therefore have a higher energy state than the grain boundary phase or the matrix material [2]. The grain boundary phase is equivalent to a thin film [3], however the thickness of a grain boundary may vary greatly. Under corrosive conditions, a material is most likely to be attacked at higher energy locations first therefore the grain boundaries are usually the first to be attacked [1]. Of importance in the current research is the resistance to chemical attack [4]. If the chemical resistance of the grains surpasses the chemical resistance of the grain boundary phase, corrosion damage will occur within the grain boundary phase. If the damage is not occurring at the grain boundaries but instead within the grains then the purity may not be of as much interest. 1.2 Hydrothermal corrosion of alumina ceramics [5] The corrosion behavior and the corresponding degradation of alumina ceramics, by low temperature water were examined by Oda. The bulk alumina (Table 1.2.1) was sectioned into testing coons (3 mm X 4 mm X 35 mm) with a diamond, cutting wheel. The coons were then tested within an autoclave for up to 10 days. The specimens were tested at a temperature of 300°C with a corresponding pressure of 8.6 MPa. The concentrations of ions dissolved in the water during the hydrothermal treatment were determined by inductively coupled plasma atomic spectroscopy (ICP-AES). X-ray diffraction was used to determine the crystalline phases present at the surface, and scanning electron microscopy (SEM) was utilized for surface microscopy. Room temperature fracture strengths, of the corroded and un-corroded samples, were determined using a three-point bending test (30 mm span, 0.5 mm/min crosshead speed). Table 1.2.1: Characteristics of different purities of Aluminat Ceramics [5] 99% pure 99.9% pure 99.99% pure Specimen alumina alumina alumina Al203 (wt%) * 98.7 99.92 99.98 SI02 (wt%) * 1.17 0.04 0.01 N320 (wt%) * 0.09 0.02 0.01 M90 (wt%) * 0.01 0.01 <0.01 CaO (wt%) * 0.03 0.01 <0.01 F9203 (wt%) * <0.01 <0.01 <0.01 Density (g/cma) 3.34 3.93 3.95 Flexural Strength (MPa) * 304 490 588 7 Kyocera Corporation, Kyoto, Japan * Room temperature * EDS Analysis Yi > 0.5 pm) has outlined the individual grains as the grain boundaries have been partially dissolved. There are no obvious polishing artifacts although the arrow indicates an AFM relic (Figure 3.1.8b) along with some dust on the surface. The micron sized circular to rhomboid shaped structures on the surface are the early stages of a scale being formed on the surface, this scale will become more prevalent and apparent in later stages of corrosion. Following 2.5 hours of hydrothermal corrosion, Figure 3.1.9, a large percentage (~45%) of the specimen is obscured by a surface layer. The individual “leafs” of the scale range from 0.25-2 um in-plane lengths and readily form on each other. 77 10 I l l T I a Figure 3.1.7: AFM micrograph of Coors AD-94, hydrothermally attacked for 1.5 hours at 250°C, 20.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 um /division, b) top view. 78 Figure 3.1.7: Continued. 79 I :. I I I T I a Figure 3.1.8: AFM micrograph of Coors AD-94, hydrothermally attacked for 2 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 pm /division, b) top view. 80 Figure 3.1.8: Continued. 81 “\ l I IN I a Figure 3.1.9: AFM micrograph of Coors AD-94, hydrothermally attacked for 2.5 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 pm /division, b) top view. 82 Figure 3.1 .9: Continued. 83 Following three hours of exposure the ADS-995 displays some pitting on the surface (Figure 3.1.10). In Figure 3.1.103, the grain structure is barely visible, along with polishing scratches and some specks of dust on the surface. Grain triple points as associated with roughly triangularly shaped depressions on the specimen surface. The grains have different crystallographic orientations so they appear to have different degrees of contrast under the AFM tip, since the contrast of the AFM image is a function of the electric potential of the surface, which in turn depends on crystallographic orientation. AD-94 corroded for 3 hours, yields a more severely damage structure (Figure 3.1.11). Approximately 98% of the specimen surface is covered in the surface layer. What appears to be the original surface (as indicated by the arrows Figure 3.1.11b) may be areas that are relatively flat in comparison to the other areas of surface scale and not the original specimen surface. At this point, the specimen is no longer experiencing only hydrothermal grooving, but the progressive accumulation of a surface layer. Note the number of peaks, in relation to Figure 3.1.10, within the surface layer indicating a peak growth height ~0.5 um accompanied by valleys of the same scale. However, the grooves that occur during the hydrothermal process are covered by the surface layer so the actual depth of the grooves is difficult to determine. Additionally, there are some artifacts from the AFM on the substrate surface. 84 I l I l a Figure 3.1.10: AFM micrograph of Coors ADS-995, hydrothermally attacked for 3 hours at 250°C, 10.00 pm scan size, 9.766 H2 scan rate, 512 samples; a) surface view - z-range: 1 pm /division, b) top view. 85 Figure 3.1.10: Continued. 86 a Figure 3.1.1 1: AFM micrograph of Coors AD-94, hydrothermally attacked for 3 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 pm /division, b) top view. 87 Figure 3.1.11: Continued. 88 After 3.5 hours of corrosion, the AD-94.has a few AFM discontinuities, yet the surface is still completely occluded by the surface layer (Figure 3.1.12). There are some smooth areas as in Figure 3.1.11, however, the size of the growths is much larger, 1-3 pm in-plane lengths, with an average peak height of ~0.5 pm. The growths are apparently developing as a series parallelograms growing on each other, with the majority of these parallelograms being square or rectangular. Day [Day, 1982], noted that for 96% pure alumina, a cubic wairakite structure developed on the surface after less than 10 days of hydrothermal treatment at 250°C (Section 1.5). After four hours of hydrothermal attack an AD-94 specimen lacked pitting or the signs of corrosion (Figure3.1.13). Grain boundary etching is mostly absent except for a single grain boundary triple point alongside polishing marks. However, there were difficulties acquiring AFM micrographs of AD-94 after 3 hours of corrosion: the surfaces have been degraded so severely that there is considerable noise when the surfaces were being resolved. In particular, the 4- hour AD-94 specimen was very difficult to image in the AFM. Therefore, the surface shown is probably not representative of the majority of the surface of the AD-94 specimen at this level of corrosion. 89 a Figure 3.1.12: AFM micrograph of Coors AD-94, hydrothermally attacked for 3.5 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.1.12: Continued. 91 l I I 1 a Figure 3.1.13: AFM micrograph of Coors AD-94, hydrothermally attacked for 4 hours at 250°C, 10.00 pm scan size, 16.28 Hz scan rate, 512 samples; a) surface view - z-range: 1 uni/division, b) top view. 92 Figure 3.1.13: Continued. 93 AD-94 that has been hydrothermally degraded for 4.5 hours (Figure 3.1.14), again displays damage similar to that seen in Figures 3.1.11-12. Some smooth portions of the surface visible (Figure 3.1.14b) yet they are not believed to be the AD-94 surface when compared to Figure 3.1.14a. The surface layers peak heights reach 1 pm with troughs on the order of ~0.6 urn deep. The in- plane lengths vary from 04-3 pm, however the majority of the scales have an in- plane length above 1.5 pm. Figure 3.1.15ab portrays polishing scratches, dust, and possible scale developing on the surface of ADS-995 with 5 hours of exposure. However, the grains themselves are not readily apparent in this micrograph. There also appears to be a few (~4) grains that erupt from the surface of the specimen. A 5-hour trial of AD-94, continues to display complete coverage by the surface layer with an average peak height of ~1 um (Figure 3.1.16). Figure 3.1.14 is mostly composed of “tall” peaks in comparison to the peak heights in Figure 3.1.16, which is due to the fact, that the AFM could not resolve an image in a more disrupted area of the specimen surface. 94 l l I I l a Figure 3.1.14: AFM micrograph of Coors AD-94, hydrothermally attacked for 4.5 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. 95 Figure 3.1.14: Continued. 96 I . I I I a Figure 3.1.15: AFM micrograph of Coors ADS-995, hydrothermally attacked for 5 hours at 250°C, 10.00 pm scan size, 4.883 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. 97 Figure 3.1.15: Continued. 98 I I l a Figure 3.1 .16: AFM micrograph of Coors AD-94, hydrothermally attacked for 5 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. 99 Figure 3.1.16: Continued. For AD-94 hydrothermally treated for 5.5 hours, AFM distortion could not be avoided at this or more severe levels of attack as the steepness of the surrounding areas of the specimen affected the movement of the tip on the specimen surface (Figure 3.1.17). Notably, the entire surface is still covered in a film, primarily composed of 1 pm basal scales. There are occasional large grains (2 x 3 x 1 urn tail) that exceed the average layer level of the specimen. Quite likely, wairakite-type cubes are forming on the surface and the micrograph is of an area that rests between the structures. The 6-hour hydrothermal attack of AD-94 (Figure 3.1.18), has distortions within the image from the AFM. The average surface layer height is low, as the surface layer is forming into possibly wairakite-type structures. The surface pitting is present at the grain boundary triple points. After 6.5 hours of corrosion displays AD-94 with a mostly low (<0.1 um) surface layer height (Figure 3.6.19). Also present are structures (~7 x 5 pm with a height of ~1 pm). There are conglomerations of numerous rectangular scales. Following 7 hours of hydrothermal attack (Figure 3.1.20), the AD-94 surface is completely coated by a surface layer that contains structures (~1 um tall) with significant lateral (~5 x 9 pm) dimensions. The large pit present is over 1 pm deep and 2 pm wide may correspond to a grain boundary triple point. At 7.5 hours of continuous attack the AD-94 surface is revealed that is completely covered by the surface layer (Figure 3.1.21). The surface layer reached 1.3 pm in height and has a triple point pit roughly ~1.5 um deep. 101 ”M l I I f I a Figure 3.1.17: AFM micrograph of Coors AD-94, hydrothermally attacked for 5.5 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. 102 Figure 3.1.17: Continued. 103 a Flgure 3.1.18: AFM micrograph of Coors AD-94, hydrothermally attacked for 6 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. Figure 3.1.18: Continued. 105 L I v i l a Figure 3.1.19: AFM micrograph of Coors AD-94, hydrothermally attacked for 6.5 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. 106 Figure 3.1.19: Continued. 107 I l l a Figure 3.1.20: AFM micrograph of Coors AD-94, hydrothermally attacked for 7 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 uni/division, b) top view. 108 Figure 3.1.20: Continued. 109 l l I a Figure 3.1 .21: AFM micrograph of Coors AD-94, hydrothermally attacked for 7.5 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 um/division, b) top view. 110 Figure 3.1 .21 : Continued. lll Subsequent to 10 hours of hydrothermal attack, the grain structure of the ADS-995 can be seen along with scratches due to polishing (Figure 3.1.22). There is some visible pitting (roughly ten ~0.2 um pits per 100 umz), and a small degree of dust (<10 specks) on the surface. The pits occur mostly at the grain boundary intersections, or triple points, are roughly triangular depressions and in this image that are less than 0.25 micron in their in-plane directions. An AD-94 specimen having experienced 10-hour hydrothermal corrosion displays complete coverage by the surface layer by a uniform distribution of ~1 um in-plane scales (Figure 3.1.23). The scales have an average height of ~0.4 um, and there are a few pits on the surface. Figure 3.1.24 depicts the ADS-995 surface exposed to 12.5 hours of treatment. The grain boundaries and pits may be seen in this micrograph alongside some specks (<10) of dust and a few grains as indicated by arrows (Figure 3.1.24a) that appear to be erupting from the surface of the specimen. However, when compared to the morphology of the AD-94 surface scale what appears to be an emerging grain may in fact be a precursor of that scale. Such may be the case as the ADS-995 has a higher purity and the scaling reaction should take longer to begin as well as progress. In order for a grain to erupt from the surface, either all the surrounding material would have to be removed or the grain would have to begin to grow, neither of these scenarios is likely however grains can still “appear" to emerge by having their crystallographic orientation being such that their surfaces appear to “come out" more from the bulk. 112 l r I l a Figure 3.1 .22: AFM micrograph of Coors ADS-995, hydrothermally attacked for 10 hours at 250°C, 10.00 pm scan size, 6.104 H2 scan rate, 256 samples; a) surface view - z-range: 1 urn/division, b) 10p view. 113 Figure 3.1 .22: Continued. 114 ”fl l L 1 1 1 a Figure 3.1.23: AFM micrograph of Coors AD-94, hydrothermally attacked for 10 hours at 250°C, 10.00 pm scan size, 9.766 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. 115 Figure 3.1 .23: Continued. 116 I : | I l j a Figure 3.1.24: AFM micrograph of Coors ADS-995, hydrothermally attacked for 12.5 hours at 250°C, 10.00 pm scan size, 16.28 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. 117 Figure 3.1.24: Continued. 118 Individual grains of the ADS-995 are readily apparent as shown within the outlined areas A and B, in Figure 3.1.25 after 15 hours of corrosive attack. Dust is not too apparent in this figure. Here the pitting at the grain boundaries is the most discernable when compared with any of the other ADS-995 figures. The mean grain size is approximately 2 pm (Table 2.2.1) with some grains larger than 2 pm and others smaller than 2 pm. Unfortunately, there are still some polishing scratches on the surface. It is expected that polishing artifacts be visible on the surface of most specimens that have not been completely occluded by a surface scale. The polishing was only conducted to 1 pm and at the magnification of the AFM, scratches in the range present (fractions of microns across) should be visible. ADS-995 after 20 hours of exposure has a highly corroded surface morphology displaying surface scale and some pitting. The scale covers the entirety of the surface and hinders viewing of the surface (Figure 3.1.26). There are no apparent polishing marks. In Figure 3.1.27, ADS-995 following 25 hours of hydrothermal corrosion, the surface remains mostly covered with scale. The corrosion of the surface appears to be less severe than that of the 20-hour specimen. Thus, either the surface was not completely covered or the surface scale was not in a particularly stable state. 119 I a. h i l a Figure 3.1.25: AFM micrograph of Coors ADS-995, hydrothermally attacked for 15 hours at 250°C, 10.00 pm scan size, 16.28 Hz scan rate, 512 samples; a) surface view — z-range: 1 uni/division, A) grain ~2um in size, B) grain larger than 2 pm, b) top view, 0) top view without lines. 120 Figure 3.1.25: Continued. 121 Figure 3.1.25: Continued 122 L I I I a Figure 3.1.26: AFM micrograph of Coors ADS-995, hydrothermally attacked for 20 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. 123 Figure 3.1 .26: Continued. 124 'lJll a Figure 3.1.27: AFM micrograph of Coors ADS-995, hydrothermally attacked for 25 hours at 250°C, 10.00 pm scan size, 12.21 Hz scan rate, 512 samples; a) surface view — z-range: 1 urn/division, b) top view. 125 Figure 3.1.27: Continued. 126 Figure 3.1.28 shows the ADS-995 surface after 30 hours of attack. Scale deposits are present that are approximately 1.5 pm in their in-plane directions (Figure 3.1.28ba). In the top view of the ADS-995 (Figure 3.1.28b) that has been treated for 30 hours, there are a few visible artifacts from the AFM. However, the surface has apparently been completely covered by the growth or surface layer as seen in the surface view (Figure 3.1.28a). The surface of the ADS-995 remains coated with the surface layer after 35 hours of hydrothermal exposure (Figure 3.1.29). However, the specimens surface topology is smoother than that shown in Figure 3.1.28. Fortunately, there are no visible scratches from the polishing. The scale deposits in this micrograph are nearly a micron in height and ~6 pm in the in-plane direction. A large number of growths and possibly dust are present on the surface of the 40-hour hydrothermally attacked specimen for ADS-995 (Figure 3.1.30). Few polishing marks are visible on the surface. The topology of the surface layer has notably changed from the surface layers revealed in previous ADS-995 micrographs (Figures 3.1.26,27,28,29) in that the grain boundaries are again visible on the surface. ADS-995 after 45 hours of hydrothermal treatment (Figure 3.1.31) reveals pits that are again apparent on the surface along with polishing marks and dust. Again the surface layer has not deposited on the alumina surface. 127 l I I I a Figure 3.1.28: AFM micrograph of Coors ADS-995, hydrothermally attacked for 30 hours at 250°C, 10.00 pm scan size, 16.28 Hz scan rate, 512 samples; a) surface view - z-range: 1 um/division, b) top view. 128 Figure 3.1.28: Continued. 129 l l I l | a Figure 3.1.29: AFM micrograph of Coors ADS-995, hydrothermally attacked for 35 hours at 250°C, 10.00 pm scan size, 16.28 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. 130 Figure 3.1.29: Continued. 131 ”ll l l l l a Figure 3.1.30: AFM micrograph of Coors ADS-995, hydrothermally attacked for 40 hours at 250°C, 10.00 pm scan size, 16.28 Hz scan rate, 512 samples; a) surface view - z-range: 1 um/division, b) top view. 132 Figure 3.1.30: Continued. 133 a Figure 3.1 .31: AFM micrograph of Coors ADS-995, hydrothermally attacked for 45 hours at 250°C, 10.00 pm scan size, 16.28 Hz scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. 134 Figure 3.1.31: Continued. 135 The surface of ADS-995 hydrothermally corroded for 50 hours (Figure 3.1.32) has a surface devoid of any surface layer. Grain boundaries and polishing artifacts are present on the surface alongside pits and dust akin to Figure 3.1.22. This supports the theory that the surface layer is deposited during the 12.5-35 hour regime, and does not deposit during the 40-50 hour regime. Conversely, for AD-94 following 50hours of hydrothermal treatment, the specimen surface is completely occluded by the surface layer (Figure 3.1.33). The largest structure present is pyramidal in shape and is more than 1 pm in height (Figure 3.1.33b). Surface layers may precipitate during the cooling cycle of the treatment. impurities in the grain boundary phase, or those impurities and water, combine to form the scale that appears on the specimen surface. The thermocouple is turned off during the cooling cycle of the furnace ands there is no thermocouple within the confines of the digestion bomb, thus reactions that would form isotherms in the cooling rate are neither recordable nor observable. Due to specimen size, isotherms are unlikely to be recordable with a thermocouple as it would be difficult to place a thermocouple on the specimen surface during experimentation. However, a thermocouple placed directly in contact with the digestion bomb may detect that a reaction was taking place. If the reaction is time dependent, which it appears to be, there would be difficulty in determining a change in the temperature as a function of time since the furnace experiences temperature fluctuations. 136 I :1 a Figure 3.1.32: AFM micrograph of Coors ADS-995, hydrothermally attacked for 50 hours at 250°C, 10.00 pm scan size, 16.28 Hz scan rate, 512 samples; a) surface view - z-range: 1 ILITl/dIVISIOi'l, b) top view. 137 Figure 3.1 .32: Continued. 138 I; I I I a Figure 3.1.33: AFM micrograph of Coors AD-94, hydrothermally attacked for 50 hours at 250°C, 10.00 pm scan size, 9.766 H2 scan rate, 512 samples; a) surface view - z-range: 1 urn/division, b) top view. 139 Figure 3.1 .33: Continued. 3.2 Roughness Analysis An analysis was made of the general surface roughness of each specimen (Figures 3.2). The roughness data was taken as a measure of the number of events that occur within a series of depths within an AFM micrograph. The roughness data was examined as histograms where the full range of depth is portrayed in 1000 steps. The data was collected and the “zero” point of the roughness data is a reference point only within a given AFM micrograph. The number of events was normalized with respect to the total number of events. The normalized number of events was then plotted with respect to the topographic depth. A quantitative analysis of the roughness data is beyond the scope of this study. However, the roughness data supports the AFM micrograph data that the degree of corrosion increases for both alumina specimens as the exposure time increases (Figures 3.2). Additionally, the shape of the AD-94 histogram curve changes as the corrosion progresses (Figure 3.2.4,8) indicating that there are multiple regimes where the depth of the corrosion is uniform. These distortions from the ideal curve may represent of grain boundaries, pits, scratches and other defects marring the surface. Due to the asymmetry of the curves, the normal distributions are believed to be the summations of multiple gaussians to accurately describe the surface roughness of the specimens. 141 Both AD-94 and ADS-995 specimens displayed a broadening of the normalization curves as corrosion progresses. Indicating that the longer a specimen is hydrothermally attacked the more damage is done to the surface. 142 ADS-995 Normalized Event Histogram - 0 hours 0.09 - . 0.08 - O .3 : 0.07 - % . 0.06 - '5 ° _ E °. 0.04 - g 3 0.03 - 2 .: 0.02 ~ 12 0.01 J -60 -40 -20 0 20 40 60 Topographic Depth (nm) Figure 3.2.1: Normalized roughness analysis as a function of the topographic depth for Coors ADS-995, with no hydrothermal treatment, polished to a 1 pm finish. AD—94 Normalized Event Histogram - 0 hours 0.025 — g 0.02 - 0 I 5 ' 0 015 .2 3. ' 8 1‘ .5 ‘ i 0 01 - a F ‘ E ' 1 O I I Z I i 0.005 4 -25 -20 -15 -10 -5 0 5 10 15 20 Topographic Depth (nm) Figure 3.2.2: Normalized roughness analysis as a function of the topographic depth for Coors AD-94, with no hydrothermal treatment, polished to a 1 pm finish. 143 ADS-995 Normalized Event Histogram - 5 hours 0.035 - 0.03 a 0.025 - O o 0"“. O 0.02 z 00 0.015 ~ «00000 0.01 . Normalized # of events 0.005 ‘ F r i l -200 -150 -100 -50 0 50 100 150 200 Topographic Depth (nm) ‘ q I Figure 3.2.3: Normalized roughness analysis as a function of the topographic depth for the of Coors ADS-995, hydrothermally attacked for 5 hours at 250°C, AD-94 Normalized Event Histogram - 5 hours 0.0045 1 - 0.004 - f. 00035 - 0.003 ~ 0.0025 - 0.002 - 0.0015 ~ 0.001 - Normalized # of events -600 -400 -200 0 200 400 600 Topographic Depth (nm) Figure 3.2.4: Normalized roughness analysis as a function of the topographic depth for Coors AD-94, hydrothermally attacked for 5 hours at 250°C. 144 ADS-995 Normalized Event Histogram - 10 hours 0.012 - g.‘ . l 1, 0.008 . f “ 0.006 - i l '. 0.004 - 0.002 5 l r l _ r l i ~150 -100 -50 0 50 100 150 200 Topographic Depth (nm) Normalized # of events 5' Figure 3.2.5: Normalized roughness analysis as a function of the topographic depth for Coors ADS-995, hydrothermally attacked for 10 hours at 250°C. AD-94 Normalized Event Histogram - 10 hours 0.012 - I 0.01 ~ % f. E, i: 0.008 ~ 0 .I at - - I 0.006 - 8 I r -- t u. 5 0.004 - 8 2 0.002 _ I l I -200 -150 -100 -50 0 50 100 150 200 Topographic Depth (nm) Figure 3.2.6: Normalized roughness analysis as a function of the topographic depth for Coors AD-94, hydrothermally attacked for 10 hours at 250°C. 145 ADS-995 Normalized Event Histogram - 50 hours 0.012 5 A 0.01 ' : 0.008 ~ I I i i I J l r r l i l l 1 l -200 -150 -100 -50 0 50 100 150 200 Topographic Depth (nm) I E I ' 0.006 ‘1 I \ 0.004 - Normalized # of events Figure 3.2.7: Normalized roughness analysis as a function of the topographic depth for Coors ADS-995, hydrothermally attacked for 50 hours at 250°C. AD-94 Normalized Event Histogram - 50 hours 0.007 - ' 0.006 - 13 . o _ P. 0.005 - > ' I 3 g I O 0.004 - at l ‘0 g 0.003 « a E 0 Z I I T -800 -600 -400 -200 0 200 400 600 800 Topographic Depth (nm) Figure 3.2.8: Normalized roughness analysis as a function of the topographic depth for Coors AD-94, hydrothermally attacked for 50 hours at 250°C. 146 3.3 X-Ray Diffraction The AD-94 and ADS-995 specimens were observed via X-ray diffraction. Specimen conditions were: as-polished, and hydrothermally attacked for 10 and 50 hours. The specimens were mounted on an aluminum slide to minimize the sources of foreign artifacts (i.e. an amorphous glass regime form a glass slide, or iron from a steel slide), within the X-ray analysis. The specimens were then compared with powder diffraction standard data (Tables 3.3.1-3) [35], data was reported (Tables 3.3.7-10) in terms of interplanar spacing (d-spacing) and relative intensities (Rel. lnt.). All of the specimens contained necessary peaks for both alumina and aluminum (Tables 3.3.1-4) [35]. Furthermore, specimens had different relative intensities at given interplanar spacings, when compared with the standards Tables 3.3.1-6 [35]. The reason for the discrepancies in the intensities is that both ADS-995 and AD-94 have a preferred orientation or texture [36]. Preferred orientations are the norm in most materials: metals, polymers and ceramics and are a result of grain rotations during plastic deformation [36]. Some of the d-spacings of the standards were very similar to those of the experimental specimens (Tables 3.3.7-10), however there are d-spacings that do not correspond to the standards: aluminum, a-alumina, boehmite, diaspore, or wairakite. Additionally, there were traces of boehmite, diaspore, and wairakite in the non-attacked ADS-995 specimen. As the purity of the ADS-995 contradicts these readings, the calibration of the diffractometer comes into question. However it can be readily determined that there is no appreciable growth of 147 either the boehmite, diaspore, or wairakite phases, as the same peaks are found at each time frame. Since the specimens have a visible growth on the surface (Figures 3.1) the growths are likely amorphous or below the resolution of the diffractometer. 148 Table 3.3.1: X-Ray Diffraction Analysis of Non-Hydrothermally Attacked Specimens [35] a—Alumina Standard ADS-995 0 Hours AD-94 0 Hours d-spacing Rel. Int. d-spacing Rel. Int. d-spacing Rel. lnt. Inc A III, III, 3.479 75 3.471 6 4 6.8805 5 2.552 90 2.5455 10 3.3984 23 2.379 40 2.3741 3 2.5082 16 2.165 1 2.3461 6 2.3433 53 2.085 100 2.2472 2 2.0584 60 1 .964 2 2.0818 1 1 2.0317 75 1 .74 45 2.0568 2 1 .7226 24 1 .601 80 2.0301 100 1 .5873 20 1 .546 4 1 .7381 6 1 .5336 3 1 .514 6 1.5999 16 1 .5025 4 1.51 8 1.5081 2 1.4344 31 1 .404 30 1 .4338 42 1 .3948 33 1 .374 50 1 .4036 5 1.3646 100 1 .337 2 1 .3727 7 1 .2228 22 1.276 4 1.2386 4 1.1839 13 1 .239 16 1 .2339 2 1 .1423 8 1.2343 8 1.2224 28 1.1211 6 1.1898 8 1.0988 2 1.0743 11 1.16 1 1.0425 3 1.0393 15 1.147 6 1.013 5 1.0145 4 1.1382 2 0.9977 3 1.0132 4 1 .1 255 6 0.9297 2 0.9954 4 1.1246 4 0.9078 3 0.9064 25 1.0988 8 0.9046 4 1.0831 4 0.898 17 1.0781 8 1.0426 14 1.0175 2 0.9976 12 0.9857 1 0.9819 4 0.9431 1 0.9413 1 0.9345 4 0.9178 4 0.9076 14 0.9052 4 0.8991 8 0.8884 1 149 Table 3.3.2: X-Ray Diffraction Analysis of 10-hour Hydrothermally Attacked Specimens [35] oz-Alumina Standard ADS-995 10 Hours AD-94 10 Hours d-spacing Rel. Int. d-spacing Rel. lnt. d-spacing Rel. Int. A m, A m, A m, 3.479 75 3.467 8 6.7412 5 2.552 90 2.5444 18 3.3984 23 2.379 40 2.3736 6 2.5082 16 2.165 1 2.3458 7 2.3433 53 2.085 100 2.0812 19 2.0584 60 1.964 2 2.0302 100 2.0317 75 1.74 45 1 .7375 1 1 1.7226 24 1 .601 80 1 .5996 28 1 .5873 20 1 .546 4 1 .5095 3 1.5336 3 1 .514 6 1 .4339 35 1.5025 4 1.51 8 1.4031 10 1.4344 31 1 .404 30 1 .3724 13 1 .3948 33 1 .374 50 1 .2383 7 1 .3646 100 1.337 2 1 .2336 2 1 .2228 22 1 .276 4 1 .2224 24 1 .1839 13 1 .239 16 1 .0987 3 1 .1423 8 1.2343 8 1.0778 2 1.1211 6 1 .1898 8 1 .0424 5 1 .0743 1 1 1.16 1 1.0131 4 1.0393 15 1.147 6 0.9976 5 1.0145 4 1 .1382 2 0.9295 2 1.0132 4 1 .1255 6 0.9077 3 0.9954 4 1 .1246 4 0.9057 5 0.9064 25 1.0988 8 0.9046 4 1.0831 4 0.898 17 1.0781 8 1.0426 14 1.0175 2 0.9976 12 0.9857 1 0.9819 4 0.9431 1 0.9413 1 0.9345 4 0.9178 4 0.9076 14 0.9052 4 0.8991 8 0.8884 1 150 Table 3.3.3: X-Ray Diffraction Analysis of 50-hour Hydrothermally Attacked Specimens [35] a-Alumina Standard ADS-995 50 Hours AD-94 50 Hours d-spacing Rel. Int. d-spacing Rel. lnt. d-spacing Rel. Int. A m, A m, A mg 3.479 75 3.4199 3.4199 7.0098 9 2.552 90 2.5205 2.5205 3.5229 7 2.379 40 2.3479 2.3479 3.4626 1 7 2.165 1 2.0657 2.0657 2.5416 14 2.085 1 00 2.056 2.056 2.3722 50 1 .964 2 2.0299 2.0299 2.3462 7 1 .74 45 1 .7273 1.7273 2.0801 49 1.601 80 1.5914 1.5914 2.03 96 1 .546 4 1 .5026 1 .5026 1.7365 19 1 .514 6 1 .4339 1 .4339 1.5988 24 1 .51 8 1.3975 1.3975 1.5443 4 1.404 30 1 .3671 1.3671 1.5124 4 1 .374 50 1 .2346 1 .2346 1 .4338 27 1 .337 2 1 .2295 1 .2295 1 .4028 32 1 .276 4 1 .2223 1 .2223 1 .3721 100 1 .239 1 6 1 .096 1 .096 1 .2222 17 1.2343 8 1 .0403 1 .0403 1 .1887 1 1 1.1898 8 1.013 1.013 1.1465 7 1.16 1 0.9959 0.9959 1.125 6 1.147 6 0.9061 0.9061 1.0777 10 1.1382 2 0.9048 0.9048 1.0423 13 1.1255 6 1.0173 4 1.1246 4 0.9078 18 1.0988 8 0.9057 4 1.0831 4 0.8993 12 1.0781 8 1.0426 14 1.0175 2 0.9976 12 0.9857 1 0.9819 4 0.9431 1 0.9413 1 0.9345 4 0.9178 4 0.9076 14 0.9052 4 0.8991 8 0.8884 1 151 Aluminum Standard Table 3.3.4: X-RayDiffraction Analysis of Aluminum Mountiqu Slide [35] Aluminum Slide d-spacing Rel. Int. d-spacing Rel. lnt. A In, A III.) 2.338 100 2.3266 5.831241 2.024 47 2.0157 100 1 .431 22 1 .4277 43.61628 1.221 24 1.2184 34.11352 1.169 7 1.0112 6.373178 1.0124 2 0.9281 4.685145 0.9289 8 0.9048 6.328016 0.9055 8 0.8266 8 152 Table 3.3.5: X-Ray Diffraction Standards for Diaspore and Boehmite [35] Basic Aluminum Oxide (Diaspore) Aluminum Oxide Hydroxide (Boehmite) d-spacing Rel. lnt. d-spacing Rel. lnt. A III.) A l/lo 4.71 13 6.11 100 9.99 100 3.164 65 3.214 10 2.346 55 2.558 30 1.98 6 2.434 3 1.86 30 2.386 5 1.85 25 2.356 8 1.77 6 2.317 56 1.662 14 2.131 52 1.527 6 2.077 49 1.453 16 1.901 3 1.434 10 1.815 8 1.412 2 1.733 3 1.396 2 1.712 15 1.383 6 1.678 3 1.369 2 1.633 43 1.312 16 1.608 12 1.303 4 1.57 4 1.224 2 1.522 6 1.209 2 1.48 20 1.178 4 1.431 7 1.171 1 2 1.423 12 1.1609 4 1.4 6 1.1337 6 1.376 16 1.1 152 2 1.34 5 1.0917 2 1.329 6 1.0459 2 1.304 3 1.0281 2 1.289 6 0.9903 2 1.279 1 0.9818 2 1.256 4 0.9506 2 1.243 5 0.931 2 1.218 2 0.9247 2 1.204 4 0.9105 2 1.1783 1 0.9023 2 1.1739 7 0.8937 2 1.1408 3 0.8907 2 1.1003 1 0.866 2 1.0923 3 0.8607 2 0.8316 153 Table 3.3.6: X-Ray Diffraction Standard for Wairakite (Pseudo-Cubic, Monoclinig [3g Wairakite (Pseudo-Cubic, Monoclinic) d-spacing Rel. Int. A l/lo 6.85 40 5.57 80 4.84 40 3.64 30 3.42 60 3.39 100 2.909 50 2.897 30 2.783 10 2.77 10 2.68 40 2.67 10 2.5 5 2.489 40 2.418 30 2.215 40 2.17 5 2.147 10 2.1 15 10 2.095 5 1.996 20 1.867 10 1.857 30 1.844 10 1.708 5 1.696 5 1.595 5 1.586 20 1.487 10 1.354 10 1.343 10 154 Table 3.3.7: Correlation between the d-spacings of the a-Alumina Standard and different stages of the Alumina Specimen Hydrothermal Attack [35] a-Alumina ADS-995 0 ADS-995 ADS-995 AD-94 0 AD-9410 AD-94 50 Standard Hours 10 Hours 50 Hours Hours Hours Hours d-spacing d-spacing d-spacing d-spacing d-spacing d-spacing d-spacing A A 3.479 2.552 2.379 2.085 1.740 1.601 1.546 1.514 1.510 1.404 1.374 1.239 1.234 1.190 1.147 1.138 1.125 1.099 1.078 1.043 1.018 0.998 0.935 0.908 0.905 0.899 3.472 2.546 2.374 2.082 1 .738 1 .600 1 .508 1.404 1.373 1.239 1 .234 1 .099 1.043 1.013 0.998 0.930 0.908 0.906 3.467 2.544 2.374 2.081 1.738 1 .600 1.510 1.403 1.372 1.238 1.234 1.099 1.078 1.042 1.013 0.998 0.930 0.908 0.906 1.591 1.503 1.398 1 .367 1.235 1 .230 1 .096 1 .040 1.013 0.996 0.906 0.905 3.490 1 .593 1 .507 1 .398 1 .368 1.186 1.144 1.122 1.075 1.040 1.015 0.996 0.907 0.905 0.898 1 .503 1 .395 1 .365 1.223 1.184 1.142 1.121 1.074 1.039 1.015 0.995 0.906 0.905 0.898 2.542 2.372 2.080 1 .737 1 .599 1 .544 1 .512 1 .403 1 .372 1.222 1.189 1.147 1.125 1.078 1.042 1.017 0.908 0.906 0.899 155 Table 3.3.8: Correlation between the d-spacings of the Boehmite Standard and different stages of the Alumina Specimen Hydrothermal Attack [35] Boehmite ADS-995 0 ADS-995 ADS-995 AD—94 0 AD-9410 A094 50 Hours 10 Hours 50 Hours Hours Hours Hours d-spacing d-spacing d-spacing d-spacing d-spacing d-spacing d-spacing 2.346 2.346 2.346 2.348 2.343 2.346 1.527 1 .539 1 .534 1 .434 1 .434 1 .434 1 .434 1.429 1 .434 1 .434 1.412 1.404 1 .396 1 .403 1.398 1 .398 1 .395 1 .403 1 .369 1 .373 1 .372 1.367 1 .368 1 .365 1 .372 1 .224 1 .222 1 .222 1 .222 1 .219 1 .223 1.222 1.178 1.186 1.184 1.189 1.134 1.122 1.121 1.125 1 .092 1.099 1 .099 1 .096 1 .046 1 .043 1 .042 1.040 0.990 0.998 0.998 0.996 0.982 0.996 0.951 0.995 0.931 0.930 0.930 0.911 0.908 0.908 0.906 0.908 0.902 0.906 0.906 0.905 0.907 0.906 0.894 0.905 0.906 0.899 0.891 0.898 0.905 0.866 0.898 156 Table 3.3.9: Correlation between the d-spacings of the Diaspore Standard and different stages of the Alumina Specimen HLdrothermal Attack [35] Diaspore ADS-995 0 ADS-995 ADS-995 AD-94 0 AD-9410 AD-94 50 Hours 10 Hours 50 Hours Hours Hours Hours d-spacing d-spacing d-spacing d-spacing d-spacing d-spacing d-spacing 4.710 3.420 2.558 2.546 2.544 2.356 2.357 2.343 2.346 2.077 2.066 2.069 2.080 1 .733 1.738 1 .738 1 .727 1 .729 1 .723 1.737 1.608 1.600 1.600 1.591 1.522 1.512 1 .431 1 .434 1 .434 1 .434 1 .429 1 .434 1.434 1 .400 1 .404 1 .403 1 .398 1 .398 1 .395 1.403 1 .376 1.373 1 .372 1.367 1 .368 1.372 1 .243 1 .239 1 .238 1 .235 1.218 1.222 1.222 1.222 1.219 1.223 1.222 1.141 1.144 1.142 1.147 1.100 1.099 1.099 1.092 1.096 Table 3.3.10: Correlation between the d-spacings of the Wairakite Standard and different stages of the Alumina Specimen Hydrothermal Attack [35] Wairakite ADS-995 0 ADS-995 ADS-995 AD-94 0 AD-9410 AD-94 50 Hours 10 Hours 50 Hours Hours Hours Hours d-spacing d-spacing d-spacing d-spacing d-spacing d-spacing d-spacing 3.420 3.420 2.500 2.508 2.350 2.346 2.346 2.348 2.357 2.343 2.346 2.095 2.082 2.081 2.080 1.732 1.738 1 .738 1.727 1.729 1.723 1.737 1 .595 1.600 1.600 1.591 1.593 1.599 1.586 1.587 1 .437 1 .434 1 .434 1 .434 1 .434 1.434 1 .407 1 .404 1 .403 1.398 1 .398 1 .395 1.403 1.354 1.365 1.215 1.222 1.222 1.222 1.219 1.223 1.222 157 3.4 Mass Changes The change in mass resulting form hydrothermal corrosion for the Coors aluminas was determined for the given research. Twin specimens of AD-94 were exposed for 100 hours at 250°C within the same hydrothermal environment, resulting in twice the amount of surface area available to attack. The initial mass of the specimens was 0.1949 g and the final mass was 0.2040 g. mf - m,“ Am% = x100 3.4.1 mi Using Equation 3.4.1, where Am°/o is the percentage mass change, m is the initial mass and m, is the final mass, the percentage mass gain was 4.66%. The observed mass gain is consistent with the formation of a hydrated phase. 3.5 Sectioning of the Atomic Force Microscopy Micrographs The raw data of the AFM micrographs was obtained from the DI software. That data was then imported into a Micorsoft Access database and placed into a 3-column format of (x,y,z), where ‘x’ and ‘y’ are the cartesian coordinates of the image, and ‘z’ is the depth or height of the point relative to the average height of the specimen. Each axis is either 256 or 512 entries long depending on the image. Next, the data was imported into a Microsoft Excel spreadsheet. Simultaneously, using another software package, WSxM 1.1 (Nanotec Electronica: Cantoblanco, Madrid.) the maximum specimen depth in meters was determined for each micrograph. The maximum depth in meters was equated to the maximum depth in the ‘z’ axis. The data was sectioned by retrieving the 158 range of data points corresponding to the depths of interest in nanometers from the database and placing the data into a spreadsheet where the data was plotted (Figures 3.5.1-5). In order to interpret, either the sectioned plot or the original image must undergo a pair of transformations: mirroring and a 90° counter- clockwise rotation (these transformations have already been done for every image within this thesis). The reason for the transformations is that the spreadsheet program plots the data in a different manner than the imaging software. The total area present at a given depth range was calculated from the number of data points present within a given depth range and the area corresponding to each pixel. The area was then analyzed with respect to: (i) the midpoint of the depth intervals (Figures 3.5.7—16), (ii) the midpoint of the depth intervals normalized with the average grain size (Figures 3.5.17-28) and (iii) the square root of the hydrothermal corrosion times (Figures 3.5.29-40). Specimens were analyzed prior to formation of the hydrothermal scale. Specimens were also analyzed after hydrothermal corrosion had etched the surface of the specimen. For the ADS-995 specimens the hydrothermal exposure times analyzed were 10, 12.5 and 15 hours. However, as the hydrothermal corrosion progressed faster in the AD-94 specimens the hydrothermal exposure times analyzed were 1, 1.5 and 2 hours. For Figures 3.5, the regression relationships have been determined through various methods: linear, power law, logarithmic, exponential, Gumbel or Cauchy. The regressions are contained within Tables 3.5.1-6, the regressions represent 159 the amount of area present at a given depth “ad” as a function of ‘dmid’ Which represents: the midpoint of the depth intervals, the midpoint of the depth intervals normalized with the average grain size and the square root of the hydrothermal corrosion times. The coefficient of determination (r2) value is a measure of how accurate the regression line depicts the data sets, with the closer the r2 value is to 1 the better the fit. The regression equations and the r2 values for ADS-995 and AD94 specimens are given (Tables 3.5.1-6). For the grain boundary etching as a function of the depth (Figures 3.5.7-16) and as a function of the depth and the average grain size (Figures 3.5.17-28) two types of distributions were determined: Gumbel (Equation 3.5.1) and Cauchy (Equation 3.5.3). The Gumbel distribution (Equation 3.5.1-2) is a doubly exponential distribution, while the Cauchy distribution (Equation 3.5.3-4) is the exponential of a power expression, where “x” is a random variable with constants “01-3,” and “k.“ Gumbel and Cauchy distributions are each a form of extremal-value statistical theory. The Gumbel distribution has been used to analyze the distribution of pit depths [37] while the Cauchy distribution is used to explain that the pit depth increases with increasing surface area. Note the r2 values of the Gumbel and Cauchy functions are independent of grain size (Tables 3.5.1-4). Tables 3.5.2,4 have additional regressions that correspond to Figures 3513,24 that are valid for the shallow depths (-5 to -275 nm). The Cauchy distribution has a better coefficient of determination while the Gumbel distribution is worse when compared to the coefficient of determination relationship near the 160 specimen surface. The Gumbel distribution is a representation of the corrosion depths, while the Cauchy distribution is descriptive of the pit depth distribution. There is more corroded surface present (Figure 3.5.11) closer (1742 A2 from -5 to —25 nm) to the surface of the specimen than in the pits (0.34 A2 from - 85 to-105 nm) as collected in Tables C.1-2. As the hydrothermal attack progresses (Figures 3.5.6.8), there is more corroded surface area present (470.75 A2) at the 10-hour trial than the 15—hour exposure (2021.37 A"). The ADS-995 (Figure 3.5.8) was corroded more at 15-hours (2021.37 A2), than the AD-94 (Figure 3.5.12) was attacked after 1.5 hours of corrosion (1504.78 A2). The ADS-995 (Figure 3.5.8) has a narrow area distribution between 665.70 A2 and 70.04 A2 independent of the grain size (Figure 3.5.19), for the depth of the trial (-10 to -110 nm). While, the AD-94 (Figure 3.5.12) has a wide area distribution between 1742.02 A2 and 0.88 A2, independent of the grain size (Figure 3.5.23), for the depth of the trial (-5 to -105 nm). Thus, the hydrothermal corrosion performs a deep, narrow etch of the AD-94 surface and a wide, shallow etch of the ADS-995 surface. These results correspond to previous work“:10 relating a decrease in strength to an increase in the impurity concentration as hydrothermal corrosion progresses when coupled with the knowledge that a material will fail under the deepest surface flaw. The AD-94 has deeper surface flaws than the ADS-995. The surface pitting of the ADS-995 showed a predominantly power-law relationship with the square root of time whereas the AD-94 had a mostly linear relationship with the square root of time. In both ADS-994 and AD-94, the 161 exponential and the power-law relationships are very similar (Figures; 3.5.29-40). Therefore, the ADS-995 and the AD-94 the nature of the grain boundary etching as a function of the square root of time was probably mixed mode (power, exponential and linear) as the coefficients of determination at the various depths are similar. Inferring that, at different depths a different relationship is dominant. 162 ad = exp(-cxp[—x]) 3.5.1 —c ex [—c d - ] 3.5.2 ad =exp( 1 (1432: Mid ) ad :exp(_x—k) 3.5.3 -k 3.5.4 = C3dmid ad em 0.4329 ) Table 3.5.1: Regressions from the grain boundary etching of ADS-995 attacked at 250°C for 10 to 15 hours as a function of the depth from the specimen surface Conditions Regression Type Regression r2 8 —4.39 10 hours Cauchy-type ad =exp(3'10 dmid ) 0.9814 0.4329 0.45 —0.075d - 10 hours Gumbel-type ad =exp(85 ex“ ""d ) 0.9224 0.4329 -1.0328 12.5 hours Cauchy-type ad = exp(3442'ldmial ) 0.6315 0.4329 .44 - . 2 - 12.5 hours Gumbel-type ad =exp(240 exp“ 00 2”de )) 0.8310 0.4329 -0.8001 15 hours Cauchy-type ad : exp(6860'4dmid ) 0.8197 0.4329 . . 4 - 15 hours Gumbel-type ad =exp(478 76ex%(;:20908 dm'd)) 0.6670 163 Table 3.5.2: Regressions from the grain boundary etching of AD-94 attacked at 250°C for 1 to 2 hours as a function of the depth from the specimen surface Conditions Relession Type Regression r2 1.0 hour Cauchy-type ad = exp( 0:31:19 ) 0.7873 1.0 hour Gumbel-type ad = exp(23254e"1’(‘°'“°5dmid ) 0.9496 0.4329 —1.8859 1.5 hours 310203d . (001) Cauchy-type ad : ex“ 0 4:31: ) 0.9339 1.5 hours _ _ 441.786xp(—0.013ldmid) (001) Gumbel type ad — exp( 0.4329 ) 0.8943 —0.9346 1.5 hours 5314.7d . (012) Cauchy-type ad = exp( 0 4321;! ) 0.8756 15 hours _ _ 247.93 exp(-0.0083dmid) (012) Gumbel type ad — exp( 0.4329 0.9547 —0.9643 14443d . 2.0 hours Cauchy-type ad = exp( 0 4:36; ) 0.8734 2.0 hours Gumbel-type ad = exp(481'756xP(—0'0068dmid )) 0.5996 0.4329 (-5 to -275 nm) Conditions Regression Type Regression r2 -l.6014 1.5 hours 102741d . Cauch - e _ mtd 0.9683 (001) Y WP ad — exp( 0.4329 ) 1.5 hours 528.4exp(—0.0155dmid) Gumbel- e = 0.7626 (001) ‘y" “d “9‘ 0.4329 ’ —0.7329 1.5 hours 2472.4d . Cauch - e _ mid 0.9274 15 hours 239.216xp(—0.0078dmid) Gumbel- e = 0.8940 (012) typ ad em 0.4329 ) 164 Table 3.5.3: Regressions from the grain boundary etching of ADS-995 attacked at 250°C for 10 to 15 hours as a function of the depth from the specimen surface and the average_grain size Conditions Regression Type Regression 12 6-10‘7d‘4-39 10 hours Cauchy-type a d = exp( mzd ) 0.9814 0.4329 . —1 . - 10 hours Gumbel-type ad = exp(850 45 ”1:43;: OSdm‘d )) 0.9224 —1.0328 12.5 hours Cauchy-type ad : exp(l'2151dmid ) 0.6315 0.4329 24 .44 —52.128 - 12.5 hours Gumbel-type ad =exp( 0 ex“ ""d )) 0.8310 0.4329 —0.8001 15 hours Cauchy-type ad z exp(l4'524dmid 0.8197 0.4329 4 . —18.412d - 15 hours Gumbel-type ad =cxp 78 766x:(4329 ""01 ) 0.6670 165 Table 3.5.4: Regressions from the grain boundary etching of AD-94 attacked at 250°C for 1 to 2 hours as a function of the depth from the specimen surface and the average grain size Conditions Regression Type Regression 12 10—9 d-4.419 1.0 hour Cauchy-type ad = exp( 0 £12“; ) 0.7873 2 2 - 2 . ° 1.0 hour Gumbel-type ad =exp( 3 54exl§4$96 ldm’d )) 0.9496 —1.8859 1.5 hours 0.0063d . (001) Cauchy-type a d = exp( 0 437121: ) 0.9339 1.5 hours _ _ 441.788XP(-156-72dmid) (001) Gumbel type ad —cxp( 0.4329 0.8943 —0.9346 1.5 hours 0.8189d . (012) Cauchy-type ad : ex“ 0 4:31:21: ) 0.8756 15 hours _ 247.93cxp(—99.811dmid) (012) Gumbel-type ad —exp( 0.4329 ) 0.9547 —0.9643 2.0 hours Cauchy-type ad = exp(L6824dmid 0.8734 0.4329 2.0 hours Gumbel-type ad =exp(481'75 exP(_81'722dmid )) 0.5996 0.4329 (-5 to -275 nm) Conditions Regression Type Regression r2 -1.6014 1.5 hours 0.0302d . (001) Cauchy-type “cl = exp( 0 4:121: ) 0.9683 2 . — . ' (1631“)w's Gumbel-type ad = exp(5 8 46x93 4138269°2dm'd) 0.7626 -0.7329 1.5 hours 2.5318d . (012) Cauchy-type ad 2 exp( 0 4:12ch ) 0.9274 15 hours 239.216xp(-94.192dmid) Gumbel- e = 0.8940 (012) typ ad cm 0.4329 ) 166 Table 3.5.5: Regressions from the grain boundary etching of ADS-995 attacked at 250°C for 10 to 15 hours Conditions Regression Type Regression 12 RT -10 to —30 nm Power ad = 3430845323 0.0314 RT -10 to —30 nm Exponential ad = 39.962 exP(0-5742dmid) 0.0423 RT —1 0 to -30 nm Linear ad = 326.87dmid — 757.39 0.1674 RT -30 to —50 nm Power ad = 00016.12)?” 0.9999 RT -30 to -50 nm Exponential ad = 0.0152 eXp(2-5464dmid) 0.9986 RT —30 to —50 nm Linear “cl = 334-95dmid - 1026.1 0.9567 RT -50 to -70 nm Power ad = 5 o10“11d;§a968 0.9806 RT —50 to —70 nm Exponential “cl = 10‘8 6XP(6-2461dmid) 0.9718 RT —50 to —70 nm Linear ad = 430.63dmid - 1387.7 0.8924 RT —70 to —90 nm Power ad = 3 010'l4d35a224 0.9800 RT -70 to —90 nm Exponential a d = 3 010-1 1 exp(7.74dm,-d ) 0.971 RT -70 to —90 nm Linear ad = 339.64dmid -1100.8 0.8504 RT -90 to —1 10 nm Power ad = 5 .10‘17 dilifw 0.9931 RT -90 to -110 nm Exponential a d = 2 .10"13 exp(9.0197dm,-d) 0.9875 RT —90 to —1 10 nm Linear ad = 251-53dmid - 820.19 0.7926 167 Table 3.5.6: Regressions from the grain boundary etching of AD-94 attacked at 250°C for 1 to 2 hours Conditions Refission Type Remession r2 RT —5 to -25 nm Power ad = 1515342127402 0.5179 RT —5 to -25 nm Exponential ad = 5954.7 exP(-l.39ld mid) 0.4692 RT -5 to —25 nm Linear ad = —l849.3d,m-d + 3413.8 0.5651 RT -25 to —45 nm Power ad = 606. 1752-14524 0.0274 RT —25 to —45 nm Exponential ad = 773-66Xp(-0-2694dmid) 0.0138 RT -25 to —45 nm Linear ad = -17l.5dm,-d + 805.05 0.0222 RT -45 to —65 nm Power ad =137.45d31-i1f33 0.7301 RT —45 to -65 nm Exponential ad = 21.1876Xp(1-8681dmid) 0.7722 RT -45 to —65 nm Linear ad = 436.69d mid - 310.89 0.7509 RT —65 to —85 nm Power ad = 6.5302dfn2j49 0.9039 RT —65 to —85 nm Exponential ad = 0.00136Xp(8.6248d mid) 0.8734 RT —65 to -85 nm Linear ad = 385.98dmid — 376.12 0.9865 RT —85 to 105 nm Power ad = 056483565 0.8994 RT —85 to 105 nm Exponential a d = 3 010—7 exp(l4,74dmid) 0.8682 RT —85 to 105 nm Linear ad = 329.06de - 330.14 0.9984 168 AD-94 Hydrothen'naly Annealed for 1.5 hours (-5.4nm to -25.4nm) 504 - , . L, . - ~- 4487 r I f \\ 392 . «‘1‘, t" 'M' u 336 - . A ’0' g 280 . $7 224 ; f 1” I ' 168 ‘ I a 112 ‘ a 56 (I / 0 r r .Fh—r—r- 0 112 168 224 280 336 392 448 504 X(nm) Figure 3.5.1: AD-94 hydrothermally attacked for 1.5 hours at 250°C at a depth of —5.4 to -25.4 nm. AD-94 Hydrothermaly Annealed for 1.5 hours (-25.4nm to -45.4nm) 3 3 . l 6 I 0 "1—M .3 .1: 3 i r r r 0 56 112 168 224 280 336 392 448 504 x (nm) Figure 3.5.2: AD-94 hydrothermally attacked for 1.5 hours at 250°C at a depth of —25.4 to —45.4 nm. 169 ADM Hydromermaily Annealed for 1.5 hours (454nm to 65.41!!!) 504 0 . 448 W\ 392 :0 o , .8 00 I I O 336 g 280 5 >- 224 168 112 - 56- W 0 c 0 M r r . . r . . . o 56 112 168 224 280 336 392 449 504 X(nm) Figure 3.5.3: AD-94 hydrothermally attacked for 1.5 hours at 250°C at a depth of —45.4 to —65.4 nm. AD—94 Hydrothermally Annealed for 1.5 hours (-65.4nm to -85.4nm) 504 W 448 s . ‘ 392 , o t s s 336 3 ' ,. .. a e 280 ' 5 >. 224 so: 168 1 s 14" . / 112 - {s 56 - o . W ' 0 P I r I I I I I I 0 56 1 12 168 224 280 336 392 448 504 X (nm) Figure 3.5.4: AD-94 hydrothermally attacked for 1.5 hours at 250°C at a depth of —65.4 to —85.4 nm. 170 AD-94 Hydrothermally Annealed for 1.5 hours (-85.4nm to -105.4nm) 448 W‘ O 392 ‘° . , so 336 ° ' ° .. “ 3! e 280 5 > 224 m I o 168 — 33>" 112 — ’ (i 56 - . . I49 0 I. e I I T I I I I I 0 56 112 168 224 280 336 392 448 504 X (nm) Figure 3.5.5: AD-94 hydrothermally attacked for 1.5 hours at 250°C at a depth of -85.4 to -105.4 nm. Grain Boundary Etching as a Function of Depth for ADS-995 Hydrothermally Annealed at 250 Celsius for 10 hours 1 000 J 100 1 10* Area (A2) 0.1 0 20 40 60 80 100 120 140 Midpoint of Intervals (nm) 0 A08995 10H -002 — Power (A08995 10H -002) - - Export. (ADSQQS 10H -002) Figure 3.5.6: Grain boundary etching as a function of depth for ADS-995 hydrothermally attacked at 250°C for 10 hours 171 Grain Boundary Etching as a Function of Depth for ADS-995 Hydrothermally Annealed at 250 Celsius for 12.5 hours 1000 - A 100 - ‘33; (U 9 < 10 ~ 1 T i i l i j 0 20 40 60 80 100 120 Midpoint of intervals (nm) 1:: ADSQQS 12H5 —Power (ADSQQS 12H5) — - Expon. (AD8995 12H5) Figure 3.5.7: Grain boundary etching as a function of depth for ADS-995 hydrothermally attacked at 250°C for 12.5 hours Grain Boundary Etching as a Function of Depth for ADS-995 Hydrothermally Annealed at 250 Celsius for 15 hours 1000 - A A 100 1 A ----- : ‘25. (U 9 < 10 - 1 i i i i l 0 50 100 150 200 250 Midpoint of intervals (nm) A ADSQQS 15H0 —- Power (#108995 15H0) - - Expoh. (ADSQQS 151-10) Figure 3.5.8: Grain boundary etching as a function of depth for ADS-995 hydrothermally attacked at 250°C for 15 hours 172 Grain Boundary Etching as a Function of Depth for ADS-995 Hydrothermally Annealed at 250 Celsius for 10 - 12.5 hours 1000 . O 100 - a a C, A O D (2‘s, 1:] to 10 - d) E o 1 _ O <> 0 0.1 l i i i i i l 0 20 40 60 80 100 120 140 Midpoint of Intervals (nm) 0 ADSQQS 10H -002 D ADSQQS 12145 Figure 3.5.9: Grain boundary etching as a function of depth for ADS-995 hydrothermally attacked at 250°C for 10-12.5 hours Grain Boundary Etching as a Function of Depth for ADS-995 Hydrothermally Annealed at 250 Celsius for 10 - 15 hours 1000 - A O D. A A A 100 - o D D 4 A A A O D ‘33; 1:3 to 10 ~ 92 < O 1 - <> 0 0 0. 1 i i I i i 0 50 100 150 200 250 Midpoint of Intervals (nm) 0 ADSQQS 10H ~002 1:2 ADSQQ‘S 12H5 A ADSQQS 15H0 Figure 3.5.10: Grain boundary etching as a function of depth for ADS-995 hydrothermally attacked at 250°C for 10-15 hours 173 Grain Boundary Etching as a Function of Depth for AD-94 Hydrothermally Annealed at 250 Celsius for 1 hour 10000 3 1000 - 22‘ 100 - (v (D 5.2 10 - 1 _ .- O 0.1 . , , , T I 0 20 40 60 80 100 120 Midpoint of intervals (nm) <> A094 1D0 -000 — Power (A094 1D0 -000) - - Expon. (AD94 1DO -000) Figure 3.5.11: Grain boundary etching as a function of depth for AD-94 hydrothermally attacked at 250°C for 1 hour Grain Boundary Etching as a Function of Depth for AD-94 Hydrothermally Annealed at 250 Celsius for 1.5 hours 10000 - 1000 ~ ‘2'; 100 - 8 3 10 - 1 _ 0.1 i l T i i 0 100 200 300 400 500 Midpoint of Intervals (nm) a ADQ41DS-001 ~. ADQ41DS -012 —Power(A094 105 -001) - - Expon. (A094 105 -001) Power (A094 105 «012) — - Expon. (A094 105 -O12) Figure 3.5.12: Grain boundary etching as a function of depth for AD-94 hydrothermally attacked at 250°C for 1.5 hours 174 Grain Boundary Etching as a Function of Depth for AD-94 Hydrothermally Annealed at 250 Celsius for 1.5 hours (-5 to -275 nm) 10000 - 1000 - 100 ~ Area (A2) 104 50 100 1 50 200 250 Midpoint of intervals [Average Grain Size Cl A094 105 -001 O A094 105 -012 - - Expon. (A094 105 -001) — Power (AD94 1D5 -001) — Power (A094 1 D5 -012) — — Expon. (A094 105 -012) Figure 3.5.13: Grain boundary etching as a function of depth for AD-94 hydrothermally attacked at 250°C for 1.5 hours (-5 to —275 nm) Grain Boundary Etching as a Function of Depth for AD-94 Hydrothermally Annealed at 250 Celsius for 2 hours 10000 5 1000 - 100 -* Area (A2) A 10- I I I 50 100 150 200 250 300 350 400 Midpoint of Intervals (nm) e A094 200 -001 — Power (AD94 200 -001) - - Expon. (A094 200 -001) Figure 3.5.14: Grain boundary etching as a function of depth for AD-94 hydrothermally attacked at 250°C for 2 hours 175 Grain Boundary Etching as a Function of Depth for AD-94 Hydrothermally Annealed at 250 Celsius for 1 - 1.5 hours 10000 - 1000 — 8 DO 100 — Area (A2) 10- ,3: 1‘ 1:1 0 . 1 I I I I I 0 1 00 200 300 400 500 Midpoint of intervals (nm) 0 A094 100 -000 D A094 105 -001 X A094 105 -012 Figure 3.5.15: Grain boundary etching as a function of depth for AD-94 hydrothermally attacked at 250°C for 1-1.5 hours Grain Boundary Etching as a Function of Depth for AD-94 Hydrothermally Annealed at 250 Celsius for 1 - 2 hours 10000 - - § 1000 8 ‘< A A > 6 & 12$ 100 — D g ,3 a (c U Q 9 D , < 10 - o C 1 7 1:1 0 0. 1 l i i l I 0 100 200 300 400 500 Midpoint of intervals (nm) 0 AD94 1D0 -000 D AD94 1 D5 -001 A AD94 200 -001 >< A094 105 -012 Figure 3.5.16: Grain boundary etching as a function of depth for AD-94 hydrothermally attacked at 250°C for 1-2 hours 176 Grain Boundary Etching as a Function of Depth and Average Grain Size for ADS-995 Hydrothermally Annealed at 250 Celsius for 10 hours 1000 - 100 ~ 10‘ Area (A2) 0.1 I I 0 0.01 0.02 0.03 0.04 0.05 0.06 Midpoint of Intervals / Average Grain Size — Power (AD8995 10H -002) — — Expon. (A05995 10H -002) Figure 3.5.17: Grain boundary etching as a function of depth and average grain size for ADS-995 hydrothermally attacked at 250°C for 10 hours 0 A08995 10H -002 Grain Boundary Etching as 3 chtion of Depth and Average Grain Size for ADS—995 Hydrothermally Annealed at 250 Celsius for 12.5 hours 1000 - 100 ~ Area (A2) 10- I 0 0.01 0.02 0.03 0.04 0.05 Midpoint of intervals I Average Grain Size 13 AD8995 12H5 — Power (ADSQQS 12H5) — - Expon. (A08995 12H5) Figure 3.5.18: Grain boundary etching as a function of depth and average grain size for ADS-995 hydrothermally attacked at 250°C for 12.5 hours 177 Grain Boundary Etching as a Function of Depth and Average Grain Size for ADS-995 Hydrothermafly Annealed at 250 Celsius for 15 hours 1000 — A 100 - ‘33. (U 9 < 10 - 1 I I I I I I 0 0.02 0.04 0.06 0.08 0.1 0.12 Midpoint of Intervals I Average Grain Size A ADSQQS 15H0 — Power (ADSQQS 151-10) — - Expon. (AD8995 15H0) Figure 3.5.19: Grain boundary etching as a function of depth and average grain size for ADS-995 hydrothermally attacked at 250°C for 15 hours Grain Boundary Etching as a Function of Depth and Grain Size for ADS- 995 Hydrothermally Annealed at 250 Celsius for 10 - 12.5 hours 1000 o O 100 - a U B A O C] a 10 - 92 < 0 1 a <> 0 0 0.1 I l I I I 1 0 0.01 0.02 0.03 0.04 0.05 0.06 Midpoint of intervals I Average Grain Size 0 A08995 10H -002 D A08995 12H5 Figure 3.5.20: Grain boundary etching as a function of depth and average grain size for ADS-995 hydrothermally attacked at 250°C for 10-12.5 hours 178 Grain Boundary Etching as a Function of Depth and Average Grain Size for ADS-995 Hydrothermally Annealed at 250 Celsius for 10 - 15 hours 1000 - A O A A A 100- o D D “ A A A A O U ‘35, 13 (U 10 4 9 < O 1 a 0 O O 0.1 I I I T I I 0 0.02 0.04 0.06 0.08 0.1 0.12 Midpoint of intervals I Average Grain Size 0 A08995 10H -002 1:3 A08995 12H5 A A08995 15H0 Figure 3.5.21: Grain boundary etching as a function of depth and average grain size for ADS-995 hydrothermally attacked at 250°C for 10-15 hours Grain Boundary Etching as a Function of Depth and Average Grain Size for AD-94 Hydrotherrnaliy Annealed at 250 Celsius for 1 hour 10000 . 1000 — 9.4: 100 . (U s < 10 - 1 _ 0.1 l i i l l 0 0.002 0.004 0.006 0.008 0.01 Midpoint of intervals I Average Grain Size 0 A094 1D0 -000 — Power (AD94 100 -000) - - Expon. (A094 100 -000) Figure 3.5.2: Grain boundary etching as a function of depth and average grain size for AD-94 hydrothermally attacked at 250°C for 1 hour 179 Grain Boundary Etching as a Function of Depth and Average Grain Size for A094 Hydrothermally Annealed at 250 Celsius for 1.5 hours 10000 - 1000 . § 100 « 79’ a 10 . 1 _ 0.1 i i l i i 0 0.01 0.02 0.03 0.04 0.05 Midpoint of intervals I Average Grain Size :1 A094 105 -001 o AD94 105 -012 -— Power (A094 105 -001i — - Exoon. (ADQ4 1D5 -0011 Figure 3.5.23: Grain boundary etching as a function of depth and average grain size for AD-94 hydrothermally attacked at 250°C for 1.5 hours Grain Boundary Etching as a Function of Depth and Average Grain Size for AD-94 Hydrothennaiiy Annealed at 250 Celsius for 1.5 hours (-5 to -275 nm) 10000 — 1000 a ‘23: «r 100 ~ 8’ < 10 - 1 1 i l I i 0 0.005 0.01 0.015 0.02 0.025 Midpoint of Intervals I Average Grain Size 1: AD94 1D5 -001 o A094 1 D5 -012 — Power (A094 105 -001) - — Expon. (AD94 105 -001) — Power (A094 105 -012) — — Expon. (A094 1D5 -012) Figure 3.5.24: Grain boundary etching as a function of depth and average grain size for AD-94 hydrothermally attacked at 250°C for 1.5 hours (-5 to -275 nm) 180 Grain Bomdary Etching as a Function of Depth and Average Grain Size for AD-94 Hydrothermally Annealed at 250 Celsius for 2 hours 10000 ~ 1000 ~ ‘35. (u 100 . 2 < 10 1 1 i l 17 i F l l 0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 Midpoint of intervals / Average Grain Size A A094 2D0 -001 — Power (AD94 2D0 -001) - — Expon. (A094 200 -001) Figure 3.5.25: Grain boundary etching as a function of depth and average grain size for AD-94 hydrothermally attacked at 250°C for 2 hours Grain Boundary Etching as a Function of Depth and Grain Size for AD-94 Hydrothermally Annealed at 250 Celsius for 1 - 1.5 hours 10000 — _ 8- 1000 < 8 (7%: 100 7 >: 8 1:1 ’0 ,. 8 D g 10 - o 33" 1 7 :1 O 0.1 i i 1 fl 1 0 0.01 0.02 0.03 0.04 0.05 Midpoint of intervals lAverage Grain Size 0 AD94 100 -000 r: AD94 1D5 -001 x A094 1D5 -012 Figure 3.5.26: Grain boundary etching as a function of depth and average grain size for AD-94 hydrotherrnaliy attacked at 250°C for 1-1.5 hours 181 Grain Boundary Etching as a Function of Depth and Average Grain Size for AD-94 Hydrothermaly Annealed at 250 Celsius for 1 - 2 hours 10000 - - 3 1000 8 x A A .1 B. A <25, 100 - D 3 4, a (U D 9 9 [:1 < 10 7 C3 0 1 _ :1 O 0.1 l i l l I 0 0.01 0.02 0.03 0.04 0.05 Midpoint of intervals / Average Grain Size 0 A094 100 -000 D A094 105 -001 A A094 200 -001 x A094 105 -012 Figure 3.5.27: Grain boundary etching as a function of depth and average grain size for AD-94 hydrotherrnaiiy attacked at 250°C for 1-2 hours Grain Boundary Etching as a Function of Depth and Average Grain Size for ADS-995 (10-15 hours) and AD-94 (1-2 hours) Hydrothermally Annealed at 250 Celsius 10000 - 1000 - + A $5.3“ X + + + + it, 100 — BEE; 4 :2 ¢ L A o + + to 0 ~ 9 10 - D >: O < o x D 1 7 X0 0 x >1: 0.1 I f I I I I 0 0.02 0.04 0.06 0.08 0.1 0.12 Midpoint of intervals I Average Grain Size 0 A094 100 -000 D A094 105 -001 A A094 200 -001 X A08995 10H -002 O A08995 12H5 + A05995 15H0 Figure 3.5.28: Grain boundary etching as a function of depth and average grain size for ADS-995 (10-15 hours) and AD-94 (1 -2 hours) hydrothermally attacked at 250°C 8 A094105 -012 182 Grain Boundary Etching as a Function of Root Time for ADS-995 Hydrothermally Annealed at 250 Celsius for 10 - 15 hours (-10 to -30 nm) 0 I T I I I 3 3.2 3.4 3.6 3.8 4 Root Time 0 -10 to -30nm — Power (-10 to -30nm) - — Expon. (-10 to -30nm) - ° - - Linear (-10 to -30nm) Figure 3.5.29: Grain boundary etching as a function of root time for ADS-995 hydrothermally attacked at 250°C for 10-15 hours (-10 to -30 nm) Grain Boundary Etching as a Function of Root Time for ADS-995 Hydrothermally Annealed at 250 Celsius for 10 - 15 hours (-30 to -50 nm) o I I I I I 3 3.2 3.4 3.6 3.8 4 Root Time 0 ~30 to -50nm — Power (-30 to -50nm) - - Expon. (-30 to -50nm) - - - - Linear (-30 to -50nm) Figure 3.5.30: Grain boundary etching as a function of root time for ADS-995 hydrothermally attacked at 250°C for 10-15 hours (-30 to -50 nm) 183 Grain Boundary Etching as a Function of Root Time for ADS-995 Hydrothermally Annealed at 250 Celsius for 10 - 15 hours (-50 to -70 nm) 450 a 400 - 350 - 300 - g 250 1 a 200 9 < 150 a 100 - 50 - 0 .- . -50 3 3.2 3.4 3.6 3.8 4 Root Time 0 -50 to -70nm — Power (-50 to -70nm) - - Expon. (-50 to -70nm) - - - - Linear (-50 to -70nm) Figure 3.5.31: Grain boundary etching as a function of root time for ADS-995 hydrothermally attacked at 250°C for 10—1 5 hours (-50 to —70 nm) Grain Boundary Etching as a Function of Root Time for ADS-995 Hydrothermally Annealed at 250 Celsius for 10 - 15 hours (-70 to -90 nm) 350 1 3001 2507 200 a 1501 100 - _ 50 - 0 a L, , i J -50 3 Area (A2) O.) N 0) p. 9° 0) 00 00 .5 Root Time A -70 to -90nm — Power (-70 to -90nm) - — Expon. (-70 to -90nm) - - - - Linear (-70 to -90nm) Figure 3.5.32: Grain boundary etching as a function of root time for ADS-995 hydrothermally attacked at 250°C for 10-15 hours (-70 to -90 nm) 184 d Grain Boundary Etching as a Function of Root Time for ADS-995 Hydrothermally Annealed at 250 Celsius for 10 - 15 hours (-90 to -110 nm) 50 ~ 0 * I o t - - I I I I is ' '3' 2 3.4 3 6 3 8 4 -50 - Root Time it -90 to -1 10nm — Power (-90 to -110nm) - — Expon. (-90 to -110nm) - - - -Linear (-90 to -110nm) Figure 3.5.3: Grain boundary etching as a function of root time for ADS-995 hydrothermally attacked at 250°C for 10—15 hours (-90 to -110 nm) Grain Boundary Etching as a Function of Root Time for ADS-995 Hydrothermally Annealed at 250 Celsius for 10 - 15 hours (-10 to -110nm) 1000 - D D S 100 a Q A O A ‘25; x (U 10 - (D 2 :1 1 d A )K 0.1 I I I I 1 3 3.2 3.4 3.6 3.8 4 Root Time 0 -10 to ~30nm o -30 to -50nm c1 -50 to -70nm A -70 to -90nm x -90 to -110nm Figure 3.5.34: Grain boundary etching as a function of root time for ADS-995 hydrothermally attacked at 250°C for 10-15 hours (-10 to -110 nm) 185 Grain Boundary Etching as a chtion of Root Time for AD-94 Hydrothermally Annealed at 250 Celsius for 1 - 2 hours (-5 to -25 nm) 2000 - 1800 a U 1600 ~ 1400 - 0 j I I I I 0.95 1.05 1.15 1.25 1.35 1.45 Root Time U -5 to -25 nm — Power (-5 to -25 nm) — — Expon. (-5 to -25 nm) - - - -Linear (-5 to -25 nm) Figure 3.5.35: Grain boundary etching as a function of root time for AD-94 hydrotherrnaily attacked at 250°C for 1-2 hours (-5 to —25 nm) Grain Boundary Etching as a Function of Root Time for AD-94 Hydrothermally Annealed at 250 Celsius for 1 - 2 hours (-25 to —45 nm) 800 - 700 - 0 600 — :- l :— L'_'..'_' ------------------------------------ 0 I I I I I 0.95 1.05 1.15 1.25 1.35 1.45 Root Time o -25 to -45 nm — Power (-25 to -45 nm) - - Expon. (-25 to -45 nm) - - - - Linear (-25 to -45 nm) Figure 3.5.36: Grain boundary etching as a function of root time for AD-94 hydrothermally attacked at 250°C for 1-2 hours (-25 to -45 nm) 186 Grain Boundary Etching as a Function of Root Time for AD-94 Hydrothermally Annealed at 250 Celsius for 1 - 2 hours (-45 to -65 nm) 0 I I I I 0.95 1.05 1.15 1.25 1.35 1.45 Root Time U -45 to -65 nm -——- Power (-45 to -65 nm) - - Expon. (-45 to -65 nm) - - - - Linear (-45 to -65 nm) Figure 3.5.37: Grain boundary etching as a function of root time for AD-94 hydrothermally attacked at 250°C for 1-2 hours (-45 to —65 nm) Grain Boundary Etching as a Function of Root Time for AD—94 Hydrothermally Annealed at 250 Celsius for 1 - 2 hours (-65 to -85 nm) 300 - 250 - 200 A 150 1 100 - 50 - 0 R-—- 0.95 1.05 1.15 1.25 1.35 1.45 Root Time Area (A2) A -65 to -85 nm — Power (-65 to -85 nm) - — Expon. (-65 to -85 nm) - - - - Linear (-65 to -85 nm) Figure 3.5.38: Grain boundary etching as a function of root time for AD-94 hydrothermally attacked at 250°C for 1-2 hours (-65 to -85 nm) 187 Grain Boundary Etching as a Function of Root Time for AD~94 Hydrothermally Annealed at 250 Celsius for 1 ~ 2 hours (~85 to ~105 nm) 350 . 300 1 250 - 200 4 150 7 100 - 50 - 0 x - . - . I. I I I 1 500335 1.05 1.15 1.25 1.35 1.45 Area (A2) Root Time 1: ~85 to ~105 nm — Power (~85 to ~105 nm) - - Expon. (~85 to ~105 nm) - - - - Linear (~85 to ~105 nm) Figure 3.5.39: Grain boundary etching as a function of root time for AD-94 hydrothermally attacked at 250°C for 1-2 hours (~85 to —105 nm) Grain Boundary Etching as a Function of Root Time for AD-94 Hydrothermally Annealed at 250 Celsius for 1 ~ 2 hours (~5 to ~105 nm) 2000 — 1800 - D 1600 — 1400 - ‘22: 1200 - E 1000 — a < ) 800 ‘i o D o 600 ' 400 - O 200 ‘ 1:) g A 0 ‘ I I fl I 0.95 1.05 1.15 1.25 1.35 1.45 Root Time D ~5 to ~25 nm 0 ~25 to ~45 nm D ~45 to ~65 nm A ~65 to ~85 nm :1: ~85 to ~105 nm Figure 3.5.40: Grain boundary etching as a function of root time for AD~94 hydrothermally attacked at 250°C for 1-2 hours (~5 to 105 nm) 188 3.6 Electron Dispersive Study The electron dispersive study (EDS) for both the ADS-995 and the AD-94 was performed at 20 W with a 17 mm working distance, exposure times were conducted in real time for 10 minutes for each specimens with representative images captured in Figures 3.6. The EDS 0f the ADS-995 and AD-94 specimens hydrothermally attacked for 0, and 5 hours displayed no appreciable quantities of elements besides aluminum. Since neither material is 100% pure and therefore contains some impurities, the impurity levels of the individual constituents are believed to be below the level, which is detectable by the available EDS array. The gold, which was found on the surface of the ADS-995 specimens, is due to the gold coating used for the analysis. This gold coating may be inhibiting the detection of the characteristic x-rays from the various impurity elements. 189 Figure 3.6.1: EDS analysis of Coors ADS-995 with no hydrothermal attack. Figure 3.6.2: EDS analysis of Coors AD-94 with no hydrothermal attack. Figure 3.6.3: EDS analysis of Coors ADS-995 hydrothermally attacked for 5 hours at 250°C. Figure 3.6.4: EDSanalysis of Coors AD-94 hydrothermally attacked for 5 hours at 250°C. 191 3.7 Scanning Electron Microscope Analysis Scanning electron microscopy (SEM) was performed for ADS-995 and AD~ 94 for the O and 5~h0ur specimens. The SEM images were imaged with a 25 kV beam and a working distance of 10 mm. The low magnification SEM images of non-hydrothermally attacked ADS-995 (Figures 3.7.1 a-b) show a surface with many (<50) small (<3 um) surface depressions (Figure 3.7.1 a). The backscattered electron (BSE) image (Figure 3.7.1 b) reveals many pores not apparent in the SEM image (Figure 3.7.1a). At higher magnifications (Figures 3.7.1c-d), the pits are still visible in the SEM image (Figure 3.7.10) and additional flaws appear in the BSE image (Figure 3.7.1 d). Non-hydrothermally attacked AD-94 (Figure 3.7.2a,c) has fewer (<10) surface depressions that are larger (<15 pm) than the pits in the ADS-995 (Figure 3.7.1a). The corresponding BSE images (Figure 3.7.1b,d), reveals more defects than the SEM image, which was also the case in the ADS-995 (Figures 3.7.1 b,d). Following 5 hours of hydrothermal corrosion, the ADS-995 surface (Figure 3.7.3) displays more surface flaws than the ADS-995 that had not been hydrothermally attacked (Figure 3.7.1). At higher magnification (Figure 3.7.3c,d), the surface has been textured by the hydrothermal corrosion. A 5-h0ur hydrothermal corrosion of AD-94 (Figures 3.7.4) has more (>30) surface flaws, than, the non-hydrothermally attacked AD~94 specimen (Figure 3.7.2). Additionally, a surface deposit has collected on the substrate. 192 I Ilium I; . _ ' Figure 3.7.1: Scanning electron image of Coors ADS-995, with no hydrothermal treatment, polished to a 1 pm finish; a) Low magnification SEM image, b) Low magnification BSE image, c) High magnification SEM image, d) High magnification BSE image. 193 Figure 3.7.1: Continued. 194 w I ”“1".“ , ,. Figure 3.7.2: Scanning electron image of Coors AD-94, with no hydrothermal treatment, polished to a 1 pm finish; a) Low magnification SEM image, b) Low magnification BSE image, 0) High magnification SEM image, d) High magnification BSE image. 195 Figure 3.7.2: Continued. 196 Figure 3.7.3: Scanning electron image of Coors ADS-995, hydrothermally attacked for 5 hours at 250°C; a) Low magnification SEM image, b) Low magnification BSE image, c) High magnification SEM image, d) High magnification BSE image. 197 Figure 3.7.3: Continued. 198 b Figure 3.7.4: Scanning electron image of Coors AD-94, hydrothermally attacked for 5 hours at 250°C; a) Low magnification SEM image, b) Low magnification BSE image, 0) High magnification SEM image, d) High magnification BSE image. 199 : Continued. Figure 3.7.4 200 3.8 Surface Morphology For AD~94 and ADS-995 the Figures 3.1 show variations in the surface morphology as a function of hydrotherrnai corrosion time. initially the surface is devoid of any scale but as the hydrothermal corrosion progresses the scale deposits on the surface. The majority of the surface scale (Figures 3.1) that was deposited on the substrates (AD-94 and ADS-995) displayed a mixed morphology of {111} and {110} planes, which appears as a cauliflower-type structure (Figure. 3.1.16b and Figure 3.1.28b) A few structures showed the {100} morphology. The differing morphologies are similar to the synthesis of diamond films though microwave discharge [38]. Depending on the temperature and the atmospheric ratios of methane and hydrogen diamond films will develop differing surface morphologies appearing as: cauliflower-type structures at low temperatures and low to high ratios of methane to hydrogen, {111}-type facets at high temperatures and low to high ratios with an intermediate {100}-type facet developing at moderate temperatures and ratios. The surface film deposited on the AD-94 and ADS-995 experiences various morphologies depending on the length of time they were subjected to the hydrothermal corrosion. A cauliflower-type structure was evident in both AD-94 (Figure 3.1.16b) and ADS-995 (Figure 3.1.28b) while different specimens AD-94 (Figure 3.1.120) and ADS-995 (Figure 3.1.27b) showed {100}-type facet. 201 3.9 Summary and Conclusions ADS-995 and AD-94 structural alumina ceramics were hydrothermally corroded for up to 50 hours at 250°C. The ADS-995 was hydrothermally corroded for 0, V2, 1, 5, 10, 12%, 15, 20, 25, 30, 35, 40, 45, and 50 hours with times of interest being 0, 5, 10-15 and 50 hours of corrosion. The AD-94 was hydrothermally corroded in half hour increments between 0 and 7.5 hours, 10 and 50 hours with times of interest being 0, 1-2, 5, 10 and 50 hours. The given periods are of interest because they correspond to the control specimens (0 hours), the most heavily corroded specimens for this study (50 hours), the ranges where the hydrothermal corrosion went from surface pitting to the deposition of a surface layer and reference times of 5 and 10 hours. For the hydrothemial corrosion of the present study, the entire surface is simultaneously being corroded by the exact same conditions. The surface uniformity is best expressed in Figure 3.7.4a-d, where the long-range order of the corrosion may be seen. Viewing the surface at lower magnification (Figures 3.7) shows the difficulty in locating and imaging a 100 - 10000 um2 area that was nominally flat. The hydrothermal environment initially has little to no effect on the alumina substrates (Figures 3.1.3-4). As the material is subjected to the hydrothermal environment, for longer periods the surface begins to be corroded (Figures 3.1.5- 6) evidenced as pitting and etching of the grain boundaries. Eventually the grains will be discernable form a once finely polished substrate (Figures 3.1.8, 25). The etching process is progressing in a closed environment so the etched 202 material, which comes from the grain boundary phases, is placed into the water solution. As the time in the hydrothermal environment proceeds the compounds in the solution precipitate out onto the substrate surface (Figures 3.1.9, 26) as evidenced through the mass gain (4.66%) experienced by the AD-94 specimens following 100 hours of treatment. Furthermore, the growth of the precipitate phase through AFM at different stages of evolution is depicted in Figures 3.1, with an electron image depiction in Figures 3.7. The X-ray and EDS analyses of the specimens were inconclusive. The X- ray peaks present in the non-hydrothennaliy attacked specimens and the hydrothermally attacked specimens could not be distinguished from one another. Meanwhile the EDS only detected alumina in the specimens. Both the E08 and the X-ray analyses should have detected changes in the specimen compositions as there was a 4.66% mass gain experienced by the AD~94 specimens and the evident growth of a surface layer on the specimens as evidenced in Figures 3.1 and Figures 3.7. Additionally, with the overall roughness data of the alumina specimens may be approximated by a series of normal distributions. However, analysis of the etched areas as function of the pit depth will require extremal-value statistics: Gumbel or Cauchy. Sections 20 nm thick, were analyzed to the maximum pit depth of the specimens (Figures3.5). ADS-995 and AD-94 specimens had similar relationships where the hydrothermal corrosion progresses from etching to depositing a surface layer (Tables 3.51 ~4). As time progresses the ADS-995 203 and AD-94 specimens surface corrosion has a mixed-mode relationship with square root of time. The surface of the alumina progressively becomes more textured as corrosion time increases. Texturing of some form is expected, as surface degradation is the purpose of an etching process. What is of interest is that the surface corrosion yields a surface that appears as many overlapping scales. The scales are composed of polyhedra in the form of {100}, {110} and {111} with the majority of the scales having a mixed {111} and {110} morphology. These polyhedra occlude increasingly more of the surface as the corrosion progresses and eventually may transform into the cube structures observed by Day in 96% pure AlSiMag 614 (Section 1.8) or diaspore and boehmite mentioned by Sinharoy et. ai. (Section 1.7). Although boehmite, diaspore and wairakite were not found in the current study, a scale did develop on the specimen surfaces. The hydrothermal environment degrades the surface of polished aluminas (ADS-995 and AD-94) at temperatures well below the maximum atmospheric use temperature (Table 2.1.1) of the ceramic. Thus, there should be a maximum use temperature for the hydrothermal environment. The progression of this attack follows extremal-value statistics, which have already been applied to failure analysis. Understanding, the mechanics and properties of hydrothermal corrosion, is fundamental in determining the service life or schedule necessary to prevent failure. 204 3.10 Future Work Future work in the hydrothermal corrosion of ceramics should encompass an in-depth X-ray analysis of the ADS-995 and AD~94 specimens hydrothermally corroded for 0 to 50 hours. Furthermore, an analysis of the scale development from 10 to 15 hours in ADS-995 and from 1 to 2 hours in AD-94 in 10-minute intervals (Le. 60, 70, 80 minutes for AD-94) should be completed. in the current work an analysis was made of the corroded surface area present at a given depth (Section 3.5), a study of the total pit should also be conducted. Finally, the exact same study conducted in the present work should be done for a non-oxide ceramic to compare how different material systems. 205 REFERENCES 206 REFERENCES . William D. Callister, “Materials Science and Engineering: An introduction,” Wiley, New York, pages 70-159, 2000. . W. D. Kingery, H. K. Bowen and D. R. Uhimann, “introduction to Ceramics,” Wiley, New York, pages 66-70, 188-201, 714, 1976. . Michel Barsoum, “Fundamentals of Ceramics,” McGraw-Hill, New York, pages 365-503, 1997. . J. F. Lynch, Ruderer C. G., and Duckworth W. H., “Engineering Properties of Selected Ceramic Materials,” American Ceramic Society incorporated, Columbus, pages 541-28, 1966. . Kohei Oda and Tetsuo Yoshio, “Hydrothermal Corrosion of Alumina Ceramics,” Journal of the American Ceramic Society, Volume 80, Number 12, pages 3222-3236. . S. Kiotaoka, Y. Yamaguchi and Y. Takahashi, “Tribological characteristics of a—Alumina in High-Temperature Water,” Journal of the American Ceramic Society, volume 75,number 11, pages 3075-80, 1992. . S. Ono, K. Suzuki, M. Kumagai, and K. Hoshi, “Corrosion and Elution behaviors in Al203 Ceramics in High-Temperature Demineralized Water,” Ziryo-to-Kankyo, volume 40, pages 655-60, 1991. . David Avigdor and S. D. Brown, “Delayed failure of a porous alumina,” Journal of the American Ceramic Society, volume 61, number 3-4, pages 97- 99, 1978. . Delbert E. Day, “Reaction of alumina ceramics with saturated steam,” Ceramic Bulletin, volume 62, number 6, pages 624-631, 1982. 10. Samar Sinharoy, Levenson L., Ballard W. and Day D. “Surface segregation of calcium in dense alumina to steam and steam-CO,” Ceramic Bulletin, volume 57 number 2, pages 231-233, 1978. 11.Fi. M. Barrer and P. J. Denny, Hydrothermal Chemistry of the Silicates: X, A Partial Study of the Field CaOoAl20304Si0202H20,” Journal of the Chemical Society (London), pages 983-1000, 1961. 12.0. S. Coombs, “X-Fiay Observations on Wairakite and Non-Cubic Analcine,” Mineral Magazine, volume 30, number 230, pages 699-708, 1955. 207 13. A. Steiner, ”Wairakite, the Calcium Analogue of Analcine,” Mineral Magazine, volume 30, number 230, pages 691-98, 1955. 14. D. Bahat, “Compositional Study and Properties of Characterization of Alkaline Earth Feldspar Glasses and Glass-Ceramics,” Journal of Materials Science, volume 4, number 10, pages 855-60, 1969. 15. R. l. Taylor, J. P. Coad and R. J. Brook, “Grain boundary segregation in A1203," Journal of the American Ceramic Society, volume 57, number 12, pages 539-540, 1974. 16. H. L. Marcus and M. E. Fine, “Grain boundary segregation in MgO-doped Al203,” Journal of the American Ceramic Society, volume 55, number 11, pages 568- 570, 1972. 17. F. A. Kroeger, “Chemistry of imperfect Crystals,” North-Holland Publishing Company, Amsterdam, Page 776, 1964. 18.W. D. Kingery, “Plausible Concepts Necessary and Sufficient for interpretation of Ceramic Grain Boundary Phenomena: 1” Journal of the American Ceramic Society, volume 57, number 1, pages 1-12, 1974. 19. P. P. Budnikov and F. l. Chatikonov, “influence of High Pressure Water Vapor on Some Properties of Corundum Materials,” Silikattechnik, volume 18, number 8, pages 243-48, 1967. 20. Hideo Hirayama, T. Kawakubo, A. Goto and T. Kaneko, “Corrosion behavior of silicon carbide in 290°C water,” Journal of the American Ceramic Society, Volume 72, number 11, pages 2049-2053, 1989. 21 .Yu Gogotsi and M. Yoshimura, “ Degradation of SiC-based fibres in high- temperature, high-pressure water,” Joumai of Materials Science, volume 13, number 6, pages 395-399, 1994. 22.0. J. Pysher, K. C. Goretta, R. S. Hodder and R. E. Tressier, “Strengths of Ceramic Fibers at Elevated-Temperatures,” Journal of the American Ceramic Society, volume 72, number 2, pages 284 - 288, 1989. 23. R. E. Tressier, N.Y. Jia, J. Zheng and H. Takahashi, in Proceedings of the tst international Symposium on the Science of Engineering Ceramics by S. Kimura and K.Niihara, page 167, 1991. 24.T. Yamamura, T. lshikawa, M. Shibuya, T. Nagasawa and K. Okamura, in Proceedings of the 1st Japan international SAMPE Symposium, page 1084, 1 989. 208 25.Yu G. Gogotsi and V. A. Lavrenko, “Corrosion of high-performance ceramics,” Springer Veriag, Berlin, pages 40-45, 1992. 26.Ciaude Contet, Kase J. l., Noma T., Yoshimura M., and Shigeyuki Somiya, “Hydrothermal oxidation of Si3N4 powder,” Joumai of Materials Science Letters, volume 6, pages 963-964, 1987. 27.S.C Singhal, Journal of American Ceramic Society, volume 59, number 1-2, pages 81-82, 1976. 28.Yung-Kuo Chao and D. E. Clark, “Hydrothermal corrosion of lithium disilicate glass and glass ceramics,” Glass Technology, Volume 25, number 3, pages 152-156, 1984. 29. BE. Clark and EC. Etheridge, American Ceramic Society Bulletin, volume 60, number 6, page 646, 1981. 30. R.J. Charles, Journal of Applied Physics, Volume 29, Number 11, page 1549, 1958. 31 .Coors Technical Ceramics Website, www.coorstekcom, 2001. 32. Parr instrument Company, “Operating Instructions: Parr Acid Digestion Bombs,” No. 249M, Parr instrument Company, December 1997. 33. Digital instruments Nanoscope iii: Atomic Force Microscope intruction Manual Version 2.2, July 8, 1992. 34. Digital instruments Website, www.c_li.com, 2001. 35.JCPDS--international Centre for Diffraction Data, “Powder diffraction file: Sets 1-43,” American Society for Testing and Materials. Swarthmore, Pa.: The Centre, 1960. 36. 8D. Culiity, “Elements of X-ray Diffraction: Second Edition,” Addison-Wesley Publishing Company, London, page 295, 1978. 37.T. Shibata, “1996 W.R. Whitney Award Lecture: Statistical and Stochastic Approaches to Localized Corrosion,” Corrosion, volume 52, number 11, pages 813-830, 1996. 38.Saeid Khatami, “Controlled synthesis of diamond films using a microwave discharge (non-equilibrium plasma),” Michigan State University, Thesis, pages 339-455, 1997. 209 39.John Frederick Nye, “Physical Properties of Crystals: Their Representation by Tensors and Matrices,” Clarendon Press, Oxford, pages 141-281, 1984. 40.J. B. Wachtman Jr., “Elastic Deformation of Ceramics and Other Refractory Materials,” pages 139 to 168, in Mechanical and Thermal Properties of Ceramics: Proceedings of a Symposium, Edited by JB Wachtman Jr., US Government Printing Office, Washington DC, page 146, 1969. 210 APPENDIX 211 APPENDIX A: Slip Systems for Alumina Optimaliy there are six ({10—11}o<11~20>) independent slip systems in a hexagonal (a = b at c, a = B = 90°, y=120°) close-packed structure, although most hexagonal close-packed structures have only three ({10-10}o<11-20> or {0001}-<11-20>) independent slip systems [1]. In the case of alumina, there are only two independent slip systems {0001}-<11-20> [2]. The structure of alumina is not close-packed but a hexagonal close-packed sub-lattice with two-thirds of the octahedral sites filled as the other sub-lattice [2]. Normally three independent slip systems are associated with the {0001} slip plane, alumina has “lost” an independent slip system. Unfortunately, the hexagonal matrix is a rhombohedral system [39]. The other rhombohedral system is trigonal (a = b at c, or = B = 90°, y=120°) [39]. Both of these crystallographic systems use the four coordinate Miller-Bravais indices {hkil}. if the material in question is not pure then the hexagonal close-packed lattice becomes a hexagonal sub-lattice, which in some cases may be represented as a trigonal structure. Alumina crystals are capable of being represented as trigonal or hexagonal [40]. Crystallographicaily, the major difference between the trigonal and hexagonal systems is that the trigonal structure has a single, 3-foid axis of rotation, whereas the hexagonal structure has a single, 6-fold axis of rotation [39]. in both cases, the axis of rotation is parallel to the z-axis [39]. As a result, most alumina data is recorded in the hexagonal system but some is reported in the trigonal system. The system in which the data was 212 recorded determined how the specific analysis was performed. We will treat the hexagonal system as the norm for alumina and we will note any data based on the trigonal system. 213 APPENDIX B: Elastic Moduli of Alumina Much research was done to obtain the anisotropic elastic moduli of alumina crystals. The single crystal elastic constants, compliance and stiffness coefficients have been reported in Table B1 in the trigonal system [40]. The compliance data was then placed within the symmetric elasticity matrix for a trigonal body, Table 8.2. The compliances may then be placed within the appropriate expression (Equation B.1) for the reciprocal of the elastic modulus for a trigonal structure of the appropriate class [39]. The compliances have been denoted as s,,- and the 1 values refer to the value of the respective Miller indices. Given the trigonal lattice parameters (a = b at c), the Miller-Bravais indices of interest are [0001] and either [1010] or [0110], where the boldface denotes a negative index. For use in equation 8.1 the relative indices are: [001] and either [100] or [010]. In for the non-basal direction, [001], equation 1.20.1 has been reduced into 3.2 using the appropriate values for l. The resultant elastic modulus in the non-basal direction was 460.83 GPa. Furthermore, 33 is the reduction of 8.1 for either [100] or [010] basal directions. As these directions are equivalent, their elastic modulus was 425.53 GPa. 214 Table 8.1: Single crystal elastic constants of alumina in the monal system [40] 1.1 Q . 514; B E .1_4 Units Compliance 2.35 2.17 6.94 -0.716 -0.364 0.489 10'” m2/N Stiffness 49.7 49.8 14.7 16.4 11.1 -2.35 1010 N/m2 Table 3.2: The reduced trigonal elasticity matrix representing crystallographic classes 32, -3m and 3m, using compliance constants the matrix is symmetric about the main diagonal [39] S11 S12 S13 S14 0 0 S11 523 " S14 0 0 $33 0 0 0 S44 0 0 S44 '2(S14) 2(11-12) — l B ‘l (1 —1_,2)2 s” +13%33 + 1,2 (1 —132 )(2513 + S“) + 21,1,(31,2 -1§s,,) ° 1 E = 4 8.2 ’3 533 1 B3 — (1—z§)2s,, 215 APPENDIX C: Sectioning Data Table C.1: Area present for the sectioning data of ADS-995 A08995 ADS995 ADS995 Depth Range 10H -002 12H5 15H0 (nm) (A2) (A2) (A2) -10.20 to -30.20 418.32 99.14 665.70 -30.20 to -50.20 47.00 128.90 286.52 -50.20 to -70.20 3.78 74.04 312.92 -70.20 to -90.20 1.03 41.58 245.36 -90.20 to -110.20 0.31 16.33 181.73 -110.20 to -130.20 0.31 117.00 -160.20 to -180.20 70.04 -210.20 to -230.20 142.10 Root time (hoursy“) 3.16 3.54 3.87 Table 02: Area present for the sectioninqdata of AD-94 AD94 AD94 A094 A094 Depth Range 100 -000 105 -001 105 -012 200 -001 (nm) (A2) (A2) (A2) (A2) -5.4 to -25.4 1742.02 1202.93 319.02 1009.06 -25.4 to -45.4 758.17 501.02 144.12 710.33 -45.4 to -65.4 153.31 186.12 141.49 339.32 -65.4 to -85.4 4.92 66.99 147.82 163.88 -85.4 to -105.4 0.34 53.33 86.17 136.91 -155.4 to -175.4 29.53 61.80 83.31 -255.4 to -275.4 16.67 33.80 44.29 -355.4 to -375.4 6.94 10.95 96.44 -455.4 to -475.4 0.88 Root time Lhoursy”) 1.0000 1.2247 1.41 216