,. .u... 33%... V n... w... 37.... Fm. a 1.... 2 a «5.36.3.3... 3.3.2... 5 53”. 3&5..me mm. 1 ¢ .. .. E... v. . 3 3:23.... ! x... 4 .. 3.. .32.... . Z .. . .9 J . . .5 J u: , ‘1 t... triafr An. 3.56”: .33 . a 3\. 3. .11. x . ”I... 1 31?. . ‘33.. 1):...” “an... .42. 3.... 1...; V: 5%...» 3:1,] 2 v33... 3 l:.1.}.:1 t I: 1 35.4.14 . 3 .3 111.31...» .3 z 3} Sift “3* J J: -’ t ‘53} 5., I :9: E 3;}: f 3 "mists” ‘ :48 51?: 3. 1.319 I, 39.2.29, 1%.... I .32..» {if}... 1‘ 73.. 3.. 1.. K.- ..|. 3r . .13... so» I... . 3. Q2. .9. 31.0.3. 3 A 3!!“ 4‘! *3 :19 0 3 3.33.33...“ 32 1.10.3.3! 1...; ?. ii. Iolrlrrc 33. 3531. 3! 3)... II.- ‘3.93~Ill\l 3) 3.3 v v. I is, 1 :25; . fa. ... ,...§.n._...b.3..%n..wu..-.u.. win :3... 1.1.1.»..513 . :vnauuumovh .4»! A . - fig? .39.... “tribal. I. h). .ulubl - I .iniL lllllllllllllllllHIIIHHIIIIHIHHIllllllllllllllllllllllll 31293 01410 7282 This is to certify that the thesis entitled The Characterization of Thermal Fatigue Effects by the Observation of Growth of Indentation Cracks in Unreinforced Polycrystalline A1203 presented by Won-Sik Kim has been accepted towards fulfillment of the requirements for Master's degree in Materials Science éécéovxbtcb/UL Major professor Date 9/11/95 0—7639 MS U is an Affirmative Action/Equal Opportunity Institution ‘ v» - - — _ ___._—-4_—~———-——-——‘.—4—. '__.—__—_.———w WT‘ __—-—.—-._——_.—.__ ~__._——.._..-..-_. -fiM LIBRARY Michigan State University ~“-M'O.; PLACE DI RETURN BOXtoromavothb chockwtfrom you'rooord. TO AVOID FINES mum on or baton date due. 1“ DATE DUE DATE DUE DATE DUE THE CHARACTERIZATION OF THERMAL FATIGUE EFFECTS BY THE OBSERVATION OF GROWTH OF INDENTATION CRACKS IN UNREINFORCED POLYCRYSTALLINE Al203 By Won-Sik Kim A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Materials Science and Mechanics August 1995 ABSTRACT THE CHARACTERIZATION OF THERMAL FATIGUE EFFECTS BY THE OBSERVATION OF GROWTH OF INDENTATION CRACKS IN UNREINFORCED POLYCRYSTALLINE A1203 By Won-Sik Kim In this study, we characterized the thermal fatigue damage for unreinforced polycrystalline alumina specimens by measuring the cyclic thermal shock induced growth of Vickers indentation cracks placed in the specimens prior to thermal shock. The severity of thermal shock damage increased with increasing quench temperature difference. A minor amount of crack growth occurred after the introduction of the Vickers indentation cracks prior to commencing the thermal fatigue testing. The crack initially grew an average of 4.3 percent but crack growth saturated within 30 minutes. Grain bridging during thermal fatigue was investigated using a scanning electron microscope (SEM). Although grain bridging is an often-cited mechanism for energy dissipation for cracks growing in unreinforced alumina, grain bridging was infrequent in the specimens included in this study. Thus grain bridging may not account for the saturation in the thermal fatigue crack length observation in this study. iii Acknowledgment I would like to thank my adviser, Dr. Eldon case for his advice and editing this thesis throughout my study. I also thank my fellow researchers Brett Wilson, Kiyong Lee, and Changlin Liu for their help during the my research. I would also thank my parents for their support during my study. I would like express my special thanks to my wife, Jeongioo Moon for her help and patience. Table of Contents Page List of Tables viii List of Figures xii Section 1 Introduction 1 1.1. Thermal shock resistance parameter 2 1.1.1. Thermal shock resistance without crack initiation 2 1.1.2. Thermal shock resistance related to surface flaws 4 1.2. Comparing the strength degradation behavior as temperature difference increases: Hasselman plot and statistical method 10 1.2.1. Strength degradation behavior: the Hasselman plot 10 1.2.2. Strength degradation behavior as temperature difference increases: statistical method 14 1.3. Cyclic thermal shock effects A 18 Section 2 Experimental Procedure 27 2.1. Materials 27 2.2. Polishing 27 2.3. Indentation 30 2.4. The observation of slow crack growth after initial indentation but before commencing thermal shock 32 2.5. The observation of slow crack growth behavior in room temperature deionized water and in 80 0C deionized water in absence of thermal shock 32 2.6. Thermal shock testing 33 2.7. Comparison of transverse and longitudinal crack length for a set of Vickers cracks subjected to thermal fatigue 2.8. The observation of grain bridging in for specimens thermally shocked during this study Section 3 Results and Discussions 3.1. Slow crack growth after initial indentation but before commencing thermal shock 3.2. Crack growth behavior in room temperature and 80 °C deionized water in the absence of thermal shock 3.3. Thermal fatigue behavior for unreinforced alumina 3.3.1. Thermal shock testing for a total number of ten thermal cycles 3.3.2. Thermal shock testing for a total number of twenty thermal cycles 3.3.3. The comparison of rate constant with Lee’s data [12] and Ash’s data [32] 3.3.4. The severity of thermal fatigue damage as a function of the quench temperature difference 3.3.5. Comparison of crack lengths for a given Vickers crack subject to thermal fatigue 3.4. Grain bridging in the brittle materials 3.4.1. An overview of grain bridging 3.4.2. Grain bridging model 3.4.3.1. The effect and observation of grain bridging by Lathabai et a1. [34] 3.4.3.2. The effect and observation of grain bridging by Hay e! a]. [35] 3.4.3.3. The effect and observation of grain bridging by Kagawa [33] 3.4.3.4. The effect and observation of grain bridging by Swanson et a1. [36] 39 41 42 42 51 54 54 69 78 81 83 85 85 86 86 89 93 96 vi 3.4.3.5. The effect and observation of grain bridging by Kishimoto et a1. [37] 3.4.3.6. The effect and observation of grain bridging by Dauskardt [38,39] 3.4.3.7. The effect and observation of grain bridging by Vekinnis et al. [44] 3.4.3.8. The effect and observation of grain bridging by Braun et a1. [45] 3.4.3.9. The effect and observation of grain bridging by Steinbrech et al. [43] 3.4.3.10. The effect and observation of grain bridging by Rodel et al. [46] 3.4.3.11. The effect and observation of grain bridging by Reichl et al. [47] 3.4.5. Crack propagation behavior and grain bridging of the unreinforced polycrystalline alumina included in this study 3.4.6. Grain bridging related to thermal shock for unreinforced alumina Section 4 Summary and Conclusions 4.1. Slow crack growth after initial indentation before commencing thermal shock 4.2. Crack growth behavior in room temperature and 80 °C water in the absence of thermal shock 4.3. Thermal fatigue behavior unreinforced polycrystalline alumina 4.4. Grain bridging in the brittle materials Section 5 Appendices Appendix A. The dimension and label of all specimens Appendix B. The result of slow crack growth testing after initial indentation but before commencing thermal shock 97 100 102 104 105 108 109 113 119 125 125 125 125 127 128 128 129 vii Appendix C. Crack length versus time for Vickers indented specimens immersed in the room temperature deionized water and in 80 0C deionized water 133 Appendix D. The thermal shock data 135 Section 6 References 149 viii List of Tables page Table 2.2.1. Heating history for demounting polycrystalline alumina specimen from aluminum plate and alumina plate. 30 Table 3.1.1. Slow crack growth (in air) after initial indentation for unreinforced polycrystalline alumina specimens. 43 Table 3.3.1. The results of the least-squares fitting to equation 3.3.1 for the data of a total of 10 thermal cycles. (where co = initial crack length, c; = final crack length, N = data points). 67 Table 3.3.2. The least-squares fitting of equation 3.3.1 for 20 quench specimens (where an = initial crack length, a; = final crack length, N = data points). 76 Table 3.3.3. The comparison of rate constants for data of Young’s modulus (Lee’s Work [12]) and data of crack length (this study and Ash’s work [32]) for unreinforced alumina specimen. 78 Table 3.3.4. The results of linear regression of normalized saturation crack length for unreinforced polycrystalline alumina specimen using PLOT-IT program. 81 Table 3.4.1. Lifetime under cyclic loading of alumina for 1 Hz and 100 Hz. A Vickers indentation crack was inserted with 30 N load [34]. 88 Table 3.4.2. Scanning electron microscope in situ observation of crack for alumina specimen under cyclic load [34]. 89 Table 3.4.3. The stress behavior versus displacement of polycrystalline alumina specimen by postfracture test [35]. 91 Table 3.4.4. The effect of grain size and Young’s modulus on crack closure stress for ceramic materials using an FEM program [33]. 96 Table 3.4.5. The energy release rate behavior for alumina specimens with different grain size using double torsion testing [44]. 103 Table 3.4.6. The strength behavior for cracks produced by different indentation loads to observe flaw resistance for Al203-Al2Ti05 composite [45]. 105 Table 3.4.7. The enery release rate as a function of normalized crack length (crack length/width of specimen) in SENB alumina specimen which had 16 pm average grain size where the width of specimen was 7 mm [43]. 108 Table A The dimensions and labels for all polycrystalline alumina specimens as measured by vemier caliper. 128 Table B-1. The crack lengths and elapsed time to examine the slow crack growth after introducing Vickers indentation crack for aulmina specimen (specimen A1-11). 129 Table B-2. The crack lengths and elapsed time to examine the slow crack growth after introducing Vickers indentation crack for alumina specimen (specimen A1-13). 130 Table B-3. The crack lengths and elapsed time to examine the slow crack growth after introducing Vickers indentation crack for alumina specimen (specimen A1-14). 130 Table B4. The crack lengths and elapsed time to examine the slow crack growth after introducing Vickers indentation crack for alumina specimen (specimen A2-8e), where “e” means the edge of specimen. 131 Table B—5. The crack lengths and elapsed time to examine the slow crack growth after introducing Vickers indentation crack for alumina specimen (specimen A2-8m), where “m” means the middle of specimen. 131 Table B-6. The crack lengths and elapsed time to examine the slow crack growth after introducing Vickers indentation crack for alumina specimen (specimen A2-9e), where “e” means the edge of specimen. 132 Table B-7. The crack lengths and elapsed time to examine the slow crack growth after introducing Vickers indentation crack for alumina specimen (specimen A2-9m), where “m” means the middle of specimen. 132 Table C-l. Crack length versus time for Vickers indentation immersed in room temperature deionized water (specimen B1-3). Table C-2. Crack length versus elapsed time for alumina specimen in hot water (~ 80 °C). The specimen held in 80 °C for 2 hours at each times (see section 2.5, specimen Bl-4). Table D-l. Crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A4-13 AT = 250 °C). Table D-2. Crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A4—14 AT = 270 °C). Table D-3. Crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A4-15 AT = 290 °C). Table D-4. Crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A6—2 AT = 295 °C). Table D-5. Crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A5-3 AT = 300 0C). Table 0-6. Crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A5-12 AT = 305 °C). Table 0-7. Crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A5-4 AT = 310 0C). Table D-8. Crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A5-11 AT = 315 0C). Table D-9. Crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A5-5 AT = 320 °C). Table D-10. Crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A5-8 AT = 325 °C). Table 0-1 1. Crack length after cyclic thermal shock test which maximum cycling number is 3 (specimen A5-7 AT = 330 °C). Table 0-12. Crack length after cyclic thermal shock test which maximum cycling number is 20 (specimen A6-3 AT = 250 °C). Table D-13. Crack length after cyclic thermal shock test which maximum cycling number is 20 (specimen A6-4 AT = 270 °C). 133 134 135 135 136 136 137 137 138 138 139 139 140 141 142 xi Table D-14. Crack length after cyclic thermal shock test which maximum cycling number is 20 (specimen B1-5 AT = 290 °C). Table 0-15. Crack length after cyclic thermal shock test which maximum cycling number is 20 (specimen A5-10 AT = 295 0C). Table 0-16. Crack length after cyclic thermal shock test which maximum cycling number is 20 (specimen A5-14 AT = 300 °C). Table D-17. Crack length after cych thermal shock test which maximum cycling number is 20 (specimen A6-1 AT = 305 °C). Table D-18. The crack length of bar-shape specimen under the thermal shock, L1. Table 0-19. The crack length of bar-shape specimen under the thermal shock, L2. Table 0-20. The crack length of bar-shape specimen under the thermal shock, L3. 143 144 145 146 147 147 148 xii List of Figures Figure number page Figure 1.1.1. Temperature and stress distribution for a plate cooled from the surface when specimen is heated and quenched [17,18]. 3 Figure 1.2.1. The Hassleman plot which is a schematic of temperature difference behavior for a ceramic materials [7,21]. 11 Figure 1.2.2. A plot of the mean retained strength versus quench temperature difference for aluminosilicate after thermal shock [8]. 13 Figure 1.2.3. The average retained strength after thermal shock for cylindrical polycrystalline 86 % alumina specimen. The quenching medium is room temperature water [22]. 16 Figure 1.2.4. The average retained strength after thermal shock for borosilicate glass specimens. The quenching medium is liquid nitrogen [23]. 17 Figure 1.3.1. The crack number density (number of cracks/inch) as a function of cumulative thermal cycles for graphite epoxy specimens [10,25]. The solid line was obtained by least-squares fit to equation 1.3.1 [10] 19 Figure 1.3.2. Thermal shock damage in terms of modulus of rupture versus the cumulative number of thermal fatigue cycles for MgO, Al203, and MgO-Cr203 refractories [10,27]. The solid line was obtained by least-squares fit to equation 1.3.1 [10]. 21 Figure 1.3.3. The Young’s modulus versus cumulative thermal shock cycles observed for 60 % alumina refractory [10,28]. The solid line was obtained by least-squares fit to equation 1.3.1 [10]. 22 Figure 1.3.4. Young modulus versus the cumulative number of thermal fatigue cycles for unreinforced polycrystalline alumina quenched in a room temperature deionized water bath. The solid curves represents least-squares fit of the data to equation 1.3.1 [12]. 25 Figure 1.3.5. Schematic diagram of damage saturation behavior subject to thermal fatigue damage for ceramic materials [10-11, 29-30]. 26 Figure 2.2.1. a. Position of specimens on the aluminum plate for polishing. specimens were superglued onto the plate. 28 xiii Figure 2.2.1. b Alumina plate used to place specimens on when burning off superglue residue. Figure 2.2.2. The heating history to demount alumina specimens from alumina plate by burning super glue using an electrical resistance furnace. Figure 2.3.1. The position of indentations on the polycrystalline alumina specimen’s surface for observation of indentation crack growth behavior. Figure 2.6.1. Schematic of the electrical resistance heated furnace used in the thermal fatigue experiments. The average heating rate was about 2 °C/min. Figure 2.6.2. Schematic diagram of thermal shock testing specimen holder which has three parts of steel wire, tongs, and wire screen. Figure 2.6.3.a. The photograph of (a) entire specimen holder which has three parts: steel wire, tongs, and wire net for thermal shock testing Figure 2.6.3. b. The photograph of (b) body of steel wire which was twisted 3 times. Figure 2.7.1. The dimension of bar-shaped polycrystalline alumina specimen and the position of indentation cracks. This part of the study investigated possible asymmetry in the group in the x and y directions during thermal fatigue. Figure 3.1.1. Indentation crack length as a function of elapsed time following indention for unreinforced polycrystalline alumina specimen (specimen Al-l 1). Figure 3.1.2. Indentation crack length as a function of elapsed time following indention for unreinforced polycrystalline alumina specimen (specimen A1-13). Figure 3.1.3. Indentation crack length as a function of elapsed time following indention for unreinforced polycrystalline alumina specimen (specimen A1-14). Figure 3.1.4. Indentation crack length as a function of elapsed time following indentation for unreinforced polycrystalline alumina specimen (specimen A2—8e where “e” means the indentation was placed near the edge of the specimen). 28 29 31 34 36 37 37 45 47 xiv Figure 3.1.5. Indentation crack length as a function of elapsed time following indentation for unreinforced polycrystalline alumina specimen (specimen A2-8m where “m” means the indentation was placed near the middle of the specimen). Figure 3.1.6. Indentation crack length as a function of elapsed time following indentation for unreinforced polycrystalline alumina specimen (specimen A2-9e where “e” means the indentation was placed near the edge of the specimen). Figure 3.1.7. Indentation crack length as a function of elapsed time following indentation for unreinforced polycrystalline alumina specimen (specimen A2-9m where “m” means the indentation was placed near the middle of the specimen). Figure 3.2.1. Crack growth behavior versus elapsed time in room temperature deionized water for unreinforced alumina specimen (specimen B1-3). Figure 3.2.2. The crack growth behavior in deionized water at 80 °C held for 2 hours (specimen B1-4). Crack length measurements were made after the specimen had been at 80 °C (see section 2.5) Figure 3.3.1. The crack growth behavior versus the cumulative number of shock cycles for a total of 10 thermal shock cycles, AT = 250 "C (specimen A4-13). The solid line represents the least-squares fit to equation 3.3.1. Figure 3.3.2. The crack growth behavior versus the cumulative number of thermal shock cycles for a total of 10 thermal shock cycles, AT = 270 “C (specimen A4-14). The solid line represents the least-squares fit to equation 3.3.1. Figure 3.3.3. The crack growth behavior versus the cumulative number of thermal shock cycles for a total of 10 thermal shock cycles, AT = 290 °C (specimen A4-15). The solid line represents the least-squares fit to equation 3.3.1. Figure 3.3.4. The crack growth behavior versus the cumulative number of thermal shock cycles for a total of 10 thermal shock cycles, AT = 295 °C (specimen A6-2). The solid line represents the least-squares fit to equation 3.3.1. 49 50 52 53 57 58 59 XV Figure 3.3.5. The crack growth behavior versus the cumulative number of thermal shock cycles for a total of 10 thermal shock cycles, AT = 300 °C (specimen A5-3). The solid line represents the least-squares fit to equation 3.3.1. Figure 3.3.6. The crack growth behavior versus the cumulative number of thermal shock cycles for a total of 10 thermal shock cycles, AT = 305 0C (specimen A5-12). The solid line represents the least-squares fit to equation 3.3.1. Figure 3.3.7. The crack growth behavior versus the cumulative number of thermal shock cycles for a total of 10 thermal shock cycles, AT = 310 °C (specimen A54). The solid line represents the least-squares fit to equation 3.3.1. Figure 3.3.8. The crack growth behavior versus the cumulative number of thermal shock cycles for a total of 10 thermal shock cycles, AT = 315 °C (specimen A5-11). The solid line represents the least-squares fit to equation 3.3.1. Figure 3.3.9. The crack growth behavior versus the cumulative number of thermal shock cycles for a total of 10 thermal shock cycles, AT = 320 °C (specimen A5-5). The solid line represents the least-squares fit to equation 3.3.1. Figure 3.3.10. The crack growth behavior versus the cumulative number of thermal shock cycles for a total of 10 thermal shock cycles, AT = 325 0C (specimen A5-8). The solid line represents the least-squares fit to equation 3.3.1. Figure 3.3.11 The crack growth behavior versus the cumulative number of thermal shock cycles for a total of 3 thermal shock cycles, AT = 330 °C (specimen A5-7). The solid line represents the least-squares fit to equation 3.3.1. Figure 3.3.12. a. The results of crack propagation behavior due to the temperature difference range of 250 - 300 °C for a total of 10 thermal shock cycles. The solid line represents the least- squares fit to equation 3.3.1. Figure 3.3.12.b. The results of crack propagation behavior due to the temperature difference range of 305-330 °C for a total of 10 thermal shock cycles. The solid line represents the least- squares fit to equation 3.3.1. 60 61 62 63 65 68 68 xvi Figure 3.3.13. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles, AT = 250 °C (specimen A6-3). The solid line represents the least- squares fit to equation 3.3.1. Figure 3.3.14. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles, AT = 270 °C (specimen A64). The solid line represents the least- squares fit to equation 3.3.1. Figure 3.3.15. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles, AT = 290 °C (specimen Bl-S). The solid line represents the least- squares fit to equation 3.3.1. Figure 3.3.16. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles, AT = 295 °C (specimen A5-10). The solid line represents the least- squares fit to equation 3.3.1. Figure 3.3.17.a. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles, AT = 300 0C (specimen A5-10). The solid line represents the least- squares fit to equation 3.3.1. Figure 3.3.17.b. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles, AT = 300 °C (specimen A5-10) without abrupt in crack length at 10th thermal quenching. The solid line represents the least- squares fit to equation 3.3.1. Figure 3.3.18. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles, AT = 305 °C (specimen A6-1) without abrupt increase in crack length at 11 thermal quenching. The solid line represents the least- squares fit to equation 3.3.1. Figure 3.3.19. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles with quench temperature difference range of 250 °C to 305 °C. The solid line represents a least-squares fit to equation 3.3.1. 70 71 72 73 74 74 75 77 xvii Figure 3.3.20. Comparison of rate constants as a function of quench temperature difference for three thermal fatigue studies: this study and Ash’s work [32] and Young’s modulus data (Lee’s work [12]). Each of these three studies involved the thermal fatigue of unreinforced polycrystalline alumina into room temperature deionized water. The solid line was obtained from linear regression of data for this study, dotted line represents linear regression for Lee’s data and double dotted line is a linear regression of Ash’s data using PLOT IT program. 79 Figure 3.4.21 Rate constants (3b) behavior for 20 quenches as AT increases for unreinforced alumina specimen. The quenching medium is deionized water. 80 Figure 3.3.22 Normalized saturation crack length behavior versus quench temperature difference for polycrystalline alumina specimens The solid line (10 quenches), dotted line (20 quenches), and double dotted line (5 quenches [32]) represent a linear regression of the data using PLOT IT program. 82 Figure 3.3.23. The crack propagation behavior of bar-shaped specimen of alumina under cyclic thermal shock testing. L1, L2, and L3 represent the position of indentation crack (see section 2.7). 84 Figure 3.4.1. a. Schematic diagram of a possible evolution of bridging crack (a) deflection for a SiC platelet composite specimen [41]. 87 Figure 3.4.1. b. Schematic diagram of a possible evolution of bridging (b) crack bridging for a SiC platelet composite specimen [411- 87 Figure 3.4.2. Schematic diagram of modified four point bend test specimen of polycrystalline alumina for postfracture tensile test [35]. 90 Figure 3.4.3. The crack closure stress behavior with increasing normalized distance where normalized distance is the ratio of the crack separation to half of grain size. The closure stress versus normalized distance curve shown here was generated by an FEM program [33]. The definition of the crack deflection angle w is shown the above. 94 Figure 3.4.4. The change of bridging area for alumina specimen with increasing crack length as calculated by a finite element method by Kishimoto et a1. [37]. 99 xviii Figure 3.4.5. Experimental technique for observing crack length and stress intensity factor where K..." = maximum stress intensity factor, K. = crack closure stress intensity factor, and Kn“. = minimum intensity factor [48]. The specimens were alumina with grain sizes of 8 and 13 pm [38,39]. 101 Figure 3.4.6. Hysteresis curve of applied load and crack opening displacement as suggested by Dauskardt [38,39]. 102 Figure 3.4.7. Energy release rate as crack extension increases for alumina s-DCB alumina specimen [43]. Crack growth was quasi-static from initial notch to fracture. 106 Figure 3.4.8. Load-displacement behavior and the modified double cantilever specimen used by Reichl to study grain bridging in alumina specimen [47]. 111 Figure 3.4.9. The assumption of geometry near the crack tip to calculate the average bridging stress by Reichl et a1. [47]. 112 Figure 3.4.10. a. The position of SEM observation for unreinforced polycrystalline alumina specimen after ten thermal quenching (a) AT = 320 0C behind the crack tip. 115 Figure 3.4.10. b. The position of SEM observation for unreinforced polycrystalline alumina specimen after ten thermal quenching tip (b) AT = 320 “C at the middle of the radial crack. 115 Figure 3.4.10. c. The position of SEM observation for unreinforced polycrystalline alumina specimen after ten thermal quenching (c) AT = 270 °C around the crack tip. 115 Figure 3.4.10. d. The position of SEM observation for unreinforced polycrystalline alumina specimen after ten thermal quenching (d) AT = 300 °C behind the crack tip. 1 15 Figure 3.4.11. a. Crack propagation shape for polycrystalline alumina specimen after 10 quenched thermal shock at (a) AT =270 116 Figure 3.4.11. b. Crack propagation shape for polycrystalline alumina specimen after 10 quenched thermal shock at (b) AT = 320. 116 xix Figure 3.4.12. a. The crack path for polycrystalline alumina specimen after 10 quenched thermal shock at AT = 320 °C (a) behind the crack tip. 117 Figure 3.4.12. b. The crack path for polycrystalline alumina specimen after 10 quenched thermal shock at AT = 320 “C (b) at the middle of the radial crack. 117 Figure 3.4.13. a. The bridging site for polycrystalline alumina specimens after 10 quenched thermal shock. (a) AT = 270 °C around the crack tip. 1 18 Figure 3.4.13. b. The bridging site for polycrystalline alumina specimens after 10 quenched thermal shock. (b) AT = 300 °C behind the crack tip. 118 Figure 3.4.14. The schematic diagram of the comparison of stress-time behavior for the mechanical loading and thermal loading [30] 121 Figure 3.4.15. The profile of stress distribution for thermal loading and bend test. The slope (do/dy) is much stiffer in thermal loading than in bend test. 123 1. Introduction When brittle materials are heated or cooled rapidly, they are subjected to thermal shock which can lead to microcracking or catastrophic failure [1-6]. In addition, many engineering applications should be considered thermal shock effect if they are under the condition that the ambient temperature differences are large. Thermal shock tests as be classified as either single quench thermal shock tests [7-9] or cyclic thermal shock tests [1 0-1 3]. The single quench thermal shock tests involve heating the specimen and quenching the specimen into water [5,6], oil [14], or air [3,15]. The retained strength is measured typically by a 3 or 4 point bend test. The other is cyclic thermal shock test. Single quench tests are unsatisfactory for applications such as ceramic turbine, computer main frame memory [16], or solar system collector where the materials are periodically heated and cooled. Thermal shock damages can be monitored in variety of methods. Retained strength measurements are often used [6-7], but the retained strengths fluctuate depending on the geometry of specimen. Non—destructive measurements, such as elastic modulus and internal friction, can be used to monitor the evolution of thermal shock damage in a given specimen [IO-l 3]. In this study, Vickers indentation cracks were induced in unreinforced polycrystalline alumina specimens. The crack length was measured during thermal fatigue (cyclic thermal shock). 1.1. Thermal shock resistance parameter Thermal shock resistance parameters indicate properties critical to thermal shock damage. These parameters are approximately quantitative measurements of thermal shock resistance. 1.1.1. Thermal shock resistance without crack initiation When ceramic materials are heated or cooled rapidly, a temperature gradient appears through the specimen thickness (Figure 1.1.1) In general, the thermal stress, on. caused by the difference of the ambient temperature is given as follows for an infinite slab [17,18]. 0'”, = ill—AT (1.1.1) l—v where on. = thermal stress v = Poisson’s ratio a = thermal expansion coefficient AT = temperature difference between furnace and water bath. When fracture occurs, AT = ATm-t and am = can such that [17,18] 0' . (1 — v) AT = C"! en! Ea (1.1.2) where ATcm = critical temperature difference 6cm = fracture strength. From Equation 1.1.1, thermal shock resistance parameter, R, is defined as [17,18] Compression Tension <——> m1“ high T \ Surface of slab, low T Figure 1.1.1. Temperature and stress distribution for a plate cooled from the surface when specimen is heated and quenched [18,19]. am, (1 — v) kam, (1 — V) —— r R' - ———-— R = _ Ea 0 Ea (1.1.3) where k = thermal conductivity. High R and R’ values favor avoidance of crack initiation. Therefore, for high R and R’ values, the strength and thermal conductivity of materials should be high while thermal expansion coefficient, Poisson’s ratio and Young’s modulus should be low. 1.1.2. Thermal shock resistance related to surface flaws Compressibility, B, is the fractional volume change per unit of pressure and is defined as 31—2v = —(—E——2 . (1.2.1) where v = Poisson’s ratio E = Young’s modulus The volume changes can be expressed in terms of compressibility [19] d(V0 — VP) = flVodp (1.2.2) where V0 = volume of body Vp = volume of porosity p = external pressure. Equation 1.2.2 can be rearranged to yield = w (1.2.3) V0 dp and 1 dV, 1 dV, Vodp— +Edp. (1 .2.4) If the lefi side of equation 1.2.4 is the actual compressibility, [3, for a porous elastic material, then equation 1.2.4 is pd VP flu = fi+ ° VoPdP (1.2.5) Walsh gave the strain energy for N small penny shaped cracks with half length, c [19] as dW= 'BZW pdp . (1.2.6) pde = dw= ,6 9(I_ 2V) pdp (1.2.7) where N = number of crack. If equation 1.2.1 and 1.2.7 are inserted into equation 1.2.5, we can obtain equation 1.2.8 since pde is work related to crack formation which is identical to an increase in the strain energy [19]. 1+16(1—v2)Nc3}= 3(1- 2v){1+16(1— v2)Nc3 fl“ =m 9(1—2v)V0 E0 9(1—2v)V0 } (1.2.8). where E0 is Young’s modulus of the crack-free material and B, is as defined for equation 1 1.2.4. The total energy per unit volume is the sum of elastic energy? E (aAT)2 , and the fracture energy of the cracks, 2N1tc2I‘ (the factor of 2 indicates two fracture surface) where I“ is defined surface energy if the crack shapes are circular [l9] 3(aAT)2 E, [1+ 16(1— v2)Nc3 ]" +27ch’r (1.2.9) 2(1- 2v) 90— 2V) W, = %E(aAT)2 +27rNc21“ = . . M where F rs fracture surface energy and N 18 the number of cracks. If we set I = 0, then using equation 1.2.9 we can obtain _ 3(aAT) E, [1 + 16(1— v2)Nc3]_2[16(1-—v2)Nc2 20- 2V) 90- 2V) 3(1 - 2v) ]+47rNcF=0 3(aAT)2E, 16(1-v2)Nc3 _, 16(1—v2)Nc2 2(1—2v) [1+ 9(1—2v) H 3(1—2v) ]=47rNcF. (1.2.10) Rearranging equation 1.2.10 gives 2(1—2v)[1+16(1—v2)Nc3]2[16(1— 112)ch (“An = ”NJ 3E, 9(1 - 2v) 3(1— 2v) 1". (1.2.11) Equation 1.2.11 can be expressed by AT as following nr1—2 2 111- 2N3 _1 2E,c(zz(1_Vi/)2]2[l+ «901/2106 10 2' (1'2'12) AT=[ d W Equation 1.2.12 suggests a critical temperature when -d—c' = 0, thus critical temperature above which crack instability occurs is [7] 7:1"(1—2v)2 ‘ 16(1—v2)Nc3 -1 ATC=[2EOa2(1__V)2]2[1+ 9(1_2V) ]c2. (l.2.l3.a) The excess energy from the difference between the elastic energy release and surface the fracture energy is transferred into kinetic energy of crack growth such that the released energy equals to the surface fracture then the crack propagation terminated when energy is balanced; release of elastic energy = increase in surface fiacture energy thus initial elastic energy - final elastic energy = final surface energy - initial surface energy. From equation 1.2.9, we can obtain the following equation 1.2. 14 [7] for final (critical) crack length 3mm2 E, 16(1— v2)Nc,3 2(1- v) “1+ 9(1—2v) 16(1- v’)Nc,3 9(1—2v) )" —{1+ }"']=27£NI‘(cf2 —c,’) (1.2.14) where co = initial crack length Cf: final crack length or critical crack length. If the initial crack length(co) is much smaller than final crack length(cf), then the terms 16(1—v2)Nc,3 d1 160—1?)ch3 9(1—2v) a" H 9(1-2v) 3 }"(since Cf is very large) in equation 1.2.14 are very small, thus equation 1.2.14 can be simplified such that 3(aA7)’ E, = 2. .2.1 2(1-v) 27er c, (1 5) 16(1— v’)Nc3 9(1- 2 v) If crack length is short in equation 1.2.13, second term of [ 1+ ] can be ignored, then AT becomes 7rl"(1—2v)2 2E,a2(1- v2) T =[ 15m? (1.2.13.b) If AT from equation 1.2.13.b is inserted into equation 1.2.15, the critical crack length Cf becomes [7] 3(1—2v)2 87r(1— v’)c,N c, ={ )5. (1.2.16) Equation 1.2.16 suggests that the critical crack length is a function of initial crack length and a weak function of Poisson’s ratio. Other material properties rarely affect critical crack length. Sack [20] assumed that circular cracks were introduced into a crack- free material and calculated the change of free energy. When a uniform stress is applied normal to a brittle material containing circular cracks and if the cracks are much smaller than the whole body of brittle materials, Sack [20] gave elastic energy of the brittle body as G =_§fl:filazcs. . 3E (1.2.17) In addition, the crack surface energy is 21tc2F thus the total energy change induced by the crack is given as 81— 2 — -—-—-—-( V )02c3. = 2 G 27wF 3E (1.2.18) From equation 1.2.18, 3 critical (maximum) crack length can be obtained when dG/dc = O 81-2 _( Vlazz 47ml" E c =0. (1.2.19) Equation 1.2.19 can be expressed in terms of crack length, c IIEF = . 1.2.20. 0 202 (1 - v2) ( a) If 0 = of, where Cf is fracture stress, then equation 1.2.20 becomes IrEF = . 1.2.20.b c 20,2 (1 — v2) ( ) Equation 1.2.20 can be expressed in terms of fracture stress, 1 0', ={ FE }2. (1.2.21.a) 2c(1— V2) If we apply equation 1.2.21 to the initial flaw population, then E = E0 and c = co, of [7,20] is given by FE, 0/ = {—3— 2c,(1—v ) NI— } . (1.2.21.b) When equation 1.2.16 is divided by equation 1.2.21.b, the initial crack length, co can be removed, then the relationship between critical crack length and fracture strength is [7] 3(1— 100,2}; (1 2 22) c = -—-—-—- . . . f 47tNF E, If the cracks are circular as was assumed by equation 1.2.22, then the crack growth resistance, R’”’ is expressed as equation 1.2.21 [7,16] since crack length is proportional to 0,2(1— v) th ' f e 1nverse o F E FE R”” = --———-. 012(1— v) (1.2.21) To avoid catastrophic crack propagation, R”” should be large. Thus the surface energy and the Young’s modulus should be high and the fracture strength should be low. 10 1.2 Comparing the strength degradation behavior as temperature difference increases: Hassleman plot and statistical method 1.2.1. Strength degradation behavior: the Hasselman plot Many researchers have evaluated thermal shock damage in terms of retained strength as a function of quenching temperature difference. The typical retained strength analysis is represented by the Hasselman plot which is a schematic diagram of retained strength and temperature difference that shows there are 3 temperature regions (Figure 1.2.1). At low quenching temperature differences, there is no change in retained strength. No crack propagation occurs since the thermal stress is not high enough for cracks to grow. To initiate cracks a specific quenching temperature difference is required. The second region is related to the initiation of crack propagation. If cracks propagate, there are two kinds of behavior : (a) a catastrophic decrease in retained strength and (b) a steady decrease of retained strength. Hassleman explained the catastrophic decrease in retained strength in terms of kinetic crack propagation [7]. The steady decrease of retained strength is caused by quasi-static propagation [7]. The catastrophic decrease of retained strength tends to occur in high strength ceramic materials which have the high elastic strain energy and the low resistance to crack growth. Gupta [21] investigated strength degradation through thermal quenching. Water was used as a quenching medium. Gupta [21] observed kinetic crack propagation behavior for both single crystal alumina and polycrystalline alumina specimens with mean grain size of 10, 34, and 40 um while a gradual decrease in strength also was observed for the 11 No thermal shock damage Strength 1 \ \ \ gradual decrease ‘ e in strength I constant I strength I l AT, Temperature difference Figure 1.2.1. The Hassleman plot which is a schematic of temperature difference behavior for a ceramic materials [7,21]. 12 larger grain sized (84 um) alumina specimen. In catastrophic decrease of strength (for 10, 34, and 40 um grain sized specimens), the critical temperature differences (ATc) were about 190 °C. No change in strength was observed at the interval from AT = 190 °C to AT = 280 °C, then a gradual decrease in strength was followed. Hasselman found that the strength decreased gradually as the quench temperature differences increased (without a sudden drop of strength) for low strength materials such as insulating firebrick [7]. The third region corresponds to quasi-static crack propagation. The retained strength decreases monotonically as temperature difference increases. Bradt et al. [9] contended that an increase of strength following quenching or thermal tempering follows the second region. Bradt et al. observed an increase of strength following an interval of quench temperature difference in which strength is constant. For example, a commercial aluminosilicate quenched into silicone oil [8] showed an interval of constant strength followed by a drop of in strength (Figure 1.2.2). However, for quench temperature differences between 1100 °C and 1400 °C, the residual strength of the aluminosilicate increased (Figure 1.2.2). 200 l l r l l 1 as Q 2 150- 5’ or 5 g; 10 m 0? 'c a) .E 3 50F (11 c: 0 r l r 1 r J r l r 1 l r 0 200 400 600 800 1000 1200 13 1400 Quench temperature difference, degree C Figure 1.2.2. A plot of the mean retained strength versus quench temperature difference for aluminosilicate after thermal shock [8]. 14 1.2.2. Strength degradation behavior as temperature difference increases : statistical method The statistical approach to thermal shock damage investigates the retained strength of an entire population of thermally shocked specimens. As an example, Bradt et al. studied a set of 50 alumina specimens [22], a set of 25 aluminosilicate specimens [8] and a set of 50 borosilicates glass specimens [22,23]. All specimen strengths before and after thermal shock were measured and the strength distribution was investigated. Ashizuka et al. [23] measured the strength of borosilicate glass after a thermal shock into room temperature water or into 77 K liquid nitrogen. The strengths ranged from 3 to 115 MPa at a temperature difference of 500 °C, 10 to 112 MPa at a temperature difference of 400 °C, and 15 to 110 MPa at a temperature difference of 300 °C in the water bath while the average strength before thermal shock was about 115 MPa. Strength distribution indicated that initially low strength specimens tend to fracture at low stresses. A rapid decrease in strength after thermal shock did not occur at a specific quench temperature difference. In addition, high strength specimens were not damaged while low strength specimen were severely affected by thermal shock. The average strengths are indicative of a quasi-static crack propagation or steady decrease in retained strength rather than a kinetic crack propagation which can result in a rapid dramatic decrease in retained strength (Figure 1.2.3-4). Brat et al. [9] suggest that a specific quench temperature difference may occur for small specimens size. The specimens’ strength distribution indicated that a specific thermal stress initiated thermal shock damage for each specimen. Bradt et al. [9,23] explained there are two types of flaws after thermal shocking for each specimen. One 15 flaw type is the pre-existent flaw population which existed prior to thermal shock. The other flaw type is induced by the thermal quenching. Thermal shock damage changes the flaw population and the crack arrest behavior occurs. 16 200*- 1001- Average strength, MPa o r I 1 l 4 0 1 00 200 300 Quench temperature difference, degree C Figure 1.2.3. The average retained strength after thermal shock for cylindrical polycrystalline 86% alumina specimen. The quenching medium is room temperature water [22]. 17 200 . l 1 I ‘ ”F <0 0. 2 £2” 0') c g 100 - A a) (D or ‘8 <1) > <1: 0 r 4 1 1 r l r 1 1 0 100 200 300 400 500 Quenching temperature difference, degree C Figure 1.2.4. The average retained strength after thermal shock for borosilicate glass specimen. The quenching medium is liquid nitrogen [23]. 18 1.3. Cyclic thermal shock effects Most experiments involving thermal-stress-induced damage utilize single quench- thermal shock testing [21-25]. However, many ceramic applications are actually involved in repeated cycling in the ambient temperature difference (thermal fatigue) as is the case for mainframe computer memories and solar cells. Thermal fatigue damage is caused by the accumulation of repeated temperature difference (thermal loading). Thus one should consider thermal fatigue damage in designing ceramics for a variety of applications. Cyclic thermal shock tests have been performed with various materials with differing properties. Herakovich [26] observed the cyclic thermal shock damage in the air for graphite epoxy by measuring crack number density (number of cracks/length). Herakovich [26] thermally fatigued the graphite epoxy specimens using a temperature difference about 278 °C in an environmental chamber. Figure 1.3.1 shows. crack density increases as the thermal cycling increases indicating thermal shock damage (crack number density) is a function of cumulative number of thermal cycles. Ainsworth et al. [27] performed cyclic thermal shock testing on MgO, A1203, and MgO-Cr203 refractories with air as the quenching medium. Specimens were placed in a 1204 °C preheated furnace for 10 minutes and then the specimens were cooled in air. Ainsworth et al. [27] evaluated the thermal shock damage by measuring modulus of rupture. The strengths decreased as quenching temperature increased (Figure 1.3.2). Thermal shock damage saturation behavior also is observed in Figure 1.3.2 that is, the l9 Crack density, cracks/ inch 1 l r 0 1o 20 Thermal shock cycles Figure 1.3.1. The crack number density as the function of cumulative thermal fatigue cycles for graphite epoxy specimens]10,25]. The solid line was obtained by least-squares fit to equation 1.3.1 [10]. 20 property (strength in Figure 1.3.2) reaches a steady state value, which indicates a saturation of rnicrocrack damage. . Semler et al. [28] observed thermal fatigue in high alumina refractories by measuring strength (modulus of rupture) and Young’s modulus. Specimens were heated in air to 1000 °C (as measured by a thermocouple) by burner flame and then the specimens were cooled in air to 150-200 °C. Afier 5 thermal shock cycles, a strength decrease of 30-35 percent of strength decrease was observed. Semler et al. also evaluated degradation of Young’s modulus for high alumina refractory, in which thermal shock damage increased as the cumulative number of thermal cycles increased (Figure 1.3.3). A saturation of modulus also was observed for the refractories [28]. 21 r T O Mgo I Aluminosilicate 5 A Magnesia-chrome ‘ C» 1b C . (D .1: _ (I) “U Q) .L“. — —l (U E L. 0 Z -( 11 0 O r 1 r 1 r 0 1 0 20 30 Thermal shock cycles Figure 1.3.2. Thermal shock damage in terms of modulus of rupture versus the cumulative number of thermal fatigue cycles for MgO, A1203, and MgO-Cr203 refractory [10,27] at temperature difference 1180 °C in the air. The solid line was obtained by least-squares fit to equation 1.3.1 [10]. 22 'O m 100 , , .E .9 92 .5. o 80 _ h (D O. a)“ .2 1? o 60 E . yr 0) O c . 8 40 r 1 r l r l r l r >- o 2 4 6 8 10 Thermal shock cycles Figure 1.3.3. The Young’s modulus observed for 60% alumina refractory by quenched in air [10,28]. The solid line was obtained by least-squares fit to equation 1.3.1 [10]. 23 Case et al. [IO-12, 29-3'1] investigated thermal fatigue effects for ceramics and ceramic composites. Thermal shock damage was evaluated nondestructively by measuring elastic modulus and internal friction. In Case et al. ’5 study of thermal fatigue, water [11-12, 29-30], silicon oil [13,30] and liquid nitrogen [13,30] were used as quenching media. In this thesis, we shall compare the crack length measurements done for a water quenching medium (this study) to elastic modulus and internal friction measurements done by Case et al. [10-12, 29-30]. Case et al. also observed that as the quenching temperature difference increases, the severity of the thermal shock damage also increases (Figure 1.3.4). In addition, thermal shock damage tended to saturate afier a few cumulative cycles [IO-12, 29-30]. Figure 1.3.4 shows the Young modulus saturation during thermal fatigue testing as the thermal shock cycles increases. Case et al. obtained empirical equation 1.3.1 [10-12, 29-30] from the thermal fatigue data. E: E0 -a{1-exp(-bN)} (1.3.1) where E= Young’s modulus E0 = Young’s modulus of undamaged specimen a = the thermal shock damage constant 0" ll rate constant N = the number of thermal shock cycles 24 The solid curves in Figure 1.3.4 represent a least-squares fit of the thermal fatigue data to equation 1.3.1 for unreinforced polycrystalline alumina specimens quenched into a room temperature deionized water bath [12]. The typical schematic diagram of thermal shock damage saturation (Figure 1.3.5) may be described in terms of a generalized form of equation 1.3.1, namely (2:0, — A{1—exp(—aN)} . (1.3.2) where Q = a material property of specimen ()0 = a material property of specimen of undamaged specimen A = the thermal shock damage constant or = rate constant N = the number of thermal shock cycles. 25 350 . r r n ' a fl 0 or=2so (U I or=250 95 340“- A or =270 - w. A or: 310 2 . 3 m e e e g 330 - . 'f 4 _co or t g k >9 320 a a a e 7 310 r J L r l l r o 20 4o 60 so 100 Thermal shock cycles Figure 1.3.4. Young’s modulus versus the cumulative number of thermal fatigue cycles for unreinforced polycrystalline quenched into a room temperature deionized water bath. The solid curves represent a least-squares fit of the data to equation 1.3.1. [12]. 26 r T V l r l t l ‘ I r I Undamaged property Material property Saturation damage level Number of thremal cycles Figure 1.3.5. Schematic diagram of damage saturation behavior subject to thermal fatigue damage for ceramic materials [10-11, 29-30]. 2. Experimental Procedure 2.1. Materials Commercial polycrystalline alumina (Coors ADS 995) was cut into 10 mm x 10 mm x 1 mm specimens from 45 mm x 45 mm x 1 mm billets using K.O. Lee cutting machine. The alumina plates were sawed at a depth of 1.27 um per pass (one forward or backward recursion) and the saw speed (left/right) was less than 0.25 cm/minute. The nominal average grain size was 4.0 pm. All specimen dimensions were measured using vemier caliper (Mitutoyo) (Appendix A). 2.2. Polishing Before polishing, sets of 8-12 specimens were mounted on an aluminum plate (Figure 2.2.1) with super glue (Super Glue Corp.). Specimens were polished through 3 steps using a VARI/POL VP- 50 (Leco Co) polishing system. During polishing, the wheel speed was 125 rpm and the applied normal load was about 5-7 lbs. The specimens were polished with 15 um diamond paste (Part No. 810- 914 Leco Corp.) for about 4 hours. Then using 6 pm diamond paste (Part No. 810-873 Leco Corp.), the specimens were polished for about 4 hours. Finally, the specimens were polished with 0.3 pm alumina powder (Buehler micropolish No. 40-6305-080) for about four hours until the specimen surfaces were mirror-like in appearance. To demount the specimens from the aluminum plate, the super glue was burned off in an electrical resistance furnace (Lindberg type 51442) for 60 21:5 minutes at 350 °C. The time-temperature history of the furnace for specimen demounting is given in Figure 2.2.2. and Table 2.2.1 27 28 Alumina specimen Hole for securing aluminum plate during polishing Thickness 20 mm 192 m Alumina plate Aluminum plate (a) (b) Figure 2.2.1. (a) Position of specimens on the aluminum plate for polishing specimens. were superglued onto the plate. (b) Alumina plate used to place specimens on when burning off the superglue residue. 29 600 ‘F Illllil' U rrrrrr ""—- Aluminum plate I' Y Wfirfirhtm T fTV‘I'UI — Alumina plan Temperature, degree C 0 r 1 1111111 1 s LLMLLL . . .rurd 1 1 1.....1 10'1 10° 19‘ . 1o2 103 Elapsed Time, minutes Figure 2.2.2. The heating history to demount alumina specimens from alumina plate by burning super glue using an electrical resistance furnace. 30 Table 2.2.1. Heating history for demounting polycrystalline alumina specimen from aluminum plate and alumina plate. from aluminum ' residue T . °C min. T °C min. 0 25 0 25 27 350 49 550 60 350 60 550 873 25 1351 25 1440 25 1440 25 After demounting, there was a super glue residue on alumina specimens. Therefore, the specimens were placed on an alumina plate (Figure 2.2.1) and the plate plus the specimens were then inserted into furnace and heated for 60 :1: 5 minutes at 550 °C to remove super glue completely (Figure 2.2.2 and Table 2.2.1). To prevent the alumina specimens from “sticking” onto the alumina plate, the specimens were placed on a 1 mm thick coating of alumina powder (Buehler 0.3 micron alpha alumina) which covered the alumina plate. The thickness of the alumina powder bed was measured by a ruler with a 1 mm scale. The alumina plate contains a glassy phase that could cause sticking, but such sticking normally does not occur for temperature below 1200 °C, thus the alumina powder bed is a precaution. 2.3. Indentation A Vickers semirnacroindenter (Digital Semirnacrohardness Tester, Buehler Ltd.) was used to introduce cracks in the specimens. A 49 N load was used with loading time of 17 seconds and loading speed of 80 urn/s. Indentation cracks were introduced at the center or edge of the 1 cm x 1 cm specimen surface (Figure 2.3.1). An optical microscope 31 A y-direction >- x-direction 10mm 4*. > 25mm? ‘ e 10mm 5 mm e V —>- 25mm Figure 2.3.1. The position of indentations on the polycrystalline alumina specimen’s surface for observation of indentation crack growth behavior. 32 (x200) mounted on the indenter was utilized to measure crack lengths. A digital readout gave crack lengths to the nearest 0.1 micron. The procedure for determining crack length was as follows: The marker of the optical microscope is positioned at two crack tips. The crack lengths were then measured in the longitudinal direction (x-direction) and transverse direction (y-direction). 2.4. The observation of slow crack growth after initial indentation but before commencing thermal shock After introducing the indentation crack (see section 2.3) crack lengths were measured using an optical microscope for specimens Al-12 to 14 and A2-8 to 9. Crack growth behavior following the indentation but before commencing thermal shock was measured as a function of elapsed time (see Appendix B). 2.5. The observation of slow crack growth behavior in room temperature deionized water and in 80 0C deionized water in absence of thermal shock To observe slow crack growth in water, the specimen (Bl-3) was inserted into room temperature deionized water. The alumina specimens were positioned on the bottom of a glass beaker (400 mm Kimax) with a water depth of about 7 cm as measured by a ruler (1 mm scale). Crack lengths were measured afier the specimens were removed from the water. The crack length versus elapsed time are given in Appendix B. Crack length behavior in 80 °C also was investigated. Specimen (Bl-4) was inserted into 7 cm deep deionized water in a glass beaker (400 ml Kimax). An alumina specimen 33 was positioned 5 cm from the bottom of beaker using the steel wire screen (0.7 x 0.7 mm grid with individual wires 0.03 mm in diameter). The alumina specimen was heated to about 80 °C on Model 4658 stirrer/hot hot plate (Cole-Partner Instrument Co.). The average heating rate was 3.5 °C/minute. The temperature was measured by mercury in glass thermometer (V WR SCIENTIFIC INC 61016-208) with 1 °C scale. The specimen was kept at 80 °C for about 2 hours then cooled at an average rate of 0.35 °C/minute by turning off the hot plate. Within about 2 and 1/2 hours the water reached room temperature (23 °C). Specimen (Bl-4) was removed fi'om the water and the crack lengths were measured using an optical microscope. The 2.5 hour immersion in 80 °C water was repeated 9 times. The resulting crack lengths and temperatures are given in Appendix C. 2.6. Thermal shock testing After the indentation cracks had aged for at least 24 hours, the specimens were placed in the fumace for thermal shock testing. The furnace, which was constructed for thermal shock testing, was heated by electrical resistance (Figure 2.6.1) at a rate of about 2.0 °C per minute to temperatures ranging from 272 °C to 352 °C. The temperature was measured by a digital readout attached to power controller (CN 5001 K, Omega Engineering Inc.) using a K-type thermocouple. The thermocouple was located about 10 cm from the lid of fumace. The temperature of deionized water bath was typically 22 °C as measured by mercury in glass thermometer (V WR SCIENTIFIC m C. 61016-208) which had a 1 °C scale. Quench bath temperature measurements were made before 34 (a) Top view . Specimen Holder Thermocouple wire Brick Lid Control] - Specimen ln-Tb-——— Water bath Deionized water (b) side view Figure 2.6.1. Schematic of the electrical resistance heated furnace used in the thermal fatigue experiments. The average heating rate was about 2 °C/min. 35 specimen was placed in an electrical furnace. The thermometer was located approximately 3.5 cm from the bottom of water bath and 9.5 cm from the specimen. The specimen was introduced into the furnace using a specimen holder (Figure 2.6.2) which was made for this study. The specimen holder consisted of tongs, stainless steel wire screen, and a steel wire (Figures 2.6.2 and 2.6.3). The tongs were used to hold the specimen in a fixed position within the furnace. If the tongs were removed then the alumina specimen plunged into the deionized water bath. Using the tongs, the specimen was placed approximately 25 cm from the top of the furnace (as measured by a 1 mm scale ruler). A wire screen which has 0.7 x 0.7 mm grid made up of wires 0.03 mm diameter held the specimen. The wire screen was made into a bowl shape approximately 1.8 cm x 1.7 cm and 1.2 cm diameter, as measured by ruler with 1 mm scale. The deionized water flowed through the wire screen during quench thus steel wire screen likely did not significantly perturb the thermal shock test. A steel wire with about 64 cm in length was made from 0.85 mm diameter wire (as measured by ruler with 1 mm scale) and formed the “backbone” of the specimen holder. The steel wire (0.85 mm in diameter) was flexible thus three strands of wire were twisted together (Figure 2.6.3). The specimen remained in the furnace for about 10 to 15 minutes supported by the specimen holder. The specimen then was quenched into the water bath by removing the tongs to move the specimen from the furnace to the water bath. The specimen was located approximately 2.5 cm above (measured by ruler with 1 mm scale) the water bath. 36 T 0 Steel wire 64 cm ‘0 Tongs + 28 cm Wire screen to hold the specimen Figure 2.6.2. Schematic diagram of the thermal shock testing specimen holder which has three parts: steel wire, tongs, and wire screen. (b) Figure 2.6.3. The photograph of (a) entire specimen holder which has three parts of steel wire, tongs, and wire net for thermal shock testing (b) body of steel wire which was twisted 3 times. 38 The time elapsed from the initial specimen movement until the specimen reached the water bath was about 1 second. The water bath consisted of a 27.5 cm x 14.5 cm x 20.5 cm plastic container containing a volume of approximately 5 liters of deionized water. The depth of deionized water was about 14 cm (measured by ruler with 1 mm scale) for all of the thermal shock experiments. After 1 minute, the specimen was removed from the water bath. Water remaining on the specimen was removed with paper towels (Kimwipes 34155 Kimberly- Clark). The specimen was then placed on movable stage of semimacroindenter. Crack lengths were measured using an optical microscope with a digital readout which is attached to semimacroindenter. The elapsed time between removing specimen from the water bath and measuring crack length was about 5 minutes. The thermal shock testing was done in two groups. In the first group the specimens were individually thermally fatigued for total of ten thermal cycles. In the second group, the specimens were individually fatigued for a total of twenty shock cycles. Specimens thermally shocked for 10 cycles were at the following quench temperature differences: 250 (A4-13), 270 (A4-14), 290 (A4-15), 295 (A6-2), 300 (AS-3), 305 (AS-12), 310 (A5- 4), 315 (AS-11), 320 (AS-5), 325 (AS-8), and 330 °C (AS-7). For the specimens quenched 20 times, the quench temperature differences were 250 (A6-3), 270 (A64), 295 (A6-5), 300 (AS-10), 305 (AS-14), and 310 °c (A6-1). 39 2.7. Comparison of transverse and longitudinal crack length for a set of Vickers cracks subjected to thermal fatigue Vickers indentation cracks were introduced on the surface of 56.1 mm x 9.4 mm x 0.93 mm bar specimen (Figure 2.7.1). The dimension , the orientation of the x-y axes of the bar and the position of indentations L1, L2, and L3 are given in Figure 2.7.1 . After 24 hours in air, the specimen was placed in the quench furnace at 282 °C for 10 to 15 minutes (see section 2.6.). The specimen was quenched into a 22 °C deionized water which was about 14 cm deep. The temperature of water bath was measured by a mercury in glass thermometer (V WR SCIENTIFIC INC. 61016-208) at nearly the same position as that used to measure the water bath temperature for the square alumina specimens (see section 2.6). The same water container as that used for thermal fatigue testing (see section 2.6) was used. The specimen holder screen (see Figure 2.6.2) was 4.5 cm long to prevent the bar-shaped specimen from falling from the specimen holder. The specimen was located about 20 cm from the top of electrical resistance furnace. The bar-shaped specimen was plunged into a deionized water bath within less than 1 second. The specimen in the water bath was held at 2.5 cm from the bottom of the water bath. After 1 minute, the specimen was removed from the water bath. Water remaining on the specimen was removed with paper towels (Kimwipes 34155 Kimberly-Clark). After the residual water was blotted of the specimen, the specimen placed on the semimacroindenter stage for crack length measurements. Crack lengths were measured by optical microscope of semimacroindenter which has digital readout (see section 2.6). 40 28.6 mm 10 mm 10 mm 4.6 mm: ,L I ¢ ' > I/ 56.3 mm I 0.96 mm thick 9.1 mm Figure 2.7.1. The dimension of bar-shaped polycrystalline alumina specimen and the position of indentation cracks. This part of the study investigated possible asymmetry in the group in the x and y directions during thermal fatigue. 41 2.8 The observation of grain bridging in for specimens thermally shocked during this study Three thermally shocked specimens (A4-l4 at 270 °C, A5-3 at 300 °C, and A5-5 at 320 °C) were gold coated in sputter coater (Model SC 500, EMSCOPE) for 3 minutes. Specimens were mounted on JEOL 9.5 x 9.5 mm aluminum specimen mounts (TED PELLA, Inc.) After gold coating for 3 minutes, the thickness of gold coating was approximately 21 nm, for a deposition rate of 7 nm/min [32]. Specimens were inserted into a scanning electron microscope (SEM, JEOL, J SM 6400V) chamber using specimen holder. The accelerating voltage was 20 kV. Micrographs of the crack shapes were taken at a magnification of 150 and bridging area micrographs were taken at a magnification of 2500. 3. Results and Discussions 3.1. Slow crack growth after initial indentation but before commencing thermal shock To evaluate the initial slow crack growth for the indentation cracks prior to thermal shock testing, the initial indentation crack lengths were measured using an optical microscope. The crack lengths and time were given by tables in Appendix B. Plots of crack length versus time are shown in Figures 3.1.1-3.1.7. Following each of the Vickers indentations, the cracks grew slowly after introducing the indentation crack. The crack length typically saturated within about 30 minutes following the indentation. In Figures 3.1.1 to 3.1.7, the open circles represent crack length of transverse direction, while the open squares indicate the longitudinal direction. Closed circles represent the crack length for the average of the transverse direction and the longitudinal direction. For the 7 specimens included in the slow crack growth study, the relative change in the where Co = mean of initial crack lengths for the two radial cracks that make up the Vickers crack system. Cmax = mean of the saturated crack lengths for a given Vickers crack system. In this study, AC varied fiom 1.024 to 1.130. The AC’s and the time at which last measurement was done are given in Table 3.1.1. 42 43 Table 3.1.1. Slow crack growth (in air) after initial indentation for unreinforced polycrystalline alumina specimens. Specimen label AC Tm,“ (minutes) Al-ll 1.024 1474 Al-13 1.077 1386 A1-14 1.130 1495 A2-8eM 1.031 1424 A2-8m*** 1.034 1482 A2-9e" 1.056 1436 A2-9m*** 1.035 1435 * Tmax is the time at which last measurement was done. ** e means the edge of specimen. *** 111 means the middle of specimen. _ ‘1?) e 244 ' ‘ e h .4 meancraeklength .9 . M x-direradrlengtn g 240 _ 121/T: y-dircrack length - 5 - N U) c b 2 236 - _ x .- O 4 g . . [i ["1 232 " W N1 .1 .. U .4 l l llllll l L llllll I l l lllllj L l l lllll 10° 101 102 1o3 10‘ Elapsed time after indentaion, (minutes) Figure 3.1.1. Indentation crack lengths (x- and y-direction, see Figure 2.3.1) as a function of elapsed time following indentation for polycrystalline alumina (specimen Al-l 1). 45 250 I I I 11111] I r I IIITVI r r r I rrrrl E 240 T5.) ‘ O L- .2 E 230 - v r as. .c“ -o- _ O) 220 c ._. -~.. 2 ____—Aa—(-:b x 210 _ 0 Lu .4 meancraeklength 0 200. Q/Q x-dir crack length " (‘ Vb y-direracklength 190 g 1 l l llLll 1 l l 1_LLlll l l l I llel L l l l 10° 101 1o2 103 10‘ Elapsed time after indentation, (minutes) Figure 3.1.2. Indentation crack lengths (x- and y-direction, see Figure 2.3.1) as a function of elapsed time following indentation for polycrystalline alumina (specimen A1-13). 46 230 . .....n] T ......., . .....nT r . 1: A 6 2H g 220— “ a _ h .2 C- , E v_ 210— -. g :3) . O 1 j, 200- ' _ x _ Hmeancraeklength o . CU Mx—dlra'acklength h _ —. O 190 ,- 3’11 y-dircraek length U 180Lo ‘ ' ‘44“‘11 1111112 A 1 1 1111113 1 1 1 1““ 4 1o 10 10 10 10 Elapsed time after indentation, (minutes) Figure 3.1.3. Indentation crack lengths (x- and y-direction, see Figure 2.3.1) as a function of elapsed time following indentation for polycrystalline alumina (specimen A1-14). 47 244 - - E _ 4 9 240 - - .2 ' ‘ E i 5‘ 236-" - O) 5 . 1 ; 232 L - o . E l 0 _ Mmeancraeklength , 228r Mx—dircmdrlength 4 ’ . . ....... .D’. ..D...'.*"'°'?°"l°"?‘t‘...f 10° 101 102 103 10‘ Elapsed time after indentation Figure 3.1.4. Indentation crack lengths (x- and y-direction, see Figure 2.3.1) as a function of elapsed time following indentation for polycrystalline alumina (specimen A2-8e where “e” means the indentation was placed near the edge of the specimen). 48 240 I ' 'I'Illl r T rrrrrrl r r Irrrrrl u . Irrrrr A b _ . 8 W1 (:3 2305- _ E V S. *5) 220 C (D l- ; Hmeancracklength O 210 _ M x-dircrack length 8 Va y-dircracklength 200 r r 11_rrrrl r r_errrrL 1 Ll‘lll‘l r r rrrxr 10° 10‘ 102 103 10‘ Elapsed time after indentation Figure 3.1.5. Indentation crack lengths (x- and y-direction, see Figure 2.3.1) as a function of elapsed time following indentation for polycrystalline alumina (specimen A2-8m where “m” means the indentation was placed near the middle of the specimen). 49 I YYYIIUI I l I TTW'YUI I T I I'VTII Y I ‘r'tl' 260 - l L 240 220 r ”ER—WT ‘ ’ T’LT .. _ H mean crack length ‘ Q/‘Q x-dir crack length i 1 1 M y-dircracklength ' 200 m 1 114111 1 mint“. 1 1 11.“. 10° 101 102 103 10‘ Elapsed time after indentation Crack length, (micron) l ALIJJII 1 Figure 3.1.6. Indentation crack lengths (x- and y-direction, see Figure 2.3.1) as a function of elapsed time following indentation for polycrystalline alumina (specimen A2-9e where “e” means the indentation was placed near the edge of the specimen). 50 220 . . .f @rfinr 5:; try-4.11.9.3 . c. _ E 210 - - 2 .9 - . E \f 200 - - .C t q—r - .. 8’ _q;_ 190 r a _ g 4,3 flag? ‘ x _ H .4 mean crack length 8 . Q/O x-dir crack length 5 180: D4 y-dir crack length ‘ 170 1 1 1141111 1 1 1111111 1 11111111 1 141111. 10° 101 102 103 104 Elapsed time after indentation Figure 3.1.7. Indentation crack lengths (x- and y-direction, see Figure 2.3.1) as a function of elapsed time following indentation for polycrystalline alumina (specimen A2-9m where “m” means the indentation was placed near the middle of the specimen). 51 3.2. Crack growth behavior in room temperature and 80 °C deionized water in the absence of thermal shock Water can cause crack extension via an environmentally-assisted slow crack growth process. Water may affect crack growth during thermal shock testing when the specimens were in the water bath. Therefore, the effect of water on crack extension was investigated. A specimen (B 1 -3) was inserted into deionized water at room temperature after being indented and aged for 24 hours in air. The crack lengths were measured as a function of time (Figure 3.2.1). The circles represent mean crack length, the diamonds indicate the transverse direction crack length and squares indicate the crack length in the longitudinal direction. The indentation crack length did not change significantly (less than 1 percent, see Appendix C) during up to 92 hours of immersion in the deionized water. The crack length data as a flmction of immersion time in the deionized water are given in Appendix C. To observe the effect of hot water (~ 80 °C) on a Vickers indentation crack system, a specimen (Bl-4) was first immersed in room temperature deionized water after being indented and aged for 24 hours in air. Then the water and specimen were slowly heated together to about 80 0C in deionized water. The specimen and water were cooled to room temperature (about 23 °C) in the air. The crack length change ( Figure 3.2.2) was less than 1 percent (see Appendix C). Thus the hot water (~ 80 °C) apparently does not cause the cracks to extend. The crack lengths for each measurement during the test are given in Appendix C. 52 l ' l ' 1 fi 1 ' T meanlength ‘ 1.4 ' x-dlrcraeklength .1 %) L f y-dircracklength J C Q . . x 8 b 1 2 - - a) r . .2 q—s E G) .1 O: 1.0% ea} 5? e a; e e ‘ - l 1 4 L 1 1 1 1 1 1 0 20 40 60 80 100 Elapsed time (hour) Figure 3.2.1. Crack growth behavior versus elapsed time in room temperature deionized water for unreinforced alumina specimen (specimen B1-3). 53 ' I I I I fly I I l I I I T I O meancracklenglh 1.4”. D x-dircraeklength _. 5 + y-dlrcracklength U) C 2 l- x 8 h 12 - - O 0 .2 4.: E G) [I 1.0M a ‘31 E 53 @ m E W L l I I. I I I A I I I l I I I I l O 4 8 12 16 Total elapsed time in water, hour Figure 3.2.2. The crack growth behavior in deionized water at 80 °C held for 2 hours for specimen B1-4. Crack length measurements were made after the specimen had been at 80 °C (see section 2.5) 54 3. 3. Thermal fatigue behavior for unreinforced alumina 3. 3. 1. Thermal cycling to a maximum of ten thermal shock cycles Thermal shock tests were done for quench temperature differences ranging from 250 to 330 °C (Figure 3.3.1-3.3.11). The crack length data for specimens that underwent a total of ten thermal shock cycles are given in appendix D. From equation 1.3.2 (section 1.3), the crack length as a function of cumulative number of thermal shock cycles can be expressed as C(N) = cl -(cl —co)exp(—c2N) (3.3.1) where c = the crack length which is a function of the cumulative number of thermal shock cycles co = the initial crack length from data cl = a free parameter which corresponds to saturated crack length C2 = free parameter corresponds to rate constant. The relative crack increments, m at temperature differences ranging from 250 co °C to 330 °C are plotted in Figures 3.3.1-3.3.11. The solid curves in figures were obtained by least-squares fit to equation 3.3.1. For quench temperature differences ranging from 250 °C to 325 °C, the saturated crack length ranged 273.54 pm to 638.52 um (Table 3.3.1). The correlation coefficients of least-squares fit were higher than 0.94 except for quench temperature differences of 300, and 310 °C (Table 3.3.1). The cracks increased with an increasing number of thermal shock cycles. Eventually, the thermal 55 fatigue damage saturated. Figure 3.3.12 gives the results of crack propagation behavior due to the temperature difference at (a) 250-300 °C and at (b) 305-330 °C. At a quench temperature difference 330 °C, the crack growth behavior is different from the crack growth behavior for lower temperature differences. Afier the fourth quench at 330 °C, the crack grew abruptly to the edge of the specimen. Thus the crack length was not measured because it was out of range of the scale on microscope. Figure 3.3.11 shows the crack growth behavior at the quench temperature difference of 330 °C. 56 1.12 a . . . ., .J .. .— .3 C on 1 Relative crack length '2 1.0 l 1 L 1 I 1 l 1 o 2 4 6 3 10 Number of quenches Figure 3.3.1. The crack growth behavior versus the cumulative number of thermal shock cycles a total of 10 thermal shock cycles, AT = 250 °C (specimen A4-13). The solid line represents the least-squares fit to equation 3.3.1. 57 Relative crack length Number of quenches Figure 3.3.2. The crack growth behavior versus the cumulative number of thermal shock cycles a total of 10 thermal shock cycles, AT = 270 °C (specimen A4-l4). The solid line represents the least-squares fit to equation 3.3.1. 58 ‘ r ‘ l T r ' l r 1.12— 1i 5 i * U) l- C Q 1‘, 1.03- 9 r 0 q) r .2 ‘5 T) 1.04 a: .l 1.0 1 l L l i l 1 o 2 4 6 8 10 Number of quenches Figure 3.3.3. The crack growth behavior versus the cumulative number of thermal shock cycles a total of 10 thermal shock cycles, AT = 290 °C (specimen A4-15). The solid line represents the least-squares fit to equation 3.3.1. 59 —l~ N O) 0 Relative crack length is 1 1 l l l 1 l 4 o 2 4 6 8 10 Number of quenches — 0.8 Figure 3.3.4. The crack growth behavior versus the cumulative number of thermal shock cycles a total of' 10 thermal shock cycles, AT = 295 0C (specimen A6-2). The solid line represents the least-squares fit to equation 3.3.]. 60 1.3 7 l 7 l I r I I T . O . $ 4 . . 5 8’ _q_) 1.2 4 x O 9 . 0 o .2 E 1.1 - o a: 1.0 1 L 1 l 1 l 1 4L 1 o 2 4 6 8 10 Number of quenches Figure 3.3.5. The crack growth behavior versus the cumulative number of thermal shock cycles a total of 10 thermal shock cycles, AT = 300 °C (specimen A5-3). The solid line represents the least-squares fit to equation 3.3.1. 61 2-4 ' T I I I I ' I 5 g 2.0 2 x 0 E . O o 1.6 .2 E 0 tr 1.2 1 L 1 l 1 l 1 l 1 o 2 4 6 8 10 Number of quenches Figure 3.3.6. The crack growth behavior versus the cumulative number of thermal shock cycles a total of 10 thermal shock cycles, AT = 305 °C (specimen A5-12). The solid line represents the least-squares fit to equation 3.3.1. 62 16 l L l l L 1 o o o O 4 . -I 5 . 8’ 2 1.4 e x O 9 . 0 o 1 .2 E 1.2 - (D . a: 1.0 V I I I T I 1 I l o 2 4 6 8 10 Number of quenches Figure 3.3.7. The crack growth behavior versus the cumulative number of thermal shock cycles a total of 10 thermal shock cycles, AT = 310 °C (specimen A5-4). The solid line represents the least-squares fit to equation 3.3.1. 63 2.4 . , . , 2 I . 1 5 g 2.0 2 x O S 0 a, 1.6 .2 E 4: CK 1.2 1 l 1 l 1 l 1 J 1 o 2 4 6 8 10 Number of quenches Figure 3.3.8. The crack growth behavior versus the cumulative number of thermal shock cycles a total of 10 thermal shock cycles, AT = 315 °C (specimen AS-ll). The solid line represents the least-squares fit to equation 3.3.1. 2.4 I | I I ' l ‘ l 5 I- g 2.0 .0.) x O E 0 m 1.6 .2 E a) m . 1.2 1 l 1 I 1 l 1 J 1 0 2 4 6 8 10 Number of quenches Figure 3.3.9. The crack growth behavior versus the cumulative number of thermal shock cycles a total of 10 thermal shock cycles, AT = 320 °C (specimen AS-S). The solid line represents the least-squares fit to equation 3.3.1. 65 2-4 I I I I w I I I 5 g 2.0 2 x 0 9 0 a) 1.6 .2 E c) II 1.2 1 l 1 l 1 l 1 l 1 o 2 4 6 8 10 Number of quenches Figure 3.3.10. The crack growth behavior versus the cumulative number of thermal shock cycles a total of 10 thermal shock cycles, AT = 325 0C (specimen A5-8). The solid line represents the least-squares fit to equation 3.3.1. 66 2-4 I I I I I T I I N 0 Relative crack length a» _L N l 1 l 1 l 4 6 8 10 Number of quenches Figure 3.3.11 The crack growth behavior versus the cumulative number of thermal shock cycles a total of 3 thermal shock cycles, AT = 330 °C (specimen A5-7). The solid line represents the least-squares fit to equation 3.3.1. 67 Table 3.3.1. The results of the least-squares fitting to equation 3.3.1 for the data of a total of 10 thermal cycles. (where co = initial crack length, Cf = final crack length, N = data points). Specimen AT(° C) Co (Inn) or (Ian) mum!) 020ml 1* N A4-13 250 244.3 272.3 273.54 0.413 0.978 1 1 A4-l4 270 256.6 293.1 316.12 0.101 0.946 1 l A4-15 290 254.3 284.6 306.80 0.080 0.954 11 A6-2 295 251.4 513.7 542.54 0.260 0.973 1 1 A5-3 300 255.4 327.4 323.21 1.199 0.616 11 A5-12 305 242.3 537.1 638.52 0.157 0.954 11 A5-4 310 251.7 387.0 389.09 1.141 0.852 11 A5-11 315 244.3 538.3 538.69 0.566 0.989 11 A5-5 320 261.1 587.6 585.26 0.274 0.943 11 A5-8 325 247.3 561.2 607.73 0.188 0.979 11 A5-7 330 235.1 526.1 613.25 0.530 0.974 4 * r = correlation coefficient 68 Relative crack length l 0 2 4 6 8 10 Number of quenches (a) 3 I fl l r l l O m-31o I or-azo O (Jr-31o «3 tar-325 g C] or-315 A DT-330 o: . c 2 ’8 e 2 - d) .2 E o rt 1 1 41 1 1 1 l 1 g o 2 4 6 8 10 Number of quenches (b) Figure 3.3.12 The crack growth behavior versus the cumulative number of thermal shock cycles for a total of 10 thermal shock cycles, The solid line represents the least-squares fit to equation 3.3.1. (a) 250 - 300 °C (b) 305-330 0C. 69 3. 3. 2. Thermal cycling to a maximum of twenty thermal shock cycles To observe the crack growth and damage saturation behavior in more detail, the number of thermal shock cycles was increased to 20. Temperature differences were 250, 270 290, 295, 300, and 305 °C. The crack length data are given appendix D. The data of crack length also were least-squares fit to equation 3.3. 1. The results are given in Table 3.3.2. Figures 3.3.13 - 3.3.18 show the results of the crack propagation behavior for each temperature difference. The solid lines in figures were obtained from the least-squares fit. The crack growth behavior also showed crack length saturation. For specimens thermally shocked for 20 cycles, the saturated crack length ranged 264.35 pm to 546.7 pm. The correlation coefficient ranged from 0.804 up to 0.988. All specimens except A5-14 (AT =300) had a correlation coefficient greater than 0.931. Specimen A5-l4 was different from the other specimens in that abrupt crack grth occurred on the 10th quench for this specimen. 70 1.16 . . . . . A 1- .a d S . *5, 1.12 8 .. C . 2 x 0 ~ . g 1.08— ~ 0 . . .2 E . m 4 o: 1.04 4 1O 1 l 1 0 1o 20 Number of quenches Figure 3.3.13. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles, AT = 250 0C (specimen A6-3). The solid line represents the least-squares fit to equation 3.3.1. 71 1.4 . , .1: *5, 1.3- 4 c 9 l- x 0 g 124 _ G) .2 E 0 - m 1.1 1. 1 l 1 0 10 20 Number of quenches Figure 3.3.14. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles, AT = 270 0C (specimen A6-4). The solid line represents the least-squares fit to equation 3.3.1. 72 2.0 .s .1 O) on Relative crack length I: .11 N 1.0. . I . 0 1o 20 Number of quenches Figure 3.3.15. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles, AT = 290 °C (specimen Bl-S). The solid line represents the least-squares fit to equation 3.3.1. 73 2.0 I I 5 1.8"" a) C 2 x 1.6— ' .1 o E o g 1.4- - E . (D 95 1.2- - 1. . ' . 0 10 20 Number of quench Figure 3.3.16. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles, AT = 295 °C (specimen A5-10). The solid line represents the least-squares fit to equation 3.3.1. ' 74 2.4 fl | 5 1- o o o o o o o o o o O ‘5 . 8 — 2.0- -' x . O O C O 8 I- 3 E 0 E 1.6- - 2 d) a: 1.2 - l 0 10 20 Number of quenches (a) T .1: ._. _ a: c 2 x II o 9 4 o 0 _ .2 fl 2 d) o: l 0 1o 20 Number of quenches (b) Figure 3.3.17. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles, (a) AT = 300 °C (specimen A5-10) (b) without abrupt increase in crack length at 10th thermal quenching. The solid line represents the least-squares fit to equation 3.3.1. 75 2.4 5 .- g 2.0 .9 x 0 (U h 0 a, 1.6 .2 E d) n: . 1.2 1 l 1 o 10 20 Number of quenches Figure 3.3.18. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles, AT = 305 °C (specimen A6-l). The solid line represents the least-squares fit to equation 3.3.1. 76 Table 3.3.2. The least-squares fitting of equation 3.3.1 for 20 quench specimens (where ao = initial crack length, a; = final crack length, N = data points). Specimen ATfC) 80 (pm) a; (rim) C1 C2 r N A6-3 250 230.7 265.15 264.35 1.03 0.952 21 A6—4 270 249.7 331.3 330.56 1.17 0.988 1 l B l -5 290 227.6 442.4 442.45 1.07 0.976 21 A5-10 295 230.8 418.4 419.42 0.57 0.984 21 A5-l4 300 244.1 538.4 523.76 0.56 0.804 21 A6-1 305 237.3 546.7 543.06 0.73 0.931 21 r* correlation coefficient N is the number of data points for the least-squares fit. N includes the measurement of the crack length 24 hours after indentation and the crack lengths measured following each of N-l thermal cycles. The crack propagation behavior for the thermal fatigue is given as a function of quench temperature difference in Figure 3.4.19. The crack lengths increase as the number of quenches, eventually the crack lengths saturate. 77 3.2 1 . O DT-250 .4 . DTII270 5 2'8" E1 or=2so '- 1.1.0 0) I- . DT=295 C a) 24_ 11» or-aoo _ x ‘ orsaos 1 8 A A A A A A A A A A A A A h A m U LJ L—J I—I LJ r—I 1f ._, u U U 1.1 fij :1 I. .2 ~ A 10.: 19. m 1.6- ' - a: 1.2 n n A ’\ r\ A A A /\ A A [‘1 A A A L A 1 A V v U v V \J \J \J U U V U U u U V V x) l l l 0 10 20 Number of quenches Figure 3.3.19. The crack propagation behavior versus the cumulative number for a total of thermal shock cycles with quenching temperature difference range of 250 °C to 305 °C. The solid line represents a least-squares fit to equation 3.3.1. 78 3.3.3. The comparison of rate constant with Lee’s data [12] and Ash’s data [32] The rate constant (b for this study and Ash’s work [32], and or for Lee’s work [12]) in equation 3.3.1 is related to quenching temperature difference. The data were obtained by measuring crack lengths in this study and in Ash’s work [32], while Young’s modulus was measured for Lee’s work [12]. The calculated values of rate constant from least- squares fit to equation 3.3.1 were plotted (Figure 3.3.20). The rate constant increased as the quenching temperature difference increased. However, rate constants for 20 quenched specimens were very scattered (Figure 3.4.21). The solid curves in Figure 3.3.20 were obtained from linear regression using a PLOT-IT program (Scientific Programming Enterprise). The correlation coefficient for Young’s modulus data was 0.975 but the Ash’s and our coefficients were poor (Table 3.4.1). The difference in the scatter could stem from the fact that the Young’s modulus measurements represent the effect of an ensemble average for an entire microcrack population in a thermally shocked (not indented) specimen. In contrast to the Young’s modulus measurements by Lee [12], this study relied on a single Vickers indentation crack system to assess the state of thermal shock damage. Table 3.3.3. The comparison of rate constants for data of Young’s modulus (Lee’s Work [12]) and data of crack length (this study and Ash’s work [32]) for unreinforced alumina specimen. slope intercept correlation data points coefficient 3b for this study (10 0.017 -4.52 0.637 11 quenches) a for Lee’s data 0.012 -2.90 0.975 5 3b for Ash’s data 0.015 -2.78 0.578 6 79 3 r l ' l ' I I, l ' T U 36 for this study '3 a. for Lee's data ’41 31) for Ash's data Rate constant -1 1 L 1 I L I 1 I 1 I 1 220 240 260 280 300 320 340 Temperature difference Figure 3.3.20. Comparison of rate constants as a function of quench temperature difference for three thermal fatigue studies: this study and Ash’s work [32] and Young’s modulus data (Lee’s work [12]). Each of these three studies involved the thermal fatigue of unreinforced polycrystalline alumina into room temperature deionized water. The solid line was obtained from linear regression of data for this study, dotted line represents linear regression for Lee’s data and double dotted line is a linear regression of Ash’s data using PLOT IT program. 80 3-6 I I ; I 1 I I 3.2: 0 -I g; C . 7,; 2.8 - - c - . o . l 0 . .9 2.4 - - w . 1 n: _ . . 2.0 - - . O 1 6 1 l 1 l 1 L 1 l 1 T '250 260 270 280 290 300 310 Temperature difference, degree C Figure 3.4.21 Rate constants (3b) behavior for 20 quenches as AT increases for unreinforced alumina specimen. The quenching medium is deionized water. 81 3.3.4. The severity of thermal fatigue damage as a function of the quench temperature difference The severity of thermal shock damage due to an increase in the quench temperature difference was estimated by calculating a normalized saturation crack length for each quench temperature difference. The normalized saturation crack length was defined as saturation crack length/initial crack length. The saturation crack length increased as the quench temperature difference increased (Figure 3.5.1). The solid lines in Figure 3.5.1 were obtained by linear regression using the PLOT-IT program (Scientific Programming Enterprise) and the results are given Table 3.5.1. The slopes of normalized saturation crack length versus quench temperature difference showed good agreement with all specimens. The correlation coefficients for the specimens quenched 20 times were very good. The scatter in the data may be related to the geometry of specimen, the initial flaw distribution, and the size of the flaws. Therefore, the thermal fatigue damage is more severe as the quench temperature difference increases. Table 3.3.4. The results of linear regression of normalized saturation crack length for unreinforced polycrystalline alumina specimen using PLOT-IT program. Intercept Slope Correlation coefficient Data points for 10 quench -4.92 0.019 0.76 11 for 20 quench -5.03 0.020 0.96 5 for 5 quench“ -3.28 0.017 0.66 6 * Ash’s data [32] 82 O 10 quench «\I [ ID Normalized crack saturation length I 1 I 1 I 1 I 1 1 220 240 260 280 300 320 340 Temperature difference I —L Figure 3.3.22 Normalized saturation crack length behavior versus quench temperature difference for polycrystalline alumina specimens. The solid line (10 quenches), dotted line (20 quenches), and double dotted line (5 quenches [32]) represent a linear regression of the data using PLOT IT program. 83 3.3.5. Comparison of crack lengths for a given Vickers crack subject to thermal fatigue Using a 49 N load, three Vickers indentation crack systems were introduced into the long transverse face of a unreinforced alumina specimen (9.1 mm x 56.3 mm x 0.96 mm). The quench temperature difference was 260 °C (see section 2 for details of the experimental procedure). The longitudinal (x-direction) crack length and vertical (y- direction) crack length show a slight directional difference in crack growth (Figure 3.3.20 and Appendix D) during thermal shock testing. The crack grew more in the y-direction than in the x-direction thus the crack propagation was affected by the crack orientation. The ratio of the crack lengths in the x- direction and y-direction increased to 1.22 as the thermal shock cycles increased. After about 3 to 4 thermal shock cycles, the crack length seemed to saturate. 1-4 ‘ 1 fl l ‘ F ' F ' l fl 1 _ we Average A Q/Q L1 _ n D g 1 3 - m 12 D ' - v 4”; 13 >2 '8 to 1.2 - ' X D A A A f. “5 , . we fi421\ . O --"" 1' rem—154.5 3: / g 1.1 r - 1'0 1 I 1 I 1 I 1 I 1 I 1 I 1 0 1 2 3 4 5 6 7 Number of quenches Figure 3.3.23. The crack propagation behavior of bar-shaped specimen of alumina under cyclic thermal shock testing. L1, L2, and L3 represent the position of indentation crack (see section 2.7). 85 3.4. Grain bridging in the brittle materials 3.4.1. An overview of grain bridging Studies of toughening mechanisms in monophase ceramic materials such as alumina have shown that a major toughening mechanism is grain bridging [33-47] including crack deflection [33,41] and crack bowing [3 5]. Grain bridging is observed behind crack tips [33-38]. Grain bridging is affected by the grain size and crack opening displacement. More bridging sites exist in coarse grained specimens than in fine grain sized specimens [34,3 8]. In grain bridging, there is a critical crack opening displacement related to grain size. Steinbech et al. [43] observed bridging for crack opening displacements less than 1/4 of mean grain size and Hay er al. [35] observed bridging for crack opening displacements 1/3 of mean grain size. The crack closure stress tends to be a function of crack opening displacement rather than a function of crack length in a bridging toughened material [3 8-40]. In addition, the frictional coefficient affects the grain bridging toughening; the crack closure forces increase as the frictional coefficient increases [33]. Scanning electron microscope observations provide evidence of grain bridging behind crack tip [34,36,37,41]. Stress-induced microcracking in the wake zone also can impede crack propagation. Typically, for polycrystalline alumina specimens there is no evidence of dispersed microcracking near the crack tip by scanning electron microscope observation. Thus it has been conjectured that stress- induced microcracking near the tip of a macrocrack is not available in monophase polycrystalline alumina [36,41]. 3.4.2. Grain bridging model Chou et al. [41] suggested a model for the evolution of grain bridging as toughening mechanism in a SiC platelet alumina composite. Crack deflection occurs and crack bridging or sliding pullout follows. Chou er al. [41] observed the crack deflection and grain bridging with a scanning electron microscope (SEM). Figure 3.4.1 is a schematic diagram of the toughening mechanism in SiC platelet reinforced alumina. Kawasawa [33] suggested a similar grain bridging model (crack deflection and crack bridging) as a toughening mechanism for polycrystalline alumina. Crack bridging as the toughening mechanism of monophase ceramic material is composed of crack deflection and grain bridging. Bridging sites in turn can be degraded by cyclic loading [34,3 7]. 3.4.3.1. The effect and observation of grain bridging by Lathabai et a1. [34] Lathabai et al. [34] investigated the effect of grain bridging using commercial alumina specimens with an average grain size of 23 to 35 um. Disc shaped specimens 22 mm in diameter and 2 mm thickness used for tension-tension cyclic testing. Plate specimens 100 mm x100 mm x 6 mm were used for a long crack test. Vickers indentation cracks were introduced with a 30 N load (short crack test) and a 50 N load (long crack test). In the short crack test, specimens were loaded sinusoidally with a maximum tensile stress of 100 to 130 MPa keeping the minimum tensile stress at 20 MPa with fi'equency of le 87 (a) '2-IIV'2‘XMWZI’. .‘é. ...'.. 4...,)(<.\ .1. ::>::$:::1:;i‘tl:£fiddzo‘stb-iklzf‘d .4 --.\;-:.5»<~3 M (b) Figure 3.4.]. Schematic diagrams of a possible evolution of bridging (a) crack . deflection (b) crack bridging for a SiC platelet composite specimen [41]. and 100 Hz using electro-servo—hydraulic machine. Lathabai et al. measured lifetime (time to failure) under cyclic tensile loading. Table 3.4.] shows the maximum stress and lifetime. Table 3.4.1. Lifetime under cych loading of alumina for 1 Hz and 100 Hz. A Vickers indentation crack was inserted with 30 N load [34]. Frequency, Largest of lowest of Largest lowest N Hz maximum maximum lifetime, lifetime, (specimens) cyclic stress, cyclic stress, seconds seconds MPa MPa 1 130 110 * 10 16 100 130 100 105 2 28 * Fracture did not occur within 24 to 40 hours. Long crack testing was done using electro-servo-hydraulic machine with stress ratio 0.1 with fi'equency 10 Hz for two alumina compact tension specimens. Toughness increased as crack propagated. The fiacture toughness was 4.0 MPam"2 at a crack extension of 0.9 mm. The fracture toughness was 5.3 MPam"2 at a crack extension 7 mm. Toughness saturated at 5.3 MPamm. Grain bridging sites were observed in-situ using a scanning electron microscope (SEM) during cyclic loading test. Scanning electron microscopy (SEM) indicated an evolution of grain bridging behind the crack tip during cyclic loading. Table 3.4.2 summaries the scanning electron microscope observation. The evidence of the degradation of the grain bridging was the “cumulation of wear debris” under cyclic loading conditions. Thus frictional traction at sliding grain bridging facets gradually was reduced by the contacting grains repeatedly sliding back and forth, resulting in a degradation of specimen toughness. 89 Table 3.4.2. Scanning electron microscope in situ observation of crack for alumina specimen under cyclic load [34]. Magnification Position of Observation observation x 1000 behind crack tip gain bridging x 6500 near dislodgnent of gain bridging gain x 5000 "‘ crack paths x 7500 * crack paths x 5000 * crack paths * not available 3.4.3.2 The effect and observation of grain bridging by Hay et a1. [35] Hay et al. [35] used 99.7% alumina (fabricated by Johnson Matthey, Alpha Alumina A123) which were cut into 8.90 mm x 60 mm x 4 mm specimens to observe gain bridging effects in monophase ceramic materials. The mean gain size of alumina was about 16 pm. Post fracture tensile (PSF) testing was performed on a modified four point bend test in which a second notch is inserted at rear of specimen (Figure 3.4.2). Post fracture tensile testing identified the critical stress and stress distribution in the wake region of the alumina specimen since there is no crack-resisting material in front of the crack tip when crack tip enters second notch at rear of specimen. The results of post fracture test, stress- displacement behavior, showed that the strength increased from 2.7 MP8 to 5.6 MPa as 90 Load Load Wedge Wid otch Load Load Figure 3.4.2. Schematic diagram of modified four point bend test specimen of polycrystalline alumina for postfracture tensile test [35]. 91 the crack opening displacement increased from 5 pm to about 7.6 pm (Table 3.4.3). Strength monotonically decreased with increasing crack opening displacement up to about 50 pm. At geater CODs, there was a loss of grain bridging associated drop in strength (Table 3.4.3). Table 3.4.3. The stress behavior versus displacement of polycrystalline alumina specimen by postfracture test [35]. First measurement of Maximum stress and Last measurement of . stress and displacement stress and displacement displacement l 3.2 MPa at3 um 8.4 MPa at5 um 0.1 MPa at 47m 2 3.7 MPa at 3.5 pm 7.6 MPa at 6 um 0.1 MPa at 28 pm 3 3.1 MPaatSum 5.8 MPaat7um 0.1MPaat37um 4 2.3 MPaat6um 4MPaat9um 0.1MPaat48um 5 1.3 MPaat7.6um 2.1 MPaatll um 0.1MPaat35um Average 2.7 MPa at 5 um 5.6 MPa at 7.6 pm 0.1 MPa at 39 um The gain bridging stress is a function of crack opening displacement and is expressed as follow [35] a = (1 +L)". (3.4.1) "I um where om = critical stress for bridge rupture um = critical crack opening displacement for bridge rupture n = constant 92 In Hay er al. ’s experiments, bridging in alumina became insignificant at crack opening displacement of 1/3 of the grain size. Steinbresh er al. [43] observed significant gain bridging at COD of less than 1/4 of the gain size. The monotonic decrease in strength during the post fracture test is caused by the loss of bridging area. Hay et al. [35] compared the COD criterion of polycrystalline alumina (COD’s of about 1/3 of gain size) with monolithic MgA1203 spinel. Three specimens of monolithic MgA1203 spinel were used for postfracture testing. For the spinel, the average gain size for small grain size specimen was 35 um, 75 pm for intermediate gain size specimen, and 180 um for large gain size specimen. For spinel specimens, gain bridging effects were sigrificant at COD of less than 1/9 of the gain size. Hay et al. suggested that “Application of the COD-gain size criteria fitted to the alumina, was found to yield poor ageement with PFT result (for spinel)...[3 5].” Further, Hay et al. stated that “...Changing the COD-gain size ratio from 1/3 to 1/9 implies that the load transfer mechanism are less effective in the spinel than in the alumina and that the geometric aspects of the microstructure, namely the size, size distribution, and spatial arrangement of gains, do not uniquely control the wake load character. Rather, there may exist a universal geometry factor which describes the effects from these geometric components but is independent of the material properties. The geometric description may require further modification to account for the various material-specific contributions, such as thermal expansion anisotropy. The much larger limiting COD (a critical COD for bridge rupture) at the wake extreme characteristic of alumina probably results primarily from the residual 93 thermoelastic stresses, which promote a more effective load transfer for an equivalent gain interlocking distance. [3 5]” 3.4.3.3 The effect and observation of grain bridging by Kagawa [33] Kagawa [33] modeled gain bridging in a material for which the crack first deflects along a grain boundary. Grain bridging follows when the deflection angle (angle between the path of crack and crack deflection) is between arctangent of the inverse of the half gain size and 90°. The Kagawa model assumes the crack bridging does not occur if the angle is less than arctangent of the inverse of half grain size. To examine the gain bridging process, a finite element method was used. The Kagawa model [33] assumes a gain size between 50-100 pm, a Poisson’s ratio of 0.15 and gains that are either square or hexagonal. The MENTAT—MARC computer progam solved for gain bridging under mode I loading condition (opening mode). The applied load used was less than maximum load required for crack propagation. Figure 3.4.3 shows the effect of the crack deflection angles upon the crack closure stress. At a crack deflection angle of 70°, crack closure stress gadually increased as crack separation increased since relative sliding occurred between contacting gains as the applied loading increased. Normalized distance (Figure 3.4.3) is the ratio of crack separation and half the grain size. 94 Region I Region II Region III +——->-I‘ —~ * ~ 80° ~ Closure \ stress ’ \ ‘ O __ 90 700 Normalized distance \l’ = Crack deflection angle 3” Figure 3.4.3. The crack closure stress behavior with increasing normalized distance where normalized distance is the ratio of the crack separation to half of grain size. The closure stress versus normalized distance curve shown here was generated by an FEM program [33]. The definition of the crack deflection angle W is shown the above. The crack closure stress for gain bridging behavior as a function of crack separation is divided into three regions. In the first region, the stress increases as separation increases to normalized crack ratio of about 0.1 (crack separation/half of gain size) due to interlocking of the gain faces. The second region shows a gadual decrease of crack closure stress with increasing crack separation by virtue of rupture of the interlocked gains. The third region involves a smooth increase of crack closure stress due to relative sliding at contacted gains. Kagawa suggested that crack closure stress versus normalized crack displacement curve indicates that the first region (involving interlocking gains) is a major process in toughening ceramic materials via gain bridging. Kagawa [33] also calculated that the crack closure stress is a function of gain size using an FEM progam. The stress increased as gain size decreased. Larger elastic deformation near gain bridging sites was observed for specimens with lower Young’s moduli (190 and 380 GPa) compared to specimens with higher Young’s moduli (570 GPa). However, the maximum crack closure stress is not affected by Young’s modulus (Table 3.4.4). The friction coefficient affects third region where the crack closure stress increases with increasing crack separation, however, the coefficient of fiiction is neglected in the first and second region. Table 3.4.4. The effect of grain size and Young’s modulus on crack closure stress for ceramic materials using an FEM program [33]. w=80,E=570GPa w=80,Grainsize=60um Grain size, Maximum E, (GPa) Maximum Normalized (um) closure stress, closure stress, crack (MPa) (MPa) separation at Max. Closure stress 60 17 190 15.7 0.006 80 14.6 380 15.7 0.009 100 12.5 570 15.7 0.019 *w = crack deflection angle, E = Young’s modulus. The data were not obtained from experiment but were calculated via a FEM progam [33]. In summary, Kagawa [33] suggested that gain bridging toughing is affected by gain size, Young’s modulus, and the fiictional coefficient. 3.4.3.4. The effect and observation of grain bridging by Swanson er al. [36] Swanson er a1. [36] investigated crack interface gain bridging as a toughening mechanism using alumina having an average gain size of 20 pm. The crack evolution was monitored with in-situ optical microscopy observation of mechanically loaded tapered double cantilever beam (12 mm edge length and 2 mm thickness) specimens. In- situ optical microscopy also followed crack evolution in disks under axial loading. The “active sites” in back of the crack tip were examined to observe the evolution (formation and fissure) of gain bridging. A series of six low magnification (x 100) microgaphs showed crack bridging during crack propagation. The crack interface bridging and an 97 increasing number of gain bridging sites were observed. In the double cantilever experiment, discontinuous crack gowth was observed. To examine strength behavior using a controlled flaw test, Vickers indentation cracks were introduced at the center of a disk specimen (25 mm diameter and 2 mm thickness). Swanson et al. observed crack propagation from the initial loading until fracture. The crack gowth was distinguished by a stabilized gowth without catastrophic gowth. The fracture proceeded mainly by an interganular mode and showed also discontinuous evolution (crack deflection and grain bridging) behind crack tip as the case for the cantilever beam specimens. Microcracking can be a source of resistance of crack gowth near crack tip since the crack can be arrested by microcracks in the fi'ontal zone of cracks. However, there is no evidence of extended frontal microcracking in front of crack tip, therefore, it was unlikely that microcracking affected the frontal zone of the crack. The crack evolution observed by optical microscopy, in both of the above experiments, gave evidences of gain bridging behind the crack tip which provided a mechanism to impede fracture. Observation of the crack evolution showed the crack gew stably and fi'actured without catastrophic crack gowth. 3.4.3.5. The effect and observation of grain bridging by Kishimoto et al. [37] Kishimoto et al. [37] observed crack gowth behavior for polycrystalline alumina specimens (6 specimens for static loading test and 6 specimens for cyclic loading test) 98 having a mean gain size ranging from 1.0 pm to 19 pm in order to examine gain size effects on gain bridging. Compact tension specimens (50 mm x 40 mm x16 mm) were loaded in an electro- servo-hydraulic testing machine. Crack lengths were measured by an optical microscope. The static load ranged from 63 N to 630 N. Sinusoidal stress (stress ratio is 0.1) was used for cyclic loading, with a loading rate 2.0 N/s. The stress intensity factor for crack gowth was much higher in coarse gain sized specimen (3.5 MPamm) than in fine gain sized specimen (2.3 MPamm). The stress intensity factor showed a slight decrease just after cyclic loading because the bridging areas were, Kishimoto et al. explained, degaded by cyclic loading. The slope of the load-strain curve was slightly larger in coarse gain sized specimens than in fine gain sized specimens though Young’s modulus for both specimens were the same. The difference in load-strain slope and in the stress intensity factor can be explained by gain bridging behind of crack tip. The gain bridging is more effective (as calculated by a finite element method [37]) in coarse gain sized specimens than fine gain size specimens. For a large gain size, the bridging site area through metallurgical microscope increased to about 175 mm2 as crack length increased to 10 mm (Figure 3.4.4) while the bridging area rarely increased as the crack length increased. The gain size effect found by Kishimoto [37] agees with the Kagawa’s result [35]. I I I I I I # I I I I I l I I I I I H coarse grain sized specimen + fine grain sized specimen 200- *- 100- - Bridging area, square milimeter o—%r.s.s.e.s.e.s.ts. 2 4 6 10 Crack length, milimeter Figure 3.4.4. The change of bridging area for alumina specimen with increasing crack length as calculated by a finite element method by Kishimoto et al. [37]. 100 3.4.3.6. The effect and observation of grain bridging by Dauskardt [38,39] Dauskardt [3 8,39] examined gain bridging degadation under cyclic fatigue with two coarse gain sized alumina specimens (mean gain sizes of 8 pm and 13 um). Using a 3 mm thick compact tension specimen, testing was done by ASTM standard fatigue crack gowth procedure (ASTM, 1989) modified for ceramics (Figure 3.4.5 [46]). The load ratio (minimum/maximum load) was 0.1 with a frequency of 25 Hz. The crack propagation profile and gain bridging were examined by a scanning electron microscope. In cyclic and monotonic loading, the gain bridging appeared behind the crack tip (as observed in SEM microgaphs). The relation of crack opening displacement to bridging was investigated by measuring crack opening displacement under cyclic loading. The displacement for loading and unloading displayed a hysteresis curve, a schematic of which is shown in Figure 3.4.6. Elastic gain bridging occurred initially during loading. Frictional bridging was followed after the gain pullout. The fiictional sliding resistance decreased thus the stiffness decreased. For unloading, frictional bridging appears following the elastic gain bridging. However, the maximum crack opening displacement measured during the loading-unloading test was about 1.4 pm which is less than about 1/4 of gain size. The degadation of gain bridging under cyclic loading can be explained by repetitive sliding during frictional gain bridging. 101 0.1 pm thick metallic Im __{ DC Potential : Crack length Time Back face strain Applied K .____Km;,,___- Time Compression strain Figure 3.4.5. Experimental technique for observing crack length and stress intensity factor where K...“ = maximum stress intensity factor, K,. = crack closure stress intensity factor, and Km". = minimum intensity factor [48]. The specimens were alumina with grain sizes of 8 and 13 um [38,39]. 102 Elastic bridging Applied Load rictional Elastic bridging bridging Crack opening displacement Figure 3.4.6. Hysteresis curve of applied load and crack opening displacement as suggested by Dauskardt [38,39]. 103 3.4.3.7. The effect and observation of grain bridging by Vekinnis et al. [44] Vekinnis er al. [42] tested in double torsion 96% alumina specimens which had a MgO-CaO-A1203 glass matrix (Coors ADS96R). The specimens were cut into 25 x 55 x 0.64 mm fiorn plates. Seven specimens with an average gain size of 7 11m were loaded in a double-torsion jig mounted in scanning electron microscope (SEM). The energy release rate increased from 41.7 to 77.1 J/m2 with increasing crack length. The energy release rate saturated after cracks gew to about 8.8 mm and the saturation value increased with increasing gain size such that the saturation value of energy release rate was 67.5 .l/m2 for a 4.1 um gain sized specimen and 110 J/m2 for a 9.3 pm gain sized specimen (Table 3.4.5). Table 3.4.5. The energy release rate behavior for alumina specimens with different grain size using double torsion testing [44]. Grain size, htitial energy release Saturated energy release, Crack length at um rate, J/m2 rate J /m2 saturation, mm 4.1 34 67.5 11.2 5.6 33.5 77 10.4 6.24 32 80 10 6.6 33 81.5 8 8.43 40 98.5 8.5 9.32 57 105 7.3 9.75 53 110 7.3 The gain bridging zone length, which may be defined as the crack extension required for the energy release rate to obtain steady state, decreased as gain size increased. For 104 example, a crack extension of 11.2 mm was required for 4.1 um gain sized specimen while 7.3 mm was needed for 9.3 pm gain sized specimen. SEM observations (magrification x 2.0 k to 4.0 k) also showed evidence of gain behind crack tip which can exert a crack closure force. 3.4.3.8. The effect and observation of grain bridging by Braun et al. [45] For A1203-A12Ti05 (20 volume % of AlzTiOS) composite disk specimens 30 mm in diameter and 4 mm thick, Braun et al. [45] introduced an indentation crack at the center of specimens (5 specimens) using a load range of 20-300 N. Fracture toughness was investigated by observing crack gowth. The crack lengths were measured by an optical microscope during the applied loading. Toughness increased with increasing crack length from 1.8 MPam"2 at a crack length of 100 pm to 4.8 MPamm at a crack length of 2000 um. Scanning electron microscope (SEM) observations showed evidence of gain interlocking bridging at crack interface during crack extension. To observe flaw resistance for A1203-A12Ti05 specimen, Vickers indentation cracks were introduced on a disk specimen (20 mm diameter and 3 mm thickness) and strength was measured using a three ball point support test. Strength decreased slightly as indentation load increased from 260 MPa (at an indentation loading 2 N ) to 200 MPa (at an indentation loading 300 N) (Table 3.4.6). However, the toughness was insensitive to indentation load; the fracture toughness was 2.9 MPam”2 for a 300 N indentation load and 2.7 MPamm for 20 N indentation load. 105 Table 3.4.6. The strength behavior for cracks produced by different indentation loads to observe flaw resistance for A1203-Al2Ti05 composite [45]. BdentationloadN 3 5 10 12 100 200 300 I [Strength MP8 260 265 265 235 220 200 218 I The observation of crack length as a function of applied load showed an increasing crack resistance by gain interlocking bridging for the A1203-A12Ti05 composite. SEM microgaphs revealed the evidence of gain bridging. In addition, indentation load- strength behavior indicated A1203-A12Ti05 composite is a highly flaw resistant material. 3.4.3.9. The effect and observation of grain bridging by Steinbrech er al. [43] Steinbrech er al. [43] cold isostatically pressed and then sintered alumina power (CT 8000, Alcoa, Frankfrut am Main, FRG) in air at 1700 °C. The mean gain sizes of the alumina were 4 pm to 17 pm. Specimens were machined into single edge notched beam specimens (SENB) of dimension 7 mm x 60 mm x 5 mm and short double cantilever beam specimens (s-DCB) of dimension 24 mm x 12.5 mm x2 mm. For the s-DCB, the energy release rate increased to about 103 J/m2 as the cracks extended under load (Figure 3.4.7). The initial energy release rate was less than 20 J/m2 (Figure 3.4.7). The energy saturates when the crack opening displacement reaches about 1/4 of the gain size. If crack opening displacement was geater than 1/4 of gain size, no further gain bridging occurred and the energy release rate saturated. 106 120 . I r I I I Energy releasr rate, (Joules/square meter) o 1 I L I 1 I 1 0 1 2 3 4 Crack extension, mm Figure 3.4.7. Energy release rate as crack extension increases for alumina s-DCB alumina specimen [43]. Crack gowth was quasi-static from initial notch to fracture. 107 For SENB specimens, Steinbrech et al. observed the crack resistance behavior as a function of crack length (normalized crack length, crack length/width of specimen) at different notch depths (Table 3.4.7). Initial energy release rates were independent of initial notch depth. The slope of energy release rate versus normalized crack length increased as the initial notch depth increased. For the smallest initial notch depth (1.54 mm), the energy release rate saturated but the other SENB specimens did not show a saturation behavior. Table 3.4.7 implies that energy saturation behavior is related to crack opening displacement. The crack length, when crack opening displacements reached 1/4 of the grain size, increased as initial notch depth increased (T able 3.4.7). Thus the energy release rate saturated for smallest initial crack depth because crack opening displacement reached 1/4 of gain size at about 70 percent of final fracture crack length as explained by the above for s-DCB specimen. However, COD for rest of the SENB specimens reached 1/4 of the gain size at about 88 -100 percent of final crack length thus saturation behavior may not occur. Thus, Steinbrech et a1. [43] investigated gain bridging effects in alumina specimens by measuring the energy release rates. The gain bridging effect displayed a critical crack opening displacement at which no further gain bridging occurred. 108 Table 3.4.7. The energy release rate as a function of normalized crack length (crack length/width of specimen) in SENB alumina specimen which had 16 um average grain size where width of specimen was 7 mm [43]. Notch depth, Initial Initial Final F inal energy Normalized mm normalized energy normalized release rate, crack length crack length release rate, crack length J/m2 at COD = 4 J/m2 um 1.5 0.22 20 0.87 105 0.65 2.8 0.4 20 0.9 105 0.79 4.2 0.6 20 0.9 120 8.7 5.3 0.76 20 0.93 105 9.3 3.4.3.10. The effect and observation of grain bridging by Rodel et al. [46] Rodel et al. [46] observed bridged crack interfaces using a scanning electron microscope. Alumina disk-shaped specimens (100 mm in diameter and 4 mm thick disk) were made by hot pressing fine alumina powder (Sumitomo AKP-HP gade) at 1650 °C for 35 hours under 35 MPa. The density was geater than 99.9 % of theoretical density and the mean gain size was 11 pm. Disk specimens were machined to form 1 mm thick compact tension specimens. Grain bridgings was observed along the crack trace for applied loads of 5 N to 300 N. The measured crack opening displacement (COD) increased with crack extension fi'om 200 nm at a crack extension 190 pm to 1200 nm at crack extension 1700 um, where crack extension defined as the change of crack length from the initial crack length. The energy release rate indicated the existence of a bridging effect since the crack resistance 109 increased as the crack extension increased. The energy release rates were from 10 J/m2 at 0 crack extension and increased to 50 J/m2 for a 1700 um crack extension. 3.4.3.11. The effect and observation of grain bridging by Reichl et al. [47] Reichl et al. [47] determined crack bridging force using polycrystalline alumina whose mean gain size was 16 um. To measure the crack bridging force, double cantilever beam specimens (30 mm x 25 mm x 2 mm) were modified by adding a second notch at the rear of specimen (Figure 3.4.8). The loading-crack opening displacement (COD) curve (Figure 3.4.8) shows that at a COD of 65 pm, the load abruptly decreases when the crack tip meets the second rear notch. The load (Po) corresponding to an abrupt load decrease represents the “load bearing capacity” of wake zone since no obstruction exists at the crack tip. The maximum COD (8",) of interlocking zone and the wake zone itself are depicted in Figure 3.4.9. The bridged zone load, P1,, is calculated from P, = 2&6 l6," (3.4.2) where 8 = crack opening displacement in the direction of applied load 8m = maximum crack opening displacement for the interlocking zone. From Figure 3.4.9, bridging stress is s, = — (3.4.3) Sb = bridging stress 1 10 z = length of bridging zone w = width of specimen Length of bridging zone, 2, is related to COD (Figure 3.4.9) such that — = —— (3.4.4) where a = distance between the direction of applied load and the second notch tip. If equation 3.4.3 is inserted into equation 3.4.2, we can obtain 5 S, = Hug-)2 /wa . (3.4.4) For 12 polycrystalline alumina specimens, Reichl et al. calculated the bridging stress as 29i3 MPa. 111 Load, N Displacement, pm 1 O wedge ‘ ’ rear notch Figure 3.4.8. Load-displacement behavior and the modified double cantilever specimen used by Reichl to study grain bridging in alumina specimens [47]. 112 Crack tip of rear notch End of bridged zone Figure 3.4.9. The assumption of geometry near the crack tip to calculate the average bridging stress by Reichl er al. [47]. 113 3.4.5. Crack propagation behavior and grain bridging of the unreinforced polycrystalline alumina included in this study Using a scanning electron microscope (J EOL, J SM 6400V), crack shape and gain bridging were observed for three polycrystalline alumina specimens that had been quenched into deionized water as a part of this study. Each of the three specimens viewed in the SEM underwent a total of ten thermal shock cycles at a fixed value of AT prior to the SEM measurements. In order to determine whether the nature of the gain bridging changes as the severity of the thermal shock damage changes, the three specimens selected to represent relatively low, moderate, and high values of thermal fatigue damage. For the low thermal fatigue specimen, specimen A4-14 was selected which underwent 10 thermal shocks at a quench -temperature difference of 270 °C. The moderate and highly damaged states were represented by specimens A5-3 and specimen A5-5, respectively. Specimen A5-3 was shocked for ten cycles at AT = 300 °C while specimen A5-5 was shocked for 10 cycles at AT =320 °C. The crack length versus cumulative number of thermal shock data for A4- 14, A5-3, and A5-5 are included in section 3.3 of this thesis. The relative positions of the SEM crack observation crack is given schematically in Figure 3.4.10. The crack shape after thermal fatigue is shown in Figure 3.4.1] for (a) specimen A4-14 with a quench temperature difference of 270 0C (b) specimen A5-5 with a quench temperature difference of 320 °C. The severity of thermal fatigue damage larger 114 at AT = 320 °C than at AT = 270 °C (Figure 3.4.11) which corresponds to the crack length measurement . The crack propagation paths for specimen A5-5 (quench temperature difference 320 0C) were observed at a magrification of x 2500 and an accelerating voltage of 20 kV at positions (a) behind crack tip and (b) near the center of the radial crack (Figure 3.4.12). Both crack propagation paths show that cracks gew in a transganular mode and there is no evidence of gain bridging sites along the surface trace of the crack. Grain bridging does appear behind the crack tip for the specimen quenched at (a) AT = 270 °C and near the crack tip for the specimen quenched at (b) AT = 300 °C (Figure 3.4.13). Grain bridging can be a source of crack gowth resistance. However, the gain bridging sites were not observed in other areas of these two specimens. Thus, it appears that the gain bridging rarely affects crack propagation under cyclic thermal shock testing for the alumina specimens included in this study. 115 (b) 1 T (d) (C) Figure 3.4.10. The position of SEM observation for unreinforced polycrystalline alumina specimen after ten ”thermal quenching (a) AT = 320 °C behind the crack tip (b) AT = 320 °C at the middle of the radial crack (c) AT = 270 °C around the crack tip ((1) AT = 300 °C behind the crack tip. “6 c , , 1 1 4s . 199Pm, F2 L81 JEOL UZBKU X1§@,33mm (b) Figure 3.4.11. Crack propagation shape for polycrystalline alumina specimen after 10 quenched thermal shock at (a) AT =270 (b) AT = 320. 117 ._| E l:l L .3 E1 11 'J '.>=I 3 - - . 3 E: In in JEDL EBKU Figure 3.4.12. The micrograph of crack path for polycrystalline alumina specimen after 10 quenched thermal shock at AT = 320 °C (a) behind the crack tip (b) at middle of the radial crack. 118 Figure 3.4.13. The micrograph of bridging site for polycrystalline alumina specimens after 10 quenched thermal shock. (a) AT = 270 °C around the crack tip (b) AT = 300 °C behind the crack tip. 119 3.4.6. Grain bridging related to thermal shock for unreinforced alumina For mechanical loading of ceramics, gain bridging contributes to crack propagation resistance. Vekinnis et al. [44] measured the energy release rate for 96 % alumina specimens and showed that the energy release rate increased to about 200 % of the initial energy release rate. The increase in toughness for unreinforced alumina with increasing crack length was explained by gain bridging [44]. To evaluate the gain bridging process, a finite element method was used by Kagawa [33]. The calculated crack closure stress reached 17 MPa for ceramic materials using a FEM progam. Grain bridging sites also were observed behind the crack tip in unreinforced alumina specimen via scanning electron microscopy under mechanical loading by Lathabai [34] and Hay er al. [35]. Grain bridging can be degaded by repeated loading and unloading. Interfacial friction during cyclic loading degades the toughness of polycrystalline alumina by fracturing the bridging sites. Kishimoto [37] observed a decrease in fracture toughness, 0.4 MPam"2 immediately after cyclic loading indicating that gain bridging effect decreases under cyclic loading. Therefore, the gain bridging is an mechanism for energy dissipation during crack gowth in unreinforced alumina under mechanical loading. However, the gain bridging was relatively limited under thermal loading for unreinforced alumina specimen included in this study. Thus the gain bridging may not account for the saturation in the thermal fatigue crack length observation in this study. 120 One difference between thermal loading performed in this study and mechanical loading done by other researchers is the duration time of loading. Lee [30] calculated time dependent thermal stress. The thermal surface stress increases rapidly to a maximum stress then decreases slowly while the stress-time behavior is symmetrical in mechanical loading (Figure 3.4.14). The time required for maximum stress was calculated as 11 milliseconds for a water quench of alumina [30]. The maximum mechanical loading time is usually much larger than 10 milliseconds; for example 1000 milliseconds [34,37]. However, the loading time for mechanical loading with a frequency of 100 Hz [37] is similar (10 milliseconds) to thermal loading. The other difference related to loading time is that the slope of stress-time curve (do/dt) is much larger in thermal loading than mechanical loading (Figure 3.4.14). 121 Stress Thermal loading Stress Time Figure 3.4.14. The schematic diagram of the comparison of stress-time behavior for the mechanical loading and thermal loading [30] 122 Under thermal shock testing, the thermal fatigue damage of material properties (for example, crack length [32], strength [27], crack number density [25] and elastic modulus [IO-12]) saturated as the number of thermal cycles increases. The release of elastic energy is a driving force and the surface fracture energy exerts the retarding force during crack propagation. When the energy is balanced, the crack gowth is stopped. At a specific quench temperature difference, the crack gows to a specific value and then crack length saturates; for example a saturated normalized crack length was 1.14 at AT = 250 °C and a saturated normalized crack length was 2.25 at AT = 320 °C in this study. After damage saturation occurs, the fracture surface energy may be a barrier for crack gowth. The spatial distribution through the thickness of the specimen for thermal stress is given in Figure 3.4.15. The stress distribution results from the temperature gadient through the specimen thickness. The surface of specimen is in tension and the stress decreases rapidly from the surface toward the center of specimen (Figure 3.4.15). On the other hand, the stress distribution in mechanical bend loading is linear through the specimen thickness (Figure 3.4.15). The slope (do/dy) is much steeper for thermal loading than for mechanical loading thus the severity of damage on the surface of the specimen is geater in thermal loading than in mechanical loading. 123 Tension )1 Compression Surface \ —> 0 x / Surface (a) Thermal loading Tension )1 Compression Surface + o x Surface (b) Bend test Figure 3.4.15. The profile of stress distribution for thermal loading and bend test. The slope (do/dy) is much stiffer in thermal loading than in bend test. 124 The scatter of rate constant in measuring Vickers indentation cracks, b, is larger than in measuring Young’s modulus, or (Table 3.3.3) which can be explained by following relationship [1 1] E = E0[1—fs] (3.3.6) where E = Young’s modulus E0 = Young’s modulus of crack free material a = crack damage parameter f = a relative weak function of the material’s Poisson’s ratio. The crack damage parameter for a circular crack is defined as [49] a = N < c3 > (3.3.7) where N = number of cracks = average crack length. Therefore, Young modulus is proportional to and represents an average over the range of crack sizes. Vickers indentation crack measurements represents a single pair of cracks, not an entire distribution of cracks. Thus, using Vickers indentation crack measurements to determine the rate constants, b, for crack gowth will result in more scatter than that observed for the rate constant, a, for the Young’s modulus measurement data (Table 3.3.3). 4. Summary and Conclusion 4.1. Slow crack growth after initial indentation before commencing thermal shock A minor amount of slow crack gowth was observed for polycrystalline alumina after initial indentation before thermal shock. The relative change in the mean crack length (emu/co, see section 3.1) was 1.024 to 1.130. Within 30 minutes, the crack gowth saturated. 4.2. Crack gowth behavior in room temperature and 80 °C water in the absence of thermal shock Water can potentially cause crack extension via an environmentally-assistanted slow crack gowth process. Slow crack gowth behavior was observed in room temperature deionized water and in 80 °C deionized water. For the two polycrystalline alumina specimens tested, however, the crack extensions in the water were not sigtificant since the change of crack length was less than 1 percent in each case. Thus, in this study, the deionized water likely did not sigrificantly affect crack gowth during thermal shock test. 4.3. Thermal fatigue behavior unreinforced polycrystalline alumina Thermal shock damage for unreinforced alumina specimens was investigated by measuring the crack length change. Two goups of polycrystalline alumina specimens were used for thermal fatigue testing. For one set, thermal fatigue tests were done for 10 cumulative quenching cycles with a quench temperature difference range of 250 to 330 °C. For the other set, 20 cumulative thermal shocks were done for polycrystalline 125 126 alumina specimens with a quench temperature difference ranging from 250 °C to 305 °C. The cracks gew to 115 percent to 227 percent of initial crack length for the specimens quenched a total of 10 cycles and 115 percent to 230 percent of the initial crack for the specimens quenched a total of 20 cycles. The crack lengths were fitted via a least-squares procedure(see section 3.3.1 and 3.3.2). The rate constants obtained fiom the least-squares fit of thermal shock data to equation 3.3.1, were a function of quench temperature difference. The rate constant increased as the quench temperature increased. Lee et al.’s [12] and Ash’s [32] rate constants behavior showed similar trends. However, the correlation coefficients of this study and Ash’s [32] were very poor but the correlation coefficients for Lee et al.’s data were much better. The difference in the relative scatter of the data may be related to the fact that for Lee et al.’s study, the thermal shock damage assessed by measuring Young’s modulus which is sensitive to the ensemble average of crack length. In contrast, for this study and Ash’s work, the thermal fatigue damage was assessed in terms of a single system of Vickers indentation cracks. The severity of thermal fatigue damage is a function of quench temperature difference, as measured by the saturation crack length. The normalized saturation crack lengths (crack length/initial crack length) were linear regessed using a PLOT IT progam. The slope was 0.019/°C for 10 quenched specimen and 0.020/°C for 20 quenched specimen. Ash’s data [32] showed a good ageement ( 0.017/°C) with the slopes obtained in this study. 127 4.4. Grain bridging in the brittle materials The major toughening mechanism in monophase ceramic material such as alumina is gain bridging [31-45]. Grain bridging is observed behind of crack tip [31-3 5]. Interlocking of separating interface at the bridging is a source of a crack resistance force. Grain bridging is more effective for coarse gained material than for fine gained materials [33,37]. Crack opening displacement also is related to gain bridging showed that gain bridging may occur when the crack opening displacement is less than 1/3 [3 5] or 1/4 [43] of gain size of specimen. Evidence of gain bridging is examined through scanning electron microscope observations behind the crack tip [34, 36,37, 41]. In this study, SEM observations of crack gowth and gain bridging for three unreinforced polycrystalline alumina specimens which had been subjected to a total of 10 f thermal shock cycles for quench temperature differences of 270, 300, and 320 °C. Grain bridging appeared at 1 or 2 sites for specimens quenched at 270 °C and 300 °C. However, gain bridging was not observed in other areas of specimens. Thus, the gain bridging may not account for the crack saturation of crack length under thermal fatigue testing. 5. Appendices Appendix A. The dimension and label of all specimens The following tables list the dimension and label of all specimens used in this study. The positions of the Vickers indentations (see Figure 2.3.1) are given under the column labeled “Position.” For thermal shock testing, the temperature difference, AT, and maximum number of quenching cycles, Amax are included. All dimensions were measured by vernier caliper (Mitutoyo). Table A The dimensions and label for all polycrystalline alumina specimens as measured byvernier caliper. Label AT, °C Position Nmax Width, mm Length, mm Thickness, mm Al-ll * m * 10.15 9.94 1.02 A1-1 3 * m * 14.06 9.93 0.99 A1-14 * m * 10.09 9.96 1.01 A2-8e * e * 10.04 9.95 1.0 A2-8m * m "‘ 10.04 9.95 1.0 A2-9e * e * 10.30 9.96 1.04 A2-9m * m * 10.30 9.96 1.04 Bl-3 * m * 9.91 9.89 0.94 B1-4 * m * 10.36 9.98 0.96 A4-1 3 250 m 10 10.34 10.02 1.03 A4-14 270 m 10 10.06 10.38 1.01 A4-15 . 290 m 10 9.99 9.93 1.01 A6-2 295 m 10 9.87 9.95 1.00 . A5-3 300 m 10 9.95 9.99 1.01 A5-12 305 m 10 10.07 10.07 0.98 A5-4 310 m 10 10.0 10.45 1.02 A5-ll 315 m 10 10.05 10.25 0.99 A5-5 320 m 10 10.05 9.95 0.96 A5-8 325 m 10 10.14 9.97 1.01 A5-7 330 m 10 9.94 10.11 1.03 A6-3 250 m 20 9.95 10.40 1.04 A6-4 270 m 20 13.95 9.95 1.01 Bl-5 290 m 20 10.33 9.91 1.01 A5-10 295 m 20 9.95 10.05 0.90 A5-14 300 m 20 9.94 10.0 0.94 A6-1 305 m 20 9.95 9.90 0.92 128 129 Appendix B. The result of slow crack gowth testing after initial indentation but before commencing thermal shock Following tables are the crack lengths as time elapsed after initial indention crack. The mean crack lengths were calculated by the average of x-directional and y-directional crack lengths. The relative crack lengths are also given by dividing initial crack length. Table B shows the crack length saturated in 30 minutes. A Table B-1. The crack lengths and elapsed time to examine the slow crack growth after introduc'gg Vickers indentation crack for alumina specimen(specimen Al-l 1). Time x-dir. crack y- dir. crack A/ 234.2 B/231.3 mean crack C/232.8 minutes length (um), length (urn), length(um), A B C 2 234.2 231.3 1 1 232.8 1 7 243.7 231.6 1.041 1.001 237.7 1.021 10 244.6 232.6 1.044 1.006 238.6 1.025 13 244.8 232.5 1.045 1.005 238.7 1.025 37 244.1 231.9 1.042 1.003 238.0 1.023 64 244.3 232.1 1.043 1.003 238.2 1.023 96 244.8 232.4 1.045 1.005 238.6 1.025 1320 244.3 231.7 1.043 1.002 238.0 1.023 1357 244.5 231.2 1.044 1.000 237.9 1.022 1474 244.5 232.0 1.044 1.003 238.3 1.024 130 Table B-2. The crack lengths and elapsed time to examine the slow crack growth after introducingVickers indentation crack for alumina (specimen Al-l3). Time x-dir. crack y- dir. crack A/ B/225. mean crack C/211.0 minutes length (um), length (um), 197.0 0 lengthmm). A B C 1 197.0 225.0 1 1 211.0 1 3 203.3 228.5 1.032 1.016 215.9 1.023 5 211.5 229.2 1.074 1.019 220.3 1.044 14 209.7 239.2 1.064 1.063 224.5 1.064 21 209.0 241.3 1.061 1.072 225.2 1.067 26 211.5 241.6 1.074 1.074 226.6 1.074 38 213.5 240.7 1.084 1.070 227.1 1.076 68 212.3 241.2 1.078 1.072 226.8 1.075 837 214.0 239.8 1.086 1.066 226.9 1.075 1 178 213.7 240.2 1.085 1.068 227.0 1.076 1386 213.5 240.9 1.084 1.071 227.2 1.077 Table B-3. The crack lengths and elapsed time to examine the slow crack growth after introducing Vickers indentation crack for alumina specimen (specimen Al-l4). Time x-dir. crack y- dir. crack A/ 180.1 B/216.1 mean crack C/198.1 minutes length (um), length (um), length(um), A B C 1 180.1 216.1 1 1 198.1 1 4 184.4 222.8 1.024 1.031 203.6 1.028 8 186.8 221.6 1.037 1.025 204.2 1.031 13 202.6 221.8 1.125 1.026 212.2 1.071 20 207.2 220.3 1.150 1.019 213.8 1.079 28 208.7 225.6 1.159 1.044 217.2 1.096 35 222.2 224.3 1.234 1.038 223.3 1.127 50 223.1 225.1 1.238 1.042 224.1 1.131 537 223.3 224.7 1.240 1.040 224.0 1.131 660 223.6 224.5 1.242 1.039 224.1 1.131 1294 222.7 225.1 1.237 1.042 223.9 1 . 130 1495 223.0 224.8 1.238 1.040 223.9 1.130 131 Table B4. The crack lengths and elapsed time to examine the slow crack growth after introducing Vickers indentation crack for alumina specimen (specimen A2-8e),' where “e” means the edge of specimen. Time x-dir. crack y- dir. crack A/ 232.3 B/227.9 mean crack C/230.1 minutes length (um), length (um), lengthmm), A B C 2 232.3 227.9 1 1 230.1 1 12 235.3 229.9 1.013 1.009 232.6 1.011 22 240.0 231.6 1.033 1.016 235.8 1.025 29 239.9 230.9 1.033 1.013 235.4 1.023 37 242.4 231.2 1.043 1.014 236.8 1.029 84 241.5 231.5 1.040 1.016 236.5 1.028 143 243.2 231.6 1.047 1.016 237.4 1.032 201 243.6 231.4 1.049 1.015 237.5 1.032 359 243.1 230.9 1.046 1.013 237.0 1.030 514 243.3 231.2 1.047 1.015 237.3 1.031 1220 242.7 231.6 1.045 1.016 237.2 1.031 1424 243.4 230.9 1.048 1.013 237.2 1.031 Table B-5. The crack lengths and elapsed time to examine the slow crack growth after introducing Vickers indentation crack for alumina specimen (specimen A2- 8m), where “m” means the middle of specimen. Tlme x-dir. crack y- dir. crack A/ 206.3 B/228.9 mean crack C/217.6 minutes length (um), length (um), 1ength(p.m), A B C 1 206.3 228.9 1 1 217.6 1 1 1 206.5 230.9 1.001 1.009 218.7 1.005 22 207.7 232.1 1.007 1.014 219.9 1.011 27 208.6 232.8 1.011 1.017 220.7 1.014 34 212.2 233.0 1.029 1.018 222.6 1.023 82 217.9 232.7 1.056 1.017 225.3 1.035 141 218.7 232.2 1.060 1.014 225.5 1.036 199 218.6 232.8 1.060 1.017 225.7 1.037 358 218.4 232.4 1.059 1.015 225.4 1.036 513 218.2 232.7 1.058 1.017 225.5 1.036 1328 218.5 232.2 1.060 1.014 225.4 1.036 1482 218.1 232.1 1.057 1.014 225.1 1.034 132 Table M. The crack lengths and elapsed time to examine the slow crack growth after introducing Vickers indentation crack for alumina specimen (specimen A2-9e), where “e” means the ed e of specimen. Time x-dir. crack y- dir. crack A/ 245.7 B/210.1 mean crack C/227.9 minutes length (m). length (urn). length(m). A B C 2 245.7 210.1 1 1 227.9 1 9 250.0 218.8 1.018 1.041 234.4 1.029 13 251.2 220.2 1.022 1.048 235.7 1.034 18 259.0 221.2 1.054 1.053 240.1 1.054 23 259.9 220.6 1.058 1.050 240.3 1.054 27 260.5 220.5 1.060 1.050 240.5 1.055 31 259.4 220.2 1.056 1.048 239.8 1.052 43 258.7 221.3 1.053 1.053 240.0 1.053 202 259.6 221.5 1.057 1.054 240.6 1.056 357 260.1 221.7 1.058 1.055 240.9 1.057 1053 259.2 220.9 1.055 1.051 240.1 1.053 1264 259.7 221.2 1.057 1.053 240.5 1.055 1436 259.6 221.5 1.057 1.054 240.6 1.056 Table B-7. The crack lengths and elapsed time to examine the slow crack growth after introducing Vickers indentation crack for alumina specimen (specimen A2- 9m), where “m” means the middle of specimen. Tune x-dir. crack y- dir. crack A/ 214.9 B/179.7 mean crack C/ 197.3 minutes length (um), length (um), length(um), A B C 1 214.9 179.7 1 1 197.3 1.000 8 215.0 180.6 1.000 1.005 197.8 1.003 12 217.4 181.2 1.012 1.008 199.3 1.010 18 219.2 181.5 1.020 1.010 200.4 1.016 22 219.0 184.5 1.019 1.027 201.8 1.023 26 219.4 189.5 1.021 1.055 204.5 1.037 42 218.7 188.2 1.018 1.047 203.5 1.031 202 218.9 189.0 1.019 1.052 204.0 1.034 356 219.2 189.1 1.020 1.052 204.2 1.034 1052 219.5 188.7 1.021 1.050 204.1 1.035 1263 219.1 189.0 1.020 1.052 204.1 1.034 1435 218.8 189.6 1.018 1.055 204.2 1.035 133 Appendix C. Crack length versus time for Vickers indented specimens immersed in the room temperature deionized water and in 80 “C deionized water. The specimens were not subjected to thermal shock. Following tables give crack length versus time for an indented and aged specimen immersed in deionized water. Table 01 shows the crack length with elapsed time in room temperature deionized water. Table C-l is the crack length with water temperature at each testing. The mean values are the average of the crack lengths in the x-direction length and in the y-direction. Table C-l. Crack length versus time for Vickers indentation immersed in room temperature deionized water (specimen Bl-3). Time x-dir. y-dil'. mean A/235.4 B/242.2 C/238.9 (hours) (pm), A M). B WLC 0 235.4 242.4 238.9 1 1 1 1 233.6 238.4 236.0 0.992 0.983 0.988 2 234.7 239.6 237.2 0.997 0.988 0.993 3 233.9 238.5 236.2 0.994 0.984 0.989 4 234.2 241.1 236.0 0.995 0.995 0.995 6 233.6 238.4 237.2 0.992 0.983 0.988 17 234.6 240.7 236.7 0.997 0.993 0.995 20 234.2 240.2 236.6 0.995 0.991 0.993 33 233.7 239.6 237.7 0.993 0.988 0.991 45 234.1 241.2 237.7 0.994 0.995 0.995 49 234.2 239.0 23 7.7 0.995 0.986 0.990 75 233.8 240.9 237.4 0.993 0.994 0.994 92 234.0 240.1 237.1 0.994 0.991 0.992 134 Table C-2 is the crack length in 80 0C deionized water at each testing. The mean values are the average of the crack lengths in the x-direction and in the y-direction length. Table C-2. The crack length versus elapsed time in the hot water (~ 80 °C). The specimen held in 80 °C for 2 hours at each times (see section 2.5, specimen Bl-4). Times Temp. X-dll’. y-dir. mean A/234.1 B/237.3 C/235.7 (° C) (1M). A (tun). 3 (Wu), C 0 234.1 237.3 235.7 1 1 1 1 81 231.0 235.5 233.3 0.987 0.992 0.990 2 76 233.4 236.6 235.0 0.997 0.997 0.997 3 84 233.2 235.9 234.6 0.996 0.994 0.995 4 81 233.0 236.4 234.7 0.995 0.996 0.996 5 82 233.5 237.1 235.3 0.997 0.999 0.998 6 79 232.9 236.6 234.8 0.995 0.997 0.996 7 82 233.1 237 235.1 0.996 0.999 0.997 8 83 233 236.7 234.9 0.995 0.997 0.996 9 81 233.4 236.5 235.0 0.997 0.997 0.997 135 Appendix D. The thermal shock data Following tables are all data of thermal shock test for a total of 10 thermal cycles and a total of 20 quenched specimen using an optical microscope (see section 2.6.) Table D-l. The crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A4-13 AT = 250 °C). # of x-direction y-direction A/233.3 A/255.3 (A+B)/2 C/244.3 Quench Mm) Bum!) (00.x) (96y) Mum!) (COM) 0 233.3 255.3 1 1 244.3 1 1 238.7 260.6 1.023 1.020 249.6 1.021 2 235.4 277.3 1.009 1.086 256.3 1.049 3 241.9 287.7 1.036 1.126 264.8 1.083 4 244.7 289.6 1.048 1.134 267.1 1.093 5 244.1 292.0 1.046 1.143 268.0 1.097 6 246.1 292.6 1.054 1.146 269.3 1.102 7 246.6 291.9 1.057 1.143 269.2 1.102 8 248.8 293.9 1.066 1.151 271.3 1.110 9 247.8 295.5 1.062 1.157 271.6 1.112 10 247.8 296.9 1.062 1.162 272.3 1.114 Table D-2. The crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A4-14 AT = 270 0C) # of x- y-direction A/252.6 A/260.4 (A+B)/2 C/256.5 Quench direction Bourn) (90.x) (Cay) M(1Lm) (em) A011“) 0 252.6 260.4 1 1 256.5 1 1 263.7 262.1 1.043 1.006 262.9 1.024 2 260.1 261.4 1.029 1.003 260.7 1.016 3 275.1 275.6 1.089 1.058 275.3 1.073 4 273.1 277.6 1.081 1.066 275.3 1.073 5 273.7 288.6 1.083 1.108 281.1 1.096 6 275.9 293.9 1.092 1.128 284.9 1.110 7 278.3 299.8 1.101 1.151 289.0 1.126 8 279.3 299.9 1.105 1.151 289.6 1.129 9 280.4 304.1 1.110 1.167 292.2 1.139 10 279.8 306.4 1.107 1.176 293.1 1.142 136 Table D-3. The crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A4-15 AT = 290 0C). # of x- y—direction A/255.1 A/253.5 (A+B)/2 C/254.3 Quench direction B(um) (00.x) (90y) MGM) (Cm) A0111!) 0 255.1 253.5 1 1 254.3 1 1 257.5 268.8 1.009 1.060 263.1 1.034 2 258.2 266.9 1.012 1.052 262.5 1.032 3 260.5 268.7 1.021 1.059 264.6 1.040 4 264.5 269.4 1.036 1.062 266.9 1.049 5 268.9 273.9 1.054 1.080 271.4 1.067 6 269.3 273.0 1.055 1.076 271.1 1.066 7 278.5 283.0 1.091 1.116 280.7 1.104 8 278.3 283.4 1.090 1.117 280.8 1.104 9 279.9 284.0 1.097 1.120 281.9 1.108 10 282.4 286.9 1.107 1.131 284.6 1.119 Table D4. The crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A6-2 AT = 295 °C). # of x- y-direction A/250.3 A/252.5 (A+B)/2 C/251.4 Quench direction B(11m) (co,x) (coy) M(11m) (COM) AW) 0 250.3 252.5 1 1 251.4 1 1 363.1 331.3 1.450 1.312 347.2 1.381 2 400.3 342.0 1.599 1.354 371.1 1.476 3 433.9 360.6 1.733 1.428 397.2 1.580 4 437.1 394.0 1.746 1.560 415.5 1.652 5 438.1 477.2 1.750 1.889 457.6 1.820 6 498.7 488.5 1.992 1.934 493.6 1.963 7 516.7 502.2 2.064 1.988 509.4 2.026 8 516.3 505.7 2.062 2.002 511.0 2.032 9 520.4 500.7 2.079 1.982 510.5 2.030 10 524.4 503.1 2.095 1.992 513.7 2.043 137 Table D-S. The crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A5-3 AT = 300 °C). # of x- y—direction A/267.1 A/243.7 (A+B)/2 C/255.4 Quench direction Bum) (00.x) (00y) Mom) (00M) AW) 0 267.1 243.7 1 1 255.4 1 1 294.3 326.2 1.101 1.338 310.2 1.214 2 298.4 327.9 1.117 1.345 313.1 1.226 3 295.0 332.7 1.104 1.365 313.8 1.228 4 300.5 330.2 1.125 1.354 315.3 1.234 5 305.3 332.0 1.143 1.362 318.6 1.247 6 309.4 339.8 1.158 1.394 324.6 1.270 7 308.5 344.6 1.154 1.414 326.5 1.278 8 309.5 344.9 1.158 1.415 327.2 1.281 9 308.1 348.4 1.153 1.429 328.2 1.285 10 308.3 346.6 1 . 154 1.422 327.4 1.282 Table M. The crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A5-12 AT = 305 °C). # of x- y-direction A/249.6 A/235 (A+B)/2 C/242.3 Quench direction Bmm) (00.x) (Cay) Mom) (co...) Auun) 0 249.6 235.0 1 1 242.3 1 1 264.0 263.0 1.057 1.119 263.5 1.087 2 296.2 288.9 1.186 1.229 292.5 1.207 3 389.2 355.0 1.559 1.510 372.1 1.535 4 502.0 432.1 2.01 1 1.838 467.0 1.927 5 544.3 436.6 2.180 1.857 490.4 2.024 6 566.9 449.2 2.271 1.91 1 508.0 2.096 7 571.1 447.9 2.288 1.905 509.5 2.102 8 584.6 457.2 2.342 1.945 520.9 2.149 9 591.2 481.5 2.368 2.048 536.3 2.213 10 589.1 485.2 2.360 2.064 537.1 2.216 138 Table D7. The crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A5-4 AT = 310 °C). # of x- y-direction A/242.7 A7260.8 (A+B)/2 C/251.7 Quench direction B(um) (00.x) (Cay) M(um) (60M) AGED) 0 242.7 260.8 1 1 251.7 1 1 349.6 343.2 1.440 1.315 346.4 1.375 2 364.1 361.0 1.500 1.384 362.5 1.439 3 367.6 363.6 1.514 1.394 365.6 1.451 4 370.0 364.4 1.524 1.397 367.2 1.458 5 374.4 379.1 1.542 1.453 376.7 1.496 6 378.2 389.0 1.558 1.491 383.6 1.523 7 379.6 387.8 1.564 1.486 383.7 1.523 8 377.2 390.9 1.554 1.498 384.0 1.525 9 379.0 391.4 1.561 1.500 385.2 1.529 10 379.0 395.1 1.561 1.514 387.0 1.537 Table D-8. The crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A5-ll AT = 315 °C). # of x- y-direction A/246.9. A/241.7 (A+B)/2 C/244.3 Quench direction Bum) (co.x) (co.y) Mm!!!) (coM) A(11m) 0 246.9 241.7 1 1 244.3 1 1 389.8 315.5 1.578 1.305 352.6 1.443 2 519.8 394.4 2.105 1.631 457.1 1.871 3 527.2 457.3 2.135 1.892 492.2 2.014 4 553.0 458.0 2.239 1.894 505.5 2.069 5 581.8 469.5 2.356 1.942 525.6 2.151 6 590.2 464.6 2.390 1.922 527.4 2.158 7 595.6 470.0 2.412 1.944 532.8 2.180 8 597.9 472.0 2.421 1.952 534.9 2.189 9 600.7 471.3 2.432 1.949 536 2.194 10 607.7 469.0 2.461 1.940 538.3 2.203 139 Table D-9. The crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen AS-S AT = 320 °C). # of x- y—direction A/261.9 A/260.3 (A+B)/2 C/261.1 Quench direction B(p.m) (co,x) (coy) M(um) (0011) AOL!!!) 0 261.9 260.3 1 1 261.1 1 1 341.4 309.7 1.303 1.189 325.0 1.246 2 373.9 354.4 1.427 1.361 364.1 1.394 3 475.9 479.8 1.817 1.843 477.8 1.830 4 511.1 520.9 1.951 2.001 516.0 1.976 5 513.1 538.0 1.959 2.066 525.5 2.012 6 525.2 554.3 2.005 2.129 539.7 2.067 7 564.9 580.4 2.156 2.229 572.6 2.193 8 575.1 589.1 2.195 2.263 582.1 2.229 9 580.3 597.8 2.215 2.296 589.0 2.256 10 578.9 596.3 2.210 2.290 587.6 2.250 Table D-10. The crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A5-8 AT = 325 °C). # of x- y-direction A/249.3 A/245.4 (A+B)/2 C/247.3 Quench direction B(um) (coy) (co,y) M(p.m) (00M) Amt“) 0 249.3 245.4 1 1 247.3 1 1 313.7 359.1 1.258 1.463 336.4 1.359 2 342.3 367.5 1.373 1.497 354.9 1.434 3 355.1 395.1 1.424 1.610 375.1 1.516 4 440.1 461.8 1.765 1.881 450.9 1.822 5 481.3 499.8 1.930 2.036 490.5 1.982 6 487.5 496.7 1.955 2.024 492.1 1.989 7 504.4 504.1 2.023 2.054 504.2 2.038 8 514.5 530.2 2.063 2.160 522.3 2.111 9 517.4 569.8 2.075 2.321 543.6 2.197 10 541.9 580.6 2.173 2.365 561.2 2.268 140 Table D-ll. The crack length after cyclic thermal shock test which maximum cycling number is 3 (specimen A5-7 AT = 330 °C). # of x- y—direction A/235.9 A/234.4 (A+B)/2 C/235.1 Quench direction B(um) (cog) (coy) M(um) (00M) Alum} 0 235.9 234.4 1 1 235.1 1 1 424.3 326.7 1.798 1.393 375.5 1.597 2 598.2 409.6 2.535 1.747 503.9 2.143 3 609.8 442.6 2.584 1.888 526.2 2.238 4 i # * After fourth thermal shock, the crack grew to the edge of specimen. Thus further crack length measurements would not have been meaningful. 141 Table D-12. The crack length after cyclic thermal shock test which maximum cycling number is 20 (specimen A6-3 AT = 250 0C). # of x- y-direction A/235.0 A/226.4 (A+B)/2 C/230.7 Quench direction BUM) (00.x) (my) Mom) (00M) AOL!!!) 0 235.0 226.4 1 1 230.7 1 1 248.6 256.9 1.058 1.136 252.7 1.095 2 255.2 261.7 1.086 1.156 258.4 1.120 3 262.9 266.1 1.118 1.175 264.5 1.146 4 261.4 268.1 1.112 1.184 264.7 1.147 5 260.9 265.2 1.110 1.171 263.0 1.140 6 263.5 267.4 1.121 1.181 265.4 1.150 7 262.4 265.5 1.116 1.172 263.9 1.144 8 260.6 267.5 1.108 1.181 264.0 1.144 9 262.6 264.7 1.117 1.169 263.6 1.142 10 264.1 264.9 1.123 1.170 264.5 1.146 11 260.8 267.7 1.109 1.182 264.2 1.145 12 261.5 264.6 1.112 1.168 263.0 1.140 13 261.0 268.1 1.110 1.184 264.5 1.146 14 262.9 267.5 1.118 1.181 265.2 1.149 15 260.1 267.6 1.106 1.181 263.8 1.143 16 262.3 264.3 1.116 1.167 263.3 1.141 17 260.8 267.9 1.109 1.183 264.3 1.145 18 262.9 267.7 1.118 1.182 265.3 1.149 19 261.2 267.2 1.111 1.180 264.2 1.145 20 262.4 267.9 1.116 1.183 265.1 1.149 142 Table D-l3. The crack length after cyclic thermal shock test which maximum cycling number is 10 (specimen A6-4 AT = 270 °C). # of x- y-direction A/248.1 A/251.4 (A+B)/2 C/249.7 Quench direction B(p.m) (co,x) (00”!) M(um) (06.14) AM) 0 248.1 251.4 1 1 249.7 1 1 282.8 331.8 1.139 1.319 307.3 1.230 2 297.1 347.8 1.197 1.383 322.4 1.291 3 301.9 350.5 1.216 1.394 326.2 1.306 4 301.3 357.6 1.214 1.422 329.4 1.319 5 299.2 359.4 1.205 1.429 329.3 1.318 6 303.2 358.2 1.222 1.424 330.7 1.324 7 302.7 360.2 1.220 1.432 331.4 1.327 8 301.9 362.2 1.216 1.440 332.0 1.329 9 299.4 362.1 1.206 1.440 330.7 1.324 10 302.5 360.2 1.219 1.432 331.3 1.326 After quench 11, the specimen fell into the bottom of water bath then the crack length of y-direction grew to edge of specimen. 143 Table D-l4. The crack length after cyclic thermal shock test which maximum cycling number is 20 (specimen Bl-S AT = 290 °C). # of x- y-direction A/234.1 A/221.1 (A+B)/2 C/227.6 Quench direction B(11m) (00,.) (coy) M(um) (06M) AW) 0 234.1 221.1 1 1 227.6 1 1 408.3 345.9 1.744 1.564 377.1 1.656 2 431.2 386.4 1.841 1.747 408.8 1.796 3 455.4 393.7 1.945 1.780 424.5 1.865 4 482.2 400.4 2.059 1.810 441.3 1.938 5 481.0 402.6 2.054 1.820 441.8 1.941 6 483.4 402.2 2.064 1.819 442.8 1.945 7 482.6 401.3 2.061 1.815 441.9 1.941 8 482.0 403.5 2.058 1.824 442.7 1.945 9 482.5 404.5 2.061 1.829 443.5 1.948 10 481.8 402.0 2.058 1.818 441.9 1.941 1 1 483.5 403.0 2.065 1.822 443.2 1.947 12 482.6 402.2 2.061 1.819 442.4 1.943 13 484.8 404.2 2.070 1.828 444.5 1.952 14 483.1 404.5 2.063 1.829 443.8 1.949 15 482.5 401.5 2.061 1.815 442.0 1.942 16 482.7 404.7 2.061 1.830 443.7 1.949 17 482.5 403.4 2.061 1.824 442.9 1.946 18 481.9 404.0 2.058 1.827 442.9 1.946 19 483.3 403.7 2.064 1.825 443.5 1.948 20 481.3 403.5 2.055 1.824 442.4 1.943 144 Table D-15. The crack length after cyclic thermal shock test which maximum cycling number is 20 (specimen A5-10 AT = 295 0C). # of x- y-direction A/232.6 A/229.1 (A+B)/2 C/230.8 Quench direction B(1,1m) (co ,x) (coy) M(um) (60M) A0611) 0 232.6 229.1 1 1 230.8 1 1 319.5 317.9 1.373 1.387 318.7 1.380 2 364.9 342.0 1.568 1.492 353.4 1.531 3 400.7 344.4 1.722 1.503 372.5 1.613 4 423 .6 389.9 1.821 1.701 406.7 1.762 5 435.6 401.7 1.872 1.753 418.6 1.813 6 433.7 403.0 1.864 1.759 418.3 1.812 7 433.2 403.8 1.862 1.762 418.5 1.812 8 433.5 403.7 1.863 1.762 418.6 1.813 9 433.7 402.8 1.864 1.758 418.2 1.811 10 433.2 403.5 1.862 1.761 418.3 1.812 11 433.3 403.3 1.862 1.760 418.3 1.812 12 433.7 403.2 1.864 1.759 418.4 1.812 13 433.0 403.2 1.861 1.759 418.1 1.811 14 433.6 403.4 1.864 1.760 418.5 1.812 15 430.0 402.6 1.848 1.757 416.3 1.803 16 433.5 403.5 1.863 1.761 418.5 1.812 17 432.7 403.7 1.860 1.762 418.2 1.811 18 433.8 402.8 1.865 1.758 418.3 1.812 19 433.2 403.2 1.862 1.759 418.2 1.811 20 433.6 403.3 1.864 1.760 418.4 1.812 145 Table D-l6. The crack length after cyclic thermal shock test which maximum cycling number is 20 (specimen A5-14 AT = 300 0C). # of x- y—direction A/241 A/247.3 (A+B)/2 C/244.1 Quench direction B(um) (co,x) (coy) M(p.m) (06M) AOL!!!) 0 241.0 247.3 1 1 244.1 1 1 489.9 332.6 2.032 1.344 41 1.2 1.684 2 533.0 359.7 2.211 1.454 446.3 1.828 3 556.7 371.0 2.309 1.500 463.8 1.900 4 560.9 376.5 2.327 1.522 468.7 1.920 5 580.6 382.6 2.409 1.547 481.6 1.972 6 590.3 380.6 2.449 1.539 485.4 1.988 7 588.6 382.5 2.442 1.546 485.5 1.989 8 590.6 381.7 2.450 1.543 486.1 1.991 9 590.2 382.9 2.448 1.548 486.5 1.993 10 592.0 481.7 2.456 1.947 536.8 2.199 11 592.8 479.5 2.459 1.938 536.1 2.196 12 592.7 479.2 2.459 1.937 535.9 2.195 13 592.3 480.8 2.457 1.944 536.5 2.198 14 593.1 482.0 2.461 1.949 537.5 2.202 15 593.1 482.8 2.461 1.952 537.9 2.203 16 591.6 481.5 2.454 1.947 536.5 2.198 17 592.4 481.4 2.458 1.946 536.9 2.199 18 593.7 481.9 2.463 1.948 537.8 2.203 19 592.6 482.0 2.458 1.949 537.3 2.201 20 593.2 483.6 2.461 1.955 538.4 2.205 146 Table D-17. The crack length after cyclic thermal shock test which maximum cycling number is 20 (specimen A6-l AT = 305 °C). # of x- y-direction A/238.4 N236.3 (A+B)/2 C/237.3 Quench direction B(pm) (coy) (Co’y) M01111) (0014) AG“) 0 238.4 236.3 1 1 237.3 1 1 490.4 375.6 2.057 1.589 433.0 1.824 2 497.5 424.9 2.086 1.798 461.2 1.943 3 524.9 446.8 2.201 1.890 485.8 2.046 4 554.5 481.5 2.325 2.037 518.0 2.182 5 556.6 485.7 2.334 2.055 521.1 2.195 6 558.2 492.2 2.341 2.082 525.2 2.212 7 581.3 503.3 2.438 2.129 542.3 2.284 8 580.1 508.0 2.433 2.149 544.0 2.292 9 584.5 514.2 2.451 2.176 549.3 2.314 10 580.8 512.4 2.436 2.168 546.6 2.302 1 1 579.8 509.8 2.432 2.157 544.8 2.295 12 579.4 512.8 2.430 2.170 546.1 2.300 13 582.4 509.3 2.442 2.155 545.8 2.299 14 582.7 511.8 2.444 2.165 547.2 2.305 15 580.2 511.7 2.433 2.165 545.9 2.300 16 582.4 509.4 2.442 2.155 545.9 2.299 17 583.6 511.3 2.447 2.163 547.4 2.306 18 581.2 511.5 2.437 2.164 546.3 2.301 19 583.6 511.4 2.447 2.164 547.5 2.306 20 582.2 511.3 2.442 2.163 546.7 2.303 The following tables give the thermal fatigue behavior for an unreinforced alumina rectangular bar shaped specimen (9.1 mm x 56.3 mm x 0.96 mm ). The crack lengths were measured using an optical microscope through cyclic thermal testing. The relative Vickers indentation cracks are indicated by L1, L2, and L3 (see section 2.7). The crack lengths were measured by L1, L2, L3 order after taking the specimen out of the water. Table D-18. The crack length of bar-shape specimen under the thermal shock, L1. 147 number of quench x-dir. crack length, y—dir. crack length, Lly / le le (tun) Lly (m) 0 218 232.8 1.07 1 225.3 257.4 1.14 2 236.4 270.7 1.15 3 239.5 276.5 1.15 4 240.2 277.8 1.16 5 240.1 277 .6 1.17 6 244.5 278.5 1.14 7 243.6 277.4 1.14 Table D-l9. The crack length of bar-shape specimen under the thermal shock, L2. number of quench x-dir. crack length, y-dir. crack length, L2y / L2x L2x (Wit) L2y (m) 0 234.3 242.7 1.03 1 253.8 298.3 1.18 2 261.3 299.2 1.15 3 261.1 342.2 1.31 4 261.9 343.6 1.31 5 261.8 344.4 1.32 6 259.8 344.3 1.33 7 259.2 343 .9 1.33 Table D-20. The crack length of bar-shape specimen under the thermal shock, L3. 148 number of quench x-dir. crack length, y—dir. crack length, L3y / L3x L3x (Hm) L3y (wt!) 0 220.1 252.8 1.15 1 238.5 272.7 1.14 2 248 280.4 1.13 3 246.3 288.5 1.17 4 247.5 289.9 1.17 5 246 288.2 1.17 6 246.9 289.6 1.17 7 245.3 288.5 1.18 6. References 1. W.P. Roger and A.F. Emeny, “Contact Thermal Shock Test Ceramics,” J. of Mat. Sci. 27, 146-152 (1992). 2. EH. Lutz and M.V. Swain, “Interaction between Flaw Resistance, R-curve behavior, and Thermal Shock degradation in Ceramics,” J. Am. Ceram. Soc. 74 [11] 2859- 68 (1991). 3. K. T. Faber, M.D. Huang, and .A.G. Evans, “Quantitative of Thermal Shock in Ceramics on a Novel Test Technique,” J. Amer. Ceram. Soc. 64, 296-301 (1981). 4. T.K. Gupta, “Crack Healing in Thermally shocked MgO,” J. Am. Ceram. Soc. 58 [43] (1975). 5. RF. Lange and T.K. Gupta, “Crack Healing by Heat Treatment,” J. Am. Ceram. Soc. Discussion and Notes, 53, 54-55 (1976). 6. T.K. Gupta, “Crack Healing and Strengthening of Thermally Shocked Alumina,” J. Am. Ceram. Soc., 259-61(1976). 7. D.P.H. Hasslemann, “Unified Theory of Thermal Shock Fracture Initiation and Crack Propagation in Brittle Ceramics,” J. Am. Ceram. Soc., 52, 600-04 (1969). 8. H. Ohira and R.C. Bradt, “Strength Distribution of a Quench-strengthen Aluminasilicate,” J. Am. Ceram. Soc., 71 [1], 35-41 (1988). 9. R.C. Bradt, M. Asnizuka, T.E. Easler and H. Ohira, “Statistical Aspects of the Thermal Shock Damage and the Quench Strengthening of Ceramics,” Thermal Shock and Thermal Fatigue Beh_avior of Advanced Ceramics. G. A. Schneider and G. Petzow 447-58, Kluwer Academic Publishers, the Nederlands (1993). 10. W.J. Lee and ED. Case, “Comparison of Saturation Behavior of Thermal Shock Damage in Variety of Brittle Materials,” Mat. Sci. and Eng., A154 1-9 (1992). 11. ED. Case, Y. Kim, W.J. Lee, “Thermal Shock and Thermal Fatigue Behavior of Advanced Ceramics,” Thermg Shock and Thermal Fatigue Behavior of Adianced Ceramics, G. A. Schneider and G. Petzow 393-406, Kluwer Academic Publishers, the Nederlands (1993). 12. W.J. Lee and ED. Case, “Thermal Fatigue in Polycrystalline Alumina,” J. of Mat. Sci., 25, 5053-5054 (1990). 149 150 13. Y. Kim, “Cyclic Thermal Shock in Ceramics and Ceramic Composites,” Ph D. Dissertation, Michigan State Univ., MI (1991). 14. E.H. Lutz, M.V. Swain, and N. Claussen, “Thermal Shock Behavior of Duplex Ceramics,” J. Am. Ceram. Soc. 74[1], 19-24 ( 1991). 15. J .R. Brockenbrough, L.E. F orsythe and R.L. Rolf, “Reliability of Brittle Materials in Thermal Shock,” J. Am. Ceram. Soc., 64, 634-37(1986). 16. J .C. Coppla and R.C. Bradt, “Thermal Shock Damage in SiC,” J. Am. Ceram. Soc., 56, 214-218 (1973). 17. W.D. Kingery, H.K.Bowen, and DR Uhlmann, “Introduction to Ceramics,” 2nd ed. 816-830, John Wiley and Sons, Inc., New York (1976). 18. W.D. Kingery, “Factors Affecting Thermal Stress Resistance of Ceramics Materials,” J. Am. Ceram. Soc., 38, 3-15 (1955). 19. J .B. Walsh, “Effect of Cracks in the Compressibility of Rock,” J. Geophy. Res., 70 [2], 381-389 (1965). 20. RA. Sack, “Extension of Griffith’s Theory of Rupture to Three Dimensions,” Proc. Phys. Sci. 58 a, 729-36 (1946). 21. T. K. Gupta, “Strength degradation and crack propagation in thermally shocked A1203,” J. Am. Ceram. Soc. 55 [5], 249-253(1972). 22. W. P. Rogers, A. F. Emery, R. C. Bradt, and A. L. Kobayashi, “Statistical Study of Fracture of Ceramic materials in the Water Quench Test,” J. Am. Ceram. Soc 70[6], 406-12 (1987). 23. M. Ashizuka, T. E. Easler, and R. C. Bradt, “Statistical Study of Thermal shock damage of Borosilicate Glass,” J. Am. Ceram. Soc. 66 [8], 542-550 (1983). 24. S. Schon, H. Prielipp, R Janssen, J. Rodel, and N. Claussen, “Effect of Microstructual Scale on Thermal Shock Resistance of Aluminum-Reinforced Alumina,” J. Am. Ceram. Soc. 77 [3], 701-704 (1994). 25. J. H. Ainsworth and R. E. Moore, “Fracture Behavior of Thermally Shocked Aluminum Oxide,” J. Am. Ceram. Soc. 52 [11], 628-29 (1969). 26. L.T. Herkkovich and M.V. Hyer, “Damage -induced Property Changes in Composite Subjected to Cyclic Loading,” Eng. Fract. Mech. 25, 779-791 (1986). 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 151 J. H. Ainsworth and R. H. Herron, “Thermal Shock Damage Resistance of Refractories,” J. Am. Ceram. Soc. Bulletin 53 [7] 628-29 (1974). C. E. Semler and TH. Hawisher, “Evaluation of the Thermal Shock Resistance of Refractories Using the Ribbon Test Method,” J. Am. Ceram. Soc. Bulletin 59 [7], 732-38 (1980). Y. Kim, W. J. Lee, and E. D. Case, “Thermal Fatigue Behavior of Ceramic Matrix composite: A Comparison among Fiber Reinforced Whisker Reinforced, Particulate Reinforced, and Monolithic Ceramics”, in Proceedings of the American Society for Composites, 5th Technical Conference, Technomic Publications, Lancaster 871-881 (1990). W. J. Lee, “Thermal Fatigue in Ceramics and Ceramic Matrix Composites”, Ph.D. Dissertation, Michigan State University, East Lansing, MI (1991). K. Lee, Personal Communication NSC 820 Class Notes, 1995, Graduate Student, Materials Science and Mechanics, Michigan State University. C. E. Ash, “Cyclic Thermal Shock-Crack Growth in Alumina,” Senior Research Project, Michigan State University. (1994). Y. Kagawa, “Quantitative Analysis of Closure Stress-Crack Separation Curve in Grain Bridge Toughening of Polycrystalline Ceramics, Mat. Sci. and Eng., A176 379- 383 (1994). S. Lathabai, J. Rodel, and B. R. Lawn, “Cyclic Fatigue from Degradation at Bridging Grains in Alumina,” J. Am. Ceram. Soc., 74 [6] 1340-48 (1991). J. C. Hay and K. W. White, “Grain-Bridging Mechanism in Monolithic Alumina and Spinel,” J. Am. Ceram. Soc., 76 [7], 1849 -54 (1993). P. L. Swanson, C. J. Fairbanks, B. R. Lawn, Y. W. Mai, and B. J. Hockey, “Crack Interface Grain Boundary in Ceramics: 1, Experimental Study on Alumina,” J. Am. Ceram. Soc., 70 [4], 279-89 (1987). H. Kishmoto, A. Ueno, and S. Okawara, “Crack Behavior of Polycrystalline Alumina under Static and Cyclic Load,” J. Am. Ceram. Soc., 77 [5] 1324-28 (1994). R. H. Dauskardt, “Cyclic Fatigue-Crack Growth in Grain Growth in Grain bridging Ceramics,” J. of Eng. Mat. and Tech. 115, 244-251 (1993). R. H. Dauskardt, “A Frictional-Wear Mechanism for Crack Fatigue-Crack growth in Grain bridging Ceramics,” Acta Metall. Mater. 41, 2765-2781 (1993). 152 40. Y. W. Mai and B. R. Lawn, “Crack Interface Grain Boundary as a Fracture Resistance Mechanism in Ceramics: H, Theoretical Fracture Mechanics,” J. Am. Ceram. Soc., 70 [4], 289-94 (1987). 41 .Y. S. Chou and D. J. Green, “Silicon Carbide platelet/Alumina Composite: 1H, Toughening Mechanism,” J. Am. Ceram. Soc., 76 [8], 1985-92 ( 1993). 42. B. R. Lawn, “Microstructure, R-curve, and Strength of Monophase Ceramics,” Ceram. Eng. Sci. Proc. 11, 591-593 (1990). 43. R. W. Steinbrech, A. Reichl, and W. Achaarwachter, “R-curve Behavior of Long Cracks in Alumina,” J Am. Ceram. Soc., 73 [7], 2009-15 (1990). 44. G. Vekinnis, M. F. Ashby, and P. W. R. Beaumont, “R-curve Behavior of A1203 Ceramics,” Acta Metall. Mater. 38 [6], 1151-1 162 (1990). 45. L. M. Braun, S. J. Bennison, and B. R. Lawn, “Short Crack T-curves and Damage Tolerance in Alumina-Based Composites,” Ceram. Eng. Sci. Proc., 13, 156-163 (1992) 46. J. Rodel, J. F. Kelly, and B. R. Lawn, “In Situ Measurements of Bridged Crack Interaction in the Scanning Electron Microscope,” J. Am. Ceram. Soc., 73 [11] 3313- 18 (1990). 47. A. Reichl and R. W. Steinbrech, “Determination of Crack-Bridging Forced in Alumina,” J. Am. Ceram. Soc., 71 [6] C299-01 (1988). 48. R. H. Dauskrdt, D. B. Marshall, and R. O. Rktchie, “Cyclic Fatigue Crack Propagation in Magnesia-Partially-Sterilized Zirconia Ceramics,” J. Am. Ceram. Soc., 73 [4] 893-93 (1990). 49. Personal Communication MSM 960 Class Notes, 1994, Materials Science and Mechanics, Michigan State University. MICHIGAN STATE UNIV. LIBRARIES Ill1111111111llll“llIllW11111111111111”11111 31293014107282