.III'I'IIII - Il‘llllllllill ||||llllll‘lllllWW 3 1293 01051 783 This is to certify that the dissertation entitled Mechanical Twin Nucleation, Propagation And Mechanical Twinning During Creep Deformation In TiAl presented by Zhe Jin has been accepted towards fulfillment of the requirements for Ph.D. degree in Materials Science Major professor... ... Date 5 ’2 8v9¥ MSU is an Affirmative Action/Equal Opportunity Institution 0- 12771 LIBRARY Mlchigan State Unlverslty PLACE IN RETURN BOX to remove thle checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE Hr" ' :11, -.o 3 .‘J’a. @llw ll— M11 3337' -~ ". "w; -m‘.2M'_‘ '99:) m? a} “is MSUIeAn."' ‘ ‘ ' ’1 "fl MECHANICAL TWIN NUCLEATION, PROPAGATION AND MECHANICAL TWINNING DURING CREEP DEFORMATION IN TiAl By Zhe Jin A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Materials Science and Mechanics 1994 ABSTRACT MECHANICAL TWIN NUCLEATION, PROPAGATION AND MECHANICAL TWINNING DURING CREEP DEFORMATION IN TiAl By Zhe Jin Gamma titanium aluminide (TiAl) is a promising candidate material for high temperature and high performance applications because of its high strength, low density, good oxidation resistance, and excellent high temperature mechanical properties. In recent years, much work has been done in order to understand its deformation mechanisms at room temperature and high temperatures. The experimental results have shown that mechanical twinning is an important deformation mechanism in TiAl both at room temperature and at high temperature. Even under creep deformation conditions the mechanical twinning still has a significant contribution to the creep deformation. Therefore, in order to understand the mechanical twinning behavior during creep deformation of TiAl, mechanical twin nucleation and propagation mechanisms and the mechanical twinning contribution to the creep deformation in TiAl were studied. Mechanical twin nucleation and propagation were in situ observed in creep deformed specimens using electron beam illumination in TEM. The mechanical twins nucleated either at grain boundaries or at twin interfaces due to the local stress concentration. The nucleus was either a superlattice intrinsic stacking fault that was a result of emission of one twinning dislocation from the grain boundary or the twin interface, or an extrinsic superlattice stacking fault resulting from the emission of two twinning dislocations. The twinning dislocation was identified to be Shockley partial 1/6< 115] and the twinning plane was identified as (111). The twinning propagation mechanism was a homogeneous glide of twinning dislocations on every adjacent twinning plane. The stress state necessary for twin propagation is that the principal tensile stress axis must be orientated in the vicinity of [SST] crystal orientation. The stresses in the twin layers are classified into three categories: forward stress, back stress and external stress, (the definition of each stress is referred to section 4.4). A stress analysis on a thin twin layer shows that both the forward stress and back stress are very large at the twin tip, but they decrease as the distance from the twin tip increases. At very large distances from the twin tip, both stresses tend to be constant. The external stress is constant along the twin layer investigated. Both true-twinning and pseudo-twinning configurations are observed in creep deformed specimens. The analyses of mechanical twin (true-twin and pseudo-twin) formation during creep of TiAl show that mechanical twinning in TiAl obeys the maximum resolved shear stress criterion. Stress concentrations at grain triple points are found to be accommodated by formation of fine mechanical twins. These fine mechanical twins result in zigzagged grain boundaries at their intersections with the grain boundaries and hence inhibit the grain boundary sliding. The formation of fine mechanical twins are probably controlled more by the local stress concentration than by an externally applied stress. Copyright by ZHEJIN 1994 To My Parents Zhong-Zhi J in and Yu-J in Cui ACKNOWLEDGENIENTS I sincerely appreciate the help of my major professor, Dr. Thomas R. Bieler. He has been a good friend and has provided me with an effective and encouraging environment to finish this research. I am also very grateful to my graduate committee, Prof. Martin A. Crimp, Prof. David Grummon, Department of Materials Science and Mechanics, and Prof. Carl L. Foiles of Physics and Astronomy Department. They have provided many effective suggestions in selecting the research topic and the facilities necessary to pursue this research. Particularly, I would like to express my sincere appreciation to Howmet Corporation, Whitehall, Michigan for the financial and material support to make this research possible. I would like to express my thanks to Dr. B. London, Dr. T. Thom, Dr. NE Paton, Dr. D.A. Weeler for their contributions in pursuing this research. I would also like to express my gratitude to my wife, Jiyan An, for her patient support and encouragement during my entire graduate study. vi TABLE OF CONTENTS LIST OF TABLES ........................................................................ xiv LIST OF FIGURES ...................................................................... xvi CHAPTER ONE. INTRODUCTION ................................................. 1 CHAPTER TWO. THEORY ........................................................... 7 2.1. Titanium Aluminide TiAl ......................................................... 7 2.1.1. Crystal Structure ........................................................ 7 2.1.2. Stacking Faults .......................................................... 9 (a) APB (Antiphase Boundary) ........................................ 11 (b) SISF (Superlattice Intrinsic Stacking Fault) .................... 13 (c) CSF (Complex Stacking Fault) ................................... 13 (d) SESF (Superlattice Extrinsic Stacking Fault) .................. 14 2.1.3. Dislocations and Dislocation Core Structures ...................... 14 (a) 1/2 <110] Type ..................................................... 15 (b) <101] Type ......................................................... l6 vii (c) 1/2 <112] Type .................................................... 16 (d) <100> Type ....................................................... 17 2.1.4. Dislocation Blocking Mechanisms ................................... 17 (a) Roof-Type Blocking ................................................ 18 (b) Kear-Wilsdorf Blocking ........................................... 20 (0) Deep Peierls Valley Blocking .................................... 21 (d) SSE—Tube Type Blocking .......................................... 22 (e) Dislocation Interaction Blocking ................................. 26 2.1.5. Lamellar Structure and Its Formation ............................... 28 (a) Type-I Lamellar Structure ......................................... 28 (b) Type-II Lamellar Structure ........................................ 29 2.1.6. Phase Transformation Mechanisms .................................. 31 2.2. Mechanical Twinning Theory .................................................... 33 2.2.1. Definition ................................................................. 33 2.2.2. Seven Twinning Classes ............................................... 35 2.2.3. Classic Twinning Modes ............................................... 37 2.2.4. Crystallography of Mechanical Twinning ........................... 38 2.2.5. Atomic Movement in Twinning Shear .............................. 45 2.2.6. Strain in Twinning ...................................................... 48 2.2.7. Prediction of Twinning Elements (Twinning Criterion) .......... 52 2.2.8. Twinning in Superlattice ............................................... 53 viii 2.2.9. Mechanical Twin Nucleation and Propagation ..................... 61 (a) Dislocation Pole Mechanism in bcc Metals ................... 62 (b) Dislocation Pole Mechanism in hcp Metals ................... 63 (c) Dislocation Pole Mechanism in fcc Metals .................... 66 2.2.10. Twin Shape ............................................................. 68 (a) A Model Concerning Energy Associated With Twinning .................................................. 68 (b) A Model Based on Twinning Dislocation Interaction Forces ............................................... 72 2.3. Creep Theory ........................................................................ 75 2.3.1. Microstructural Correspondences to Creep ......................... 76 2.3.2. Mechanical Model ...................................................... 77 2.3.3. Microstructural Model ................................................. 80 CHAPTER THREE. MECHANICAL TWIN NUCLEATION AND PROPAGATION IN TiAl ............................... 82 3.1. Materials and Experimental Procedure ......................................... 83 3.2. Results and Analyses ............................................................... 85 3.2.1. In Situ Observations of Mechanical Twin Nucleation and Propagation ........................................... 85 (a) Twin Nucleation at Grain Boundaries .......................... 85 (b) Twin Nucleation at Twin Interfacies ........................... 100 (c) Twin Layer Morphology Near The Twin Tip ................. 107 3.2.2. Post-Mortem Observation of Mechanical Twinning During Creep Deformation ........................................... 107 3.3. Discussion ........................................................................... 113 3.3.1. Mechanical Twin Nucleation in TiAl ................................ 113 3.3.2. Mechanical Twinning Mechanisms in TiAl ......................... 116 3.3.3. On The DO19 phase in Fine Mechanical Twins .................... 121 3.3.4. Twin Morphology ....................................................... 122 3.4. Summary ............................................................................. 123 CHAPTER FOUR. FORCE AND STRESS ANALYSES IN A THIN TWIN LAYER ..................................... 126 4.1. Theory ................................................................................ 127 4.1.1. Forces on Each Twinning Dislocation ............................... 127 4.1.2. Dislocation Interaction Forces ........................................ 131 4.1.3. Internal Friction Force ................................................. 134 4.1.4. The Force Due to The Stacking Faults ............................. 136 4.1.5. Applied Force ........................................................... 137 4.1.6. Definitions of External Force, Forward Force and Backward Force ................................................... 138 4.2. Twinning Dislocation Distribution Within A Thin Twin Layer ............ 139 4.3. Force Calculation .................................................................. 144 4.3.1. Calculation of Forward Force Ff ..................................... 144 4.3.2. Calculation of Backward Force Fb ................................... 148 4.3.3. Calculation of External Force Fa .................................... 155 4.4. Stresses in The Thin Twin Layer ................................................ 157 4.5. Simplified Equations for Ff, F, and F“ ......................................... 162 4.6. Discussion ........................................................................... 168 4.6.1. Stress and Force Distributions Within The Twin layer .......... 168 (a) External Stress 1,, ................................................... 168 (b) Forward Stress r, .................................................... 169 (0) Back Stress 1,, ........................................................ 170 4.6.2. Dislocation Distribution in The Case of Equal Backward Force on Each Twinning Dislocation ................... 175 4.6.3. Effect of Dislocation Location on The Stress Distribution ....... 176 4.7. Summary ............................................................................. 178 CHAPTER FIVE. MECHANICAL TWINNING DURING CREEP DEFORMATION IN TiAl ......................... 183 5.1. Introduction .......................................................................... 183 5.2. Material and Experimental Procedure .......................................... 184 5.3. Results ................................................................................ 186 5.3.1. Optical Microstructure ................................................. 186 5.3.2. Cross Twinning and Parallel Twinning in Lamellar Grains ..... 187 5.3.3. Fine Mechanical Wins at Grain Triple Points .................... 190 5.4. Analysis and Discussion ........................................................... 194 5.4.1. Cross Twinning in Lamellar Grains ................................. 194 5.4.2. Cross Twinning in Equiaxed ‘y Grains .............................. 202 5.4.3. Fine Mechanical Twins at Equiaxed 7 Grain Triple Points ..... 203 xii (a) Fine Mechanical Twins Formed by Accommodation of Stress Concentration at Equiaxed 7 Grain Triple Points ........................................................ 203 (b) Formation of Fine Mechanical Twins Prevents Grain Boundary Sliding .......................................... 204 (c) On DO“, Crystal Structure in Fine Mechanical Twins ...... 207 5.5. Conclusions .......................................................................... 209 CHAPTER SIX. CONCLUDING REMARKS AND RECOMMENDED FUTURE WORK ......................... 211 LIST OF REFERENCES ................................................................ 214 xiii Table 2.1. Table 2.2. Table 2.3. Table 2.4. Table 2.5. Table 2.6. Table 2.7. Table 3.1. Table 3.2. Table 3.3 Table 4.1. Table 4.2 Table 4.3 Table 4.4 Table 4.5 Table 5.1. Table 5.2 LIST OF TABLES Dislocation behavior in TiAl at different temperature ranges ...... 18 Twinning modes in cubic lattices ........................................ 36 Twinning modes in cubic lattices ........................................ 58 Possible true twinning modes in cubic superlattices .................. 60 Twinning modes in non-cubic superlattices ............................ 61 Twinning modes in hcp metals ........................................... 66 Creep Deformation Mechanisms At Temperature T/Tm 2 0.4 79 Composition of the 'y TiAl specimen ................................... 83 The mechanical twinning elements in Fig. 3.1 ........................ 93 Schmid factors on all possible twinning systems within the (111) plane ........................................................... 93 The distance of twinning dislocations from the twin tip ............. 143 The forces on each twinning dislocation ............................... 149 The stresses at each twinning dislocation .............................. 160 The forces calculated using simplified equations ..................... 166 The modified stresses and forces for the first three twinning dislocations .................................................... 171 Schmid Factors for Mechanical Twinning Systems in 7/7 Lamellae .......................................................... 199 Misorientation and Grain Boundary Structure of Grains at the Grain Triple Point ....................................................... 205 xiv Figure 2.1 - Figure 2.2 - Figure 2.3 - Figure 2.4 - Figure 2.5 - Figure 2.6 - Figure 2.7 - Figure 2.8 - Figure 2.9 - Figure 2.10 - Figure 2.11 - Figure 2.12 - LIST OF FIGURES Unit cell of superlattice L10 .................................................... 7 Phase diagram of TiAl .......................................................... 8 Stacking fault configurations as viewed along [1T0]. (a) APB, (b) SISF, (c) CSF, (d) SESF(I), and (e) SESF(II) ......................... 9 Crystal projection as viewed along [111]. bA is a crystal vector which results in an APB, bC results in a CSF, and bS results in a SISF ........................................................................... 12 Thompson tetrahedron. The primed letters indicate the antisites with respect to the unprimed letters ........................................... 12 Dislocations in superlattice Ll0 ................................................. 15 Glissile configurations of superdislocations. The single line is a SSF, the jagged line is an APB, and the double line is a CSF ........ 19 Roof-type blocking configurations. The single line is a SSF, the jagged line is an APB, and the double line is a CSF .............. 19 Kear-Wilsdorf blocking configurations. The dashed line is the APB in the {001} plane, and the single line is the SSF in the {1 1 1} plane ........................................................................ 20 The Peierls relief along different crystallographic directions in TiAl ............................................................................. 22 Schematic diagram of the row of broken Ti-Ti bonds .................... 23 Formation of SSF—tube blocking. (a) Formation of nonaligned jogs on a superdislocatiog; (b,c) formation of a dipole; (d) a SSF—tube along [101] direction .......................................... 24 XV Figure 2.13 - Figure 2.14 - Figure 2.15 - Figure 2.16 - Figure 2.17 - Figure 2.18 - Figure 2.19 - Figure 2.20 - Figure 2.21 - Figure 2.22 - Figure 2.23 - Figure 3.1 - Figure 3.2 - Definition of orientation variants in (‘y + 02) lamellae ..................... 30 Relations between ci and pi .................................................... 39 The sign of four twinning elements .......................................... 41 The rationality of K1 in type-I twinning ..................................... 42 Schematic illustration of lattice shuffling during type-I twinning for the of q = 4 .................................................................. 46 lattice point displacement during twinning shear ......................... 49 The extension and contraction of sample length for twinning in zinc .............................................................. 51 The structure and interchange shuffles during (120) twinning in D0, .............................................................................. 5 6 Dislocation pole mechanism for twinning in bcc lattice .................. 64 Q02) twinning in zirconium. Projection of the lattice on the (1210) plane. Circles are in plane of the paper. Squares are a/2 above and below the paper. Solid symbols indicate atom positions in the twin ............................................................. 65 Dislocation pole mechanism in fee lattice ................................... 67 Mechanical twinning propagation configuration which was observed in situ in TiAl: (a) the tilted image of fine mechanical twins within a grain interior; (b) the image of fine mechanical twins viewed parallel to the twinning plane; (c) the diffraction pattern across several fine twin layers; and (d) indexes of the diffraction pattern which shows a twin relationship across the twin-matrix interfaces ........................................................... 86 Schematic diagram of mechanical twinning nucleation and propagation procedure in TiAl, which shows that the mechanical twinning nucleates at a grain boundary and propagates into the grain interior. (a) and (b) are the configuration of mechanical twinning nucleation at the grain boundary with the emission of one twinning dislocation from the grain boundary; (c) and (d) are the same configuration as (a) and (b) except that two twinning dislocations are emitted from the grain boundary simultaneously; xvi Figure 3.3 - Figure 3.4 - Figure 3.5 — Figure 3.6 - Figure 3.7 - Figure 3.8 - Figure 3.9 - (e) is an intermediate state of twinning propagation; and (f) is the final state in which the twinning propagation ceases in the grain interior as twin layers "A" and "B" in Fig. 4.1 (a) ................ 89 Schematic diagram of the incoherent twin boundary structure in terms of individual twinning dislocations, which is viewed in [110] direction. The boundary between the first dislocation and the second dislocation at the twin tip is an intrinsic stacking fault, the boundary between the second and the third dislocations is an extrinsic stacking fault, and the boundary between any two adjacent dislocations after the third dislocation is a twin plane. The twinning dislocations are denoted by numbers: the leading dislocation at the twin tip is counted as the first dislocation, the one just behind it is the second one, and so on ............................ 92 Crystal orientation relationships of the twinning elements and the possible twinning directions in (111) plane. The cross mark indicates the direction of creep tensile axis ................................. 94 Schmid firctor contours with respect to the twinning system of (111)[1 12], which indicates that a tensile principal stress should be in the vicinity of [551] crystal direction in order to operate the (111)[112] twinning system ................................................ 96 Mechanical twinning propagation mechanism in TiAl, which shows the twin layers resulting from the glide of (a) one, (b) two, (0) three, and ((1) four twinning dislocations in the (111) twinning plane during the twinning propagation. The left side column is the glide sequence of l/6[112] twinning dislocations. The middle cglumn is the atomic arrangements of resultant twins viewed in [110] direction, in which the open circles indicate the atoms located in the plane of the paper and the shaded squares indicate the atoms in the plane just beneath the paper. The right side column is the atomic stacking sequences along [111] direction before and after twinning, in which the primed letters indicate the sheared planes ................................................................ 97 The sequences of mechanical twinning nucleation and propagation in TiAl ............................................................................. 101 (a) The side view of twin layers A, B, C, and D in Figure 4.7, and (b) the diffraction pattern taken across these twin layers ............ 104 The thin twin morphology near the twin tip ................................ 108 xvii Figure 3.10 - Figure 3.11 - Figure 4.1 - Figure 4.2 - Figure 4.3 - Figure 4.4 - Figure 4.5 - Figure 4.6 - The configuration of fine mechanical twins resulting from the accommodation of local stress concentration at a grain triple point (a). Diffraction pattern across the interfaces (b) shows that these fine laths are twin-related with matrix and the existence of fine DO19 phase between them (0). Diffraction pattern taken in untwinned area within the same grain without tilting the sample ((1) shows that the untwinned region has the same crystal orientation as the matrix of fine mechanical twins as shown in (b). (e) is the indexes of (d). It also shows that the growth of fine mechanical twins as pointed out with an arrow between the completely developed fine twins and the matrix stopped in the grain interior .................................................................... (a) shows three thin twin layers indicated by letters "A", "B" and "C" within a y grain interior, which indicates that the nuclei of these twin layers were formed inside the 7 grain; (b) is a diffraction pattern taken across the twin layer ................... Forces on individual twinning dislocation in the thin twin layer. (a) On the first twinning dislocation, (b) on the second one, (c) on the third, and (d) on the fourth ...................................... (a) Variation of Peierls energy as a function of transverse displacement (u), (b) Variation of the lattice friction stress. (w, :5 b2) ..................................................................... (a) is the amplified image of a thin twin layer and (b) is a schematic drawing of the twinning dislocation distribution in the thin twin layer ............................................................. The distance of twinning dislocations from the twin tip ................. The forces distribution along the twin layer ............................... Shear stresses distribution in the thin twin layer .......................... Figure 4.7 - Comparisons of results obtained from the original equations with Figure 4.8 - The modified stress distribution along the thin twin layer ............... Figure 4.9 - The modified force distribution along the thin twin layer ................ results obtained from the simplified equations ............................ xviii .. 110 119 .. 129 .. 135 .. 141 145 .. 150 . 161 .. 167 172 173 Figure 4.10 - Figure 4.11 - Figure 5.1 - Figure 5.2 - Figure 5.3 - Figure 5.4 - Figure 5.5 - Figure 5.6 - Figure 5.7 - Figure 5.8 - Figure 5.9 - Figure 5.10 — Figure 5.11 - Dislocation distribution at equal backward force on every twinning dislocation. (Fb/Gb = 1.18 x 10'2) ................................ 177 Effect of the twinning dislocation location on the backward stress. (a) For the second dislocation, (b) for the third dislocation, and (c) for the tenth dislocation .................................................... 179 Optical micrographs of (a) initial microstructure and (b) after creep deformation ................................................................ 186 (a) Cross twinning configuration in a lamellar grain, (b) diffraction pattern of the original lamellae, and (c) orientations of the original lamellae ............................................................................ 188 Cross twinning configuration within a large equiaxed 7 grain .......... 191 Fine mechanical twin configurations at equiaxed 7 grain triple points ............................................................................... 192 Fine mechanical twins (a), diffraction pattern across several fine twin interfaces (b), and twin relation in the diffraction pattern (0) ..... 193 Atomic arrangement of true-twinning (a) and pseudo-twinning (b) ..... 196 Stereographic projection of possible twinning systems along [110] direction .................................................................... 198 Schematic mode of cross twinning: (a) initial condition of two lamellae; (b) metastable condition of twinning in lamellae 1; (c) final configuration of cross twins as seen in Fig. 5 .2 (a) ............ 200 Fine mechanical twins from accommodation of local stress concentration (a), twin end configuration at equiaxed 7 grain boundary (b), and dislocation slip near zigzagged boundary (c) ........ 206 Boundary between an equiaxed 7 grain and a lamellar grain ........... 208 Optical microstructure showing the roughness of lamellar grain boundaries ......................................................................... 208 CHAPTER ONE INTRODUCTION Gamma titanium aluminide (TiAl) is a promising candidate material in high temperature and high performance applications for its high strength, low density, good oxidation resistance, and excellent high temperature mechanical properties. The main barrier for application in practice is its brittleness at room temperature “'3'. In the last decade, most of the research effort in TiAl has been concerned with fundamental studies of dislocation structures in single phase materials [“0'. More recent studies on Ti-rich alloys consisting of two phases (7+a2) which present better room temperature ductility have also been focused mainly on the dislocation structures “”6'. However, experimental results have shown that mechanical twinning is an important deformation mechanism in TiAl both at room and high temperatures “7'2”. Particularly, under creep deformation conditions, mechanical twinning contributes significantly to the creep deformation “’23‘3”. Titanium aluminide TiAl has a L10 ordered face centered tetragonal structure. The mechanical twins observed are of the {111} < 112]' type which are the variants in "‘ The notations " < " and " > " in the indexes indicate that the two indexes nearest to " < " or " > " are permutable one to another without changing its properties. 2 which the L10 superlattices are mechanically twinned without disturbing the L1o crystal structure. It has been reported that when the L10 structure is homogeneously sheared by the variants other than {111} < 112] type, the L10 structure will be changed into a L1l superlattice structure “1. Based on this concept, the twinning operation by the variants other than {111}<112] is thought to be forbidden ”'3‘. The details of mechanical twinning behavior and its contribution to the deformation in TiAl are not completely understood at present. No fundamental investigations on mechanical twinning in TiAl have been carried out until the very recent work done by Farenc, Coujou and Couret [32‘ on twin propagation in TiAl, by Wardle, Phan and Hug ‘33] and by Sun, Hazzledine and Christian [3‘5" on twin intersections. Farenc, Coujou and Couret first investigated in situ mechanical twin propagation at room temperature and at 400 °C. They found that " twins are formed by a/6<112] partial dislocations gliding in octahedral planes" and the twin propagation is due to "a partial dislocation turning around a perfect dislocation". However, the material used in their study was an aluminum rich Ti..,,Als4 single 7 phase alloy. The basic properties of Al-rich alloys (single 7 phase) are inherent brittleness at room temperature “‘11", but the 7 phase in the 7+0:2 lamellar structure has been found to be intrinsically ductile ‘3‘”. Therefore, the mechanical twinning behavior and its role in the deformation in Al-rich single phase alloys may differ from those in Ti-rich two phase alloys. This is because the stacking fault energys of 7 phase in Al-rich single 7 phase and in Ti-rich 7+0:2 lamellar grain are different [3°]. The twin intersection has been comprehensively studied by Wardle, Phan and Hug [33‘ for both Al-rich and Ti-rich 3 alloys at room temperature and by Sun, Hazzledine and Christian [34’3“ for an alloy of stoichiometric composition (50Ti - 50A] at. %) in the temperature range -196 °C to 900 °C, respectively. In Wardle, Phan and Hug’s work ‘33', two different twin intersections were observed and it was found that the structures of twin intersections depended on "the orientation of the common direction of the twin habit planes". Similarly, in Sun, Hazzledine and Christian’s work ”4”“, two types of twin intersections have been found: type I intersection occurs along < 110] directions while type II intersection occurs along <011] directions. In the type I intersection, two different configurations are observed depending on the deformation temperatures: (i) At room temperature, one twin band is deflected and the other remains rectilinear. The lattice in the intersection region remains in L10 structure and is congruent with the lattice of the deflected twin; (ii) at high temperature, both twin bands are deflected. In type II intersection, both twin bands are deflected at both room and high temperatures. Type II intersection has been found to be associated with 'A < 110] dislocations in the twin-matrix interface. However, all these fundamental studies on mechanical twinning have been carried out on the specimens deformed at common tensile testing strain rates, for example, the strain rate in [34] was 5 x 105 s", which is much faster than normal creep strain rates. Therefore, these observations may not be directly applicable to creep deformation. For mechanical twinning in creep deformation, very limited research has been reported [22. 2”“37'39'. Also, there are two differing results concerning twinning in creep deformation of TiAl. Loiseau and Lasalmonie [22' investigated creep deformation of 4 equiaxed single phase 7 Ti46A154 and found that twinning was an important creep deformation mechanism at deformation temperatures up to 800 °C. 1 in and Bieler [28'3” have found that mechanical twinning is significant in creep deformation at 765 °C for Ti- rich alloy (Ti-48Al—2Nb-2Cr). However, Huang and Kim [39‘ studied creep behavior of two phase 7+a2 alloy with composition of Ti-47.0Al-1.0Cr—1.0V-2.5Nb at 900 °C and observed no evidence of twinning in creep deformation. Pseudo-twinning, which was thought to be "forbidden" in the literature ”5"“ and it will be discussed later, was found in a creep deformed TiAl specimen '2‘”. In addition, the creep deformation mechanisms are not clear at present. The activation energies for creep of TiAl are much larger than those for self-diffusion and interdiffusion in TiAl [39:41:42], which indicate that the creep rate may be controlled by some processes other than diffusion. The reported values of the stress exponents vary widely from about 2 to 8 [434"]. This indicates that several deformation mechanisms may be involved in creep deformation of TiAl. Mechanical twinning has been found to be an important creep deformation mechanism [22. "*3" 37'3”, but mechanical twinning is influenced very little by diffusion processes, and dislocation configurations in creep specimens are similar to those observed in short term tensile specimens “'9‘. All these indicate that mechanical twinning may play an important role in the creep deformation of TiAl. Based on the above circumstances, investigations of twin nucleation and propagation in creep deformed TiAl and the contributions of mechanical twinning to 5 creep deformation of TiAl have been carried out in this study. At first, the basic understanding of TiAl, the intensive twinning theory and a brief introduction of creep theory has been summarized in Chapter Two. In Chapter Three, in situ observations of mechanical twin nucleation and propagation in a creep deformed specimen are presented. The mechanical twinning was observed to nucleate either at grain boundaries or at twin interfaces due to the local stress concentration. The nucleus was either a superlattice intrinsic stacking fault which resulted from the emission of one twinning dislocation from the grain boundaries or the twin interfaces, or a superlattice extrinsic stacking fault resulting from the emission of two twinning dislocations. The twinning dislocation was identified to be Shockley partial 1/6 <11’2’] and twinning plane was identified as (111). The twinning propagation mechanism was a homogeneous glide of twinning dislocation on every adjacent twinning plane. The stress state necessary for twinning propagation is that the principal stress axis must be orientated in the vicinity of [SST] crystal orientation in the case of tensile stress state. The force and stress analysis on a thin twin layer is presented in Chapter Four. The result shows that both the forward stress and the back stress are very large at the twin tip and decrease quickly as the distance from the twin tip increases. At very large distances from the twin tip, both stresses tend to be constant. The external stress on the investigated twin layer was found to be a residual stress in the matrix. In the case of no external load and small residual stress, the twin propagation is controlled by the emission of twinning dislocations. 6 In Chapter Five, some characterizations of mechanical twins in creep deformed specimens are presented. The results show that mechanical twins are formed by either true-twinning or pseudo-twinning. The mechanical twinning in creep deformation of TiAl obeys the maximum resolved shear stress criterion. Stress concentrations at grain triple points can be accommodated by forming fine mechanical twins. The zigzagged grain boundary formed by mechanical twinning can inhibit grain boundary sliding. The formation of fine mechanical twins are probably controlled more by the local stress concentration than by an externally applied stress. In the last chapter, Chapter Six, some general concluding remarks and some recommended future work are stated. CHAPTER TWO THEORY Some basic understanding of titanium aluminide TiAl, mechanical twinning theory and creep theory will be summarized in this chapter. 2.1. Titanium Aluminide TiAl 2.1.1. Crystal Structure Titanium aluminide TiAl has a Llo ordered face centered tetragonal structure with a composition range of 49-66 atomic percent of aluminum, which varies depending on [47491 temperature . The Titanium and aluminum atoms alternately stack in (002) plane, o f 1 b I ———-—— - —___..__. Figure 2.1 — Unit cell of superlattice L10. 8 as shown in Fig. 2.1. At the stoichiometric composition, the c/a ratio is 1.02 and the tetragonality increases up to c/a= 1.03 with increasing aluminum concentration and decreases to 1.01 with decreasing aluminum concentration [5°52]. At off stoichiometric composition, excess titanium or aluminum atoms occupy antisites without creating vacancies ”3‘. The TiAl phase remains ordered up to its melting temperature of about 1450 °C as shown in Fig. 2.2 ‘49], which shows a central portion of the equilibrium Ti-Al phase diagram. Weight Percent of A1 20 25 30 35 40 45 50 I l l l l I l 1600 ". L _ 2900 ; f3 1400 — [3 E? ' Q, - 2500 L l g 01 - 2300 E 8 1200 - a 8, / \ 2100 8. E E O) E2 1000 _ 02 1900 H _ 1700 800 — 1500 l 1 l l 1300 30 40 50 Atomic Percent of Al Figure 2.2 - Phase diagram of TiAl. 2.1.2. Stacking Faults Since TiAl has L10 structure with ratio c/a=1.02, it is very close to the fcc structure. Like the fcc crystal, the geometric stacking sequence along <111> direction is ABCABCABCABC There exist four types of stacking faults in the TiAl crystal structure, i.e., they are antiphase boundary (APB), superlattice intrinsic stacking fault (SISF), complex stacking fault (CSF) and superlattice extrinsic stacking fault (SESF). The atomic configurations of these stacking faults are shown in Fig. 2.3, which are viewed along [1T0] direction. [111] —A' -B' -C' l u>nu Figure 2.3 - Stacking fault configurations as viewed along [1T0]. (a) APB, (b) SISF, (c) CSF, (d) SESF(I), and (e) SESF(II). 10 (Figure 2.3 continued) 11 Ill can: I >nw> (Figure 2.3 continued) (a) APB (Antiphase Boundary) The APB is created by the glide of a slip vector D’A (or CA), which is shown in Fig. 2.4 by a crystal vector bA. Fig. 2.4 shows an atomic arrangement in a (111) crystal plane where the stacking sequence is labeled with a, b, c. The slip vector D’A (or CA) is shown in Thompson tetrahedron in Fig. 2.5. In this case the stacking sequence across the APB changes into the following one: ABCABCIA’B’ C’ A’B’ C’ where the primed letters indicate the displaced planes with respect to the unsheared crystal that is indicated by letters without primes (it will be the same in the following). 12 [0“] [112] [101] 1 [1511 —->- [110] Figure 2.4 - Crystal projection as viewed along [111]. bA is a crystal vector which results in an APB, bC results in a CSF, and b3 results in a SISF. Dr L__ Figure 2.5 - Thompson tetrahedron. The primed letters indicate the antisites with respect to the unprimed letters. 13 It is more visible if one looks at the shear displacement in the [ITO] direction, which is shown in Fig. 2.3. In Fig. 2.3 (a), one can easily find the antiphase boundary configuration which lies in the (111) plane. (b) SISF (Superlattice Intrinsic Stacking Fault) The SISF is very common in TiAl. The SISF is formed by the glide of a slip vector 3A in Fig. 2.5 or by a crystal vector bS in Fig. 2.4. If one looks at the stacking sequence order, the SISF is formed by taking off one plane from the original stacking sequence as the following: ABCABCBCABC In Fig. 2.3 (b), it can be seen that the atoms in the upper half of the crystal shift to the left with respect to the lower half in an amount of BA. (c) CSF (Complex Stacking Fault) The CSF is created by the glide of HD’ (or BC’) in Fig. 2.5, which is shown in Fig. 2.4 with a crystal vector be. The stacking sequence after shearing is ABCABmerxnwr From Fig. 2.3 (c), it is found that it has an antiphase boundary feature in addition to Fig. 2.3 (b) like configuration. Therefore, the CSF energy may be equal to APB energy plus 14 SISF energy. ((1) SESF (Superlattice Extrinsic Stacking Fault) There are two possibilities to form the SESF. (i) The SESF is formed by the glide of two BA in successive close-packed planes (type I). This type SESF looks like the effect of inserting one extra plane between two original planes, I ABCABCBABCABC which is shown in Fig. 2.3 ((1). (ii) The SESF is formed by the glide of slip vectors BD’ and BC’ (type II). The type H configuration is shown in Fig. 2.3 (e). Comparing type I and type II configurations in Fig. 2.3 (d) and (e), it is easily found that type I is more stable than type 11 because type 11 includes an APB feature. 2.1.3. Dislocations and Dislocation Core Structures There are four types of dislocations in TiAl due to its ordered crystal structure, that is, 1/2 < 110] type, < 101] type, 1/2 <112] type and <100> type, as shown in Fig. 2.6. Here 1/2 < 110] dislocations are normal dislocations but both < 101] and 1/2 < 112] dislocations are superdislocations and < 100 > dislocations are cube dislocations. Notations " < " and " > " in the indexes indicate that the two indexes nearest to " < " or " > " are permutable one to another without changing its properties. The dislocation core structures are very complicated and they depend on which plane they dissociate into [011] 15 1/21112] ‘ O Afr-72+. ------------ ”-i-f-I-i-i-3 y 1/2[110] Figure 2.6 - Dislocations in superlattice L10. ”“5”. The core structure dependence upon temperature is closely associated with the temperature-dependent deformation behavior in TiAl 121. The typical dislocation dissociation reactions of these four dislocations are listed in the following: (a) 1/2 <110] Type D’C’ --—> D’B + CSF + BC’ 1/2[110] ---> 1/6[211] + CSF + 1/6[12I] [3 Note that the marks " D. c; C 30 is CSF, is SISF, is APB, is SESF, is no fault. ~ 16 (b) <1011TYP6 (i) 2 DA ---> D’A + APB + D’B + srsr + BA ---> D’B + csr + BA + APB + D’B + srsr + BA [101] ---> 1/2[101] + APB + 1/6[211] + srsr + 1/6[1I2] ---> 1/6[211] + CSF + 1/6[1I2] + APB + 1/6[211] + SISF + l/6[1I2] D'A D'E ‘ BA 1313 3A 1313 3A O———O==OVWVV\AO (ii) 2 D’A ---> D’C’ + 3 BA ---> D’C’ + BA + SESF(I) + 2 BA ---> D’C’ + 2 BA + SESF(II) + (D’B + C’B) [101] ---> 1/2[110] + 1/2[1i2] ---> 1/2[110] + 1/6[1I2] + SESF(I) + 1/3[1'1'2] ---> 1/2[110] + 1/3[112] + SESF(II) + (1/6[211] + 1/6[El]) DC 3 BA D'C' BA 2 BA D'C 2 13A new]: CD ------ O O -------- W o -------- W (c) 1/2<112] Type (1)3 BA ---> 2 BA + SISF + BA 1/2[1I2] ---> 1/3[1i2] + srsr + 1/6[1I2] (ii) 3 BA ---> D’B + CSF + BA + APB + one + SISF + BA 1/2[11'2] ---> 1/6[211] + CSF + 1/6[1I2] + APB + 1/611'2'1] + SISF + 1/6[1I2] (iii) 3 13A ---> BA + SESF + 2 BA 1/211'1'21---> 1/6[1I2] + SESF + 1/31112] 2 BA BA on BA C’B BA BA 2 BA 213A BA D'B BA C‘B BA BA 213A 17 (d) <100> Type D’a’ ---> D’C’ + AB ---> D’fi + CSF + BC’ + A6 + CSF + (B where "*" is not a Thompson notation. [100] ---> 1/2[110] + 1/2[1I0] or -—-> 1/6[211] + CSF + 1/6[12I] + 1/6[I2I] + CSF + 1/6[211] DC‘ AB D'B BC A5 813 CD ------ O o———o ---------- o———o The dislocation slip planes in TiAl are close packed {l l 1} planes. According to the crystallography 0f TiAl, dislocations with 1/2 < 110] or <101] Burgers vectors can slip in two different {111} planes while dislocations with 1/2 (112] Burgers vectors can only slip in one {111} plane ‘5’. <100> dislocations may slip on the cube planes at very high temperatures (about 1000 °C) [5". It has been found that dislocation behavior is different at different temperatures. Dislocations observed in different temperature ranges are summarized in table 2.1. 2.1.4. Dislocation Blocking Mechanisms TiAl, like many intermetallic compounds, shoWs an anomalous yield stress- temperature dependence, that is, unlike disordered metals and alloys, the yield stress of TiAl increases as the temperature increases up to about 700 °C. The anomaly of yield stress to temperature results from the glissile-sessile transformations of dislocations in. different temperature ranges. The glissile dislocations have coplanar splitting 18 Table 2.1. Dislocation behavior in TiAl at different temperature ranges ‘5'°'°” Dislocation Unblocked Blocked Unblocked 1/2 < 110] -196 to 100°C 200 to 540°C >600°C 1/2 <112] -196 to 100°C 200 to 600°C >700°C (Not observed) <011] -196 to 300°C 400 to 600°C > 700°C < 100 > Not observed Not observed > 700°C configurations as shown in Fig. 2.7. The sessile dislocations have various configurations depending on their formation. Some dislocation blocking modes will be summarized in the following. (a) Roof-Type Blocking The roof type blocking results from a resplitting of superdislocations from their planar glissile configurations into two {111} noncoplanar sessile configurations (like a peaked root), as shown in Fig. 2.8 ”4’5". Of these configurations, the one containing SISF bands on both octahedral planes possesses the lowest energy compared to glissile and other roof-type configurations [2'5“0'. In addition, a planar sessile configuration was found for 1/2 <112] dislocations 19 <101> Dislocation 1/2<112> Dislocation G 4>-O=O C? . OVV~O=O 5C B5 SC B5 SC A5 5C B8 f M $ M 5C B5 BC 5C A5 BC 0 4: c2 4) 8C B6+BC 8C A8+BC Figure 2.7 - Glissile configurations of superdislocations. The single line is a SISF, the jagged line is an APB, and the double line is a CSF. 5,: \N... B6 01C 01C 8C B6 8C BC: 8C B5 c 1:3wa KM Figure 2.8 - Roof-type blocking configurations. The single line rs a SISF, the jagged line is an APB, and the double line is a CSF. p..— 20 [5'61]. It contains a SESF band bounded by partial dislocations with parallel Burgers vectors 1/6 < 112] and 1/3 < 112]. This configuration forms double layer twin in its core and is strongly blocked. (b) Kear-Wilsdorf Blocking The Kear-Wilsdorf blocking occurs as a result of cross slip of superdislocations into the cubic plane, as shown in Fig. 2.9 [59'6"]. Since the activation energy for the destruction of Kear-Wilsdorf blocking is very high, the transformation of Kear-Wilsdorf blocking to a glissile configuration is difficult. Taking 1/2<112] dislocation as an example, this dislocation may dissociate according to the reaction (C’A + C’B)‘C"" = C’A'C'“ + C’B‘C'A’, where the superscripts indicate the dislocation line directions. Therefore, the screw type C’A‘C'“ can glide in (010) cube plane, and the other remains blocked. The C’A‘C’“ partial gliding produces APB in (010) plane. Thus, Kear-Wilsdorf blockings of <011] or 1/2 <112] dislocations can not be destroyed by cubic slip “”6”. B01 8C B8 C} 9 a I 3 , C C O 1 (,2 3 8C B8 1 5c B8 ———.——— Figure 2.9 - Kear-Wilsdorf blocking configurations. The dashed line is the APB in the {001} plane, and the single line is the SISF in the {111} plane. 21 (c) Deep Peierls Valley Blocking The covalent nature of interatomic bonds in TiAl is the cause of deep (compared to metals) Peierls potential valleys. The deep Peierls valley blocking is based on the formation of directional Ti-Ti bonds along < 110] directions in [001] plane that contains Ti atoms. The slip planes {111} in gamma TiAl are equivalent, but the <1I0> directions on this plane are not equivalent. As a result, two types of directions exist: like atom direction and unlike atom direction as shown in Fig. 2.10 (a). Accordingly, two types of dislocation families exist: one-color—set dislocations (the dislocation lines lie along the rows of like atoms), and two-color-set dislocations (the dislocation lines are parallel to the rows of unlike atoms). The presence of a dislocation of any set gives rise to a row of broken bonds in the (111) slip plane as schematically drawn in Fig. 2.11. However, the other row of covalent bonds lying in the slip plane is parallel to the dislocation line for only the one-color—set. Fig. 2.10 (b) shows the relative Peierls valleys of these two directions. The one-color—set dislocations are in deep energy valleys, and hence are more stable than two-color—set ones which are in shallow valleys. The dislocations of different sets differ also in the structure of double kinks. For a one-color-set dislocation, the double kink consists of two equivalent kinks that are oriented in two-color-set directions; for a two-color—set dislocation, the double kink contains two nonequivalent kinks: one is in one-color—set direction and the other is in two-color-set direction. Therefore, a transformation from the shallow valleys into the deep valleys is thermodynamically preferred. At low temperatures, dislocations in deep valleys are blocked. As temperature rises until the stress peak, the above transformation 22 > X (a) (b) Figure 2.10 - The Peierls relief along different crystallographic directions in TiAl. is thermally activated and more dislocations are blocked in the deep valleys so as to increase the yield stress. Above the peak temperature, however, the deep valley blocking is thermally destroyed so that the flow stress drops rapidly. (d) SSF-Tube Type Blocking The SSF-tube (Superlattice Stacking Fault tube) blocking is also an effective blocking mode. The formation of such a configuration is schematically shown in Fig. 2.12 ‘9'. A slipping superdislocation intersects with the other dislocations to form a jog in the second plane. As a result of splitting, the jog transforms into two nonaligned jogs 23 ///Ae/ /{/// //// / I//// //I f /¢ 2 /% /// 7//////// % ////// W / \\\\\\\V \\\“ . Figure 2.11 — Schematic diagram of the row of broken Ti-Ti bonds. Al atoms are removed for clarity. The gray shaded plane is a terminated half plane of the dislocation and the cross-hatched plane is (111) plane. on the trailing and leading dislocations. Between them, a SISF band is formed. On the intersection edge of primary and secondary slip planes, stair-rod dislocations with Burgers vector (AB+C6) are necessary, Fig. 2.12 (a). (Here we did not use the primed letters as used in Fig. 2.5. This will not affect the analysis of SSF-tube type blocking.) As the leading dislocation with Burgers vector 6C moves, a dipole arises. However, a similar dipole cannot be formed on the trailing dislocation 26C, Fig. 2.12 (b). As some part of the trailing dislocation comes into contact with the leading and the stair-rod 24 ((1) 28C “5 i!!! BC+BC 25C (20 (4) 8C 25C . — — _\ g \ l \ 1 \ ; . ”lflefl'r-CS n" \ ‘1 A (b) \ ' BC+BC (b) Figure 2.12 — Formation of SSF—tube blocking. (a) Formation of nonaligned jogs on_a superdislocation; (b,c) formation of a dipole; (d) a SSE-tube along [101] direction. (d) ( A _1. Vi Ap N I’ \' - “EL“ BC'I'BC BC+B§ (C) (d) 8C 25C —--5c- —"- -"— -\ \ CB CB 25c \ 511 5A : / ‘B+C5 E n SAVE 8‘51 AB 4» CB Jwa’) BC BC+BB ((1) (Figure 2.12 continued) 26 dislocations, a reconstruction of the whole configuration takes place which results in a dipole forming on the superdislocations. The dipole contains two SISF bands on the parallel primary slip planes and one SISF band on the second plane, Fig. 2.12 (c). The tube formation ends by the formation of APB bands which takes place during emission of 1/2 <101] screw dislocations (vector CB in Fig. 2.12 (d)) in parallel cross-slip planes. In practice, depending on the jog height, the tube may be observed either as two isolated SISF’s or as a SESF (if the jog is not large). (e) Dislocation Interaction Blocking When two mobile dislocations are moving toward each other, these two dislocations can react at their intersection and produce a third dislocation. This resultant dislocation is either glissile on one of primary slip planes that the two unreacted dislocations belong to, or glissible on the third slip plane, or completely sessile, i.e., to form a stair rod dislocation. Either of the last two resultant dislocations can not move in the primary slip planes of two reacting dislocations, and hence, blocks the two glissile dislocations. Kawabata and Izurni [”1 made a good summary on some possible dislocation reactions in TiAl. In addition to the dislocation reactions of Kawabata and Izumi’s, some additional dislocation reactions in terms of < 100 > type dislocations are summarized by the author. The added dislocation reactions are listed in the following. Reactions between <100> and 1/2< 110] dislocations [100100, + 1/2[i10](,,,, ---> 1/2[110](,,,, [0011000, + 1/2[I10](,,,,---> 1 2 T12 my, # 27 [001],,,,, + 1/2['1'10](,,;, ---> 1/2 in up, # [001](100, + 1/2[110](,;1, ---> 1/2 112 (11;) # Reactions between <100> and <011] dislocations [100].,,,, + [1011mm -—-> mung, + APB + 1/2[101]m,, [100],,,,, + [011](,,;,---> 11231111115) + APB + 1/2[011](,,;, # [001],,,,, + [10mm ---> 12mg“, + APB + 1/2[101],,,,, [001],,,,, + [011],,,,, ---> 1/2[011](,,,, + APB + 1/2[01i](,,,, Reactions between <100> and 1/2<112], 1/3<112], 1/6<112] dislocations [100100, + 1/2[112](,;,, ---> mam, [100100, + 1/3[112](,;,,---> 11312121521) [100],,,,, + l/6[I12](,;,, ---> 1 12 5,2, [001],,,,, + 1/2[112](,,,, ---> may, [001],,,,, + 1/3[112],,,,, ---> W5”, [001](100) + l/6[11§](1”) "'> 1/ 114 (51—1.) Reactions between < 100 > dislocations [1001(001) + [0101000) “‘> [1 101(001) # [1001(001) + [0011(1oo)---> 1.1911501) # [1001(001) + [0011(010) '"> [101](010) # Here the underlines indicate resultant stair rod dislocations and the marks "11'" at the end 28 of equations indicate that the reactions are energetically equivalent in the forward and backward directions. 2.1.5. Lamellar Structure and Its Formation The lamellar structure in TiAl is formed by a phase transformation and it can be classified into two types, type-I and type-II, according to 7 lath crystal orientation relationships within lamellae. Both types of lamellar structures have alternating a2(Ti3Al)/7(TiAl) plates at room temperature and the same crystallographic orientation relationships between 012 and 7. The difference between these two types of lamellar structures is that 7 plates in type-II have the same crystal orientations within a grain, but 7 plate crystal orientations in type-I vary from plate to plate. (a) Type-I Lamellar Structure The type-I lamellar structure is typically formed by the growth of 7 plates into or phase or (12 phase in two phase regions: a+7 and 012+7 (see phase diagram shown in Fig. 2.2). The TiAl (7) lamellar phase is formed on the basal plane of the Ti3Al (02) at the expense of the Ti3Al phase and finally a lamellar structure consisting of lamellae of TiAl and Ti3Al phases is formed in the two phase region ‘62]. The orientation relationship between the TiAl and Ti3Al is the following 1°31, {111}m // (0001)“... <110>TiAl // <1120>,,,A,. 29 The < 110] and <101] in TiAl are not equivalent to each other, but the <1120> directions in Ti3Al are all equivalent. Therefore, there exist six possible crystal orientations, as shown in Fig. 2.13 "’4‘. In Fig. 2.13, orientations E and F are crystallographically identical to the orientations C and D, respectively. Therefore, four distinguishable orientations exist in TiAl laths. With different combinations of two of them, there exist four orientation relationships between two 7 plates: (1) when orientation A in one 7 plate is parallel to orientation A in neighboring 7 plate, a translation order- fault interface or no interface is formed between them; (2) when A/IC or A/lE, a 120° rotational order-fault interface is formed where the c-axes of the two neighboring 7 plates are perpendicular to each other; (3) when A//B, two 7 plates have true-twin relationship; (4) when A/lD or A/IF, pseudo-twin relationship is formed. The pseudo-twin differs from the true-twin. In the pseudo-twin, the atomic sites are in twin positions but these sites are incorrectly occupied by anti-site atoms. Since the energy of lamellar interface with true-twin orientation relationship is lower than those with other orientation relations, the true-twin lamellar interface is more preferred. When the orientation relationships other than the true-twin are observed, a thin lamella of Ti3Al is often found to be sandwiched between the corresponding two TiAl lamellae. However, for mechanical twinning, the energetic criterion is usually not suitable, particularly at high stress and low temperature deformation. In this case a maximum resolved shear stress criterion proposed by Jin is more suitable '2‘". (b) Type-II Lamellar Structure The type-II lamellar structure is formed by Ti3Al (01,) plate growing into TiAl (7) 30 Semaiceb o 255, Sémzvarfle 63683 Aae+$ E 35%? song—etc mo comma—Lon - 2 .N Eswmm S8m2v~>86 Sémzvxrte 553$ 952$ Bémzvxzwoc _ 9:53» _ «5:88:25 S8m2v§§€ «— “55> < “55> 31 phase. This lamellar structure has the orientation relationship between the TiAl and Ti,Al as type-I does, that is, {111},.,,l // (0001)“,Al and <110>m // <1120>.,,,A,. However, since all <1120 > directions in Ti3Al are equivalent to one another, there are only one (12 crystal orientation and only one 7 crystal orientation within a grain, and the 7 crystal orientation is the orientation of the original 7 grain. There are two ways to form either type-I or type-II lamellar structure, as shown in the following. For type-I, (i) a ---> or + 7,, ---> L(a/7) ---> I.(a2/7) (ii) a "‘> (12 "'> (12 + 7p! "‘> 02 + 'Yp "‘> 14(02/7) where 7,, stands for 7 plates, L(a/7) stands for lamellar structure of 7 phase and a phase, 7pt is 7 precipitates. For type-II, (i) 7,, + aP ---> 7m +01p ---> L(7/a) ---> L(7/az) (ii) 7111 -_-> 7m + C{pt ---> 7m + apl _‘-> 117/01) --_> 147/02) where superscript "p" dedicates particles, subscript "pl" dedicates plates, subscript "m" matrix and subscript "pt" precipitates. 2.1.6. Phase Transformation Mechanisms Concerning phase transformation mechanisms, there is a well-accepted stacking 32 fault mechanism proposed by Blackburn “’5‘. A TiAl (L10) stacking fault has the equivalent stacking sequence to the Ti3Al structure (D019). For example, in a SISF (Superlattice Intrinsic Stacking Fault), the four atomic layers around L1o stacking fault turns out to be DOl9 crystal structure, LIOSISF ABCABCABIABCABCABCABC | D019 1 However, this mechanism is not suitable for other stacking faults existing in TiAl. In the case of CSF (Complex Stacking Fault), even though the stacking sequence of four atomic layers in CSF is similar to D0,, structure, LloCSF ABCABCABIABCABCABC Ipseudo-DOl9 | it is a pseudo-DO19 structure since it includes an anti-phase component, as shown in Fig. 2.3 (c). For the SESF (Superlattice Extrinsic Stacking Fault) and the APB (Antiphase Boundary), the Blackburn mechanism is also not suitable, (see Fig. 2.3 (a), (d) and (e)). But, since the SISF is the lowest interface energy condition of all possible stacking faults, and therefore very common in TiAl, it can help explaining how the 0:2 precipites in the 7 matrix. Once a DA] phase nucleates at the SISF of TiAl, T13A1 grows along the 33 octahedral plane of TiAl by increasing the separation of the bounding Shockley partial dislocations in the octahedral plane. The 7 precipitation in 012 basal plane can also be explained by this mechanism in the same way. In this case, glide of Shockley partials a/3 < 1010] in alternate basal planes of Ti3Al matrix results in L1o atomic stacking sequence (TiAl), ABABABABCABCABCABABAB D019 I Ll0 I D019 m ABABABACBACBACBABABAB D019 | L10 l D019 by operating opposite Shockley partials a/3 2, at least some lattice points must shuffle as illustrated in Fig. 2.17. For the multiple lattice structure, the shuffling is much more complicated than the single lattice structure analyzed above. In the case of multiple lattice structure, the shuffles both in the hi planes and in the directions perpendicular to hi planes are necessary to move parent lattice points to the twin lattice positions. 2.2.6. Strain in Twinning On the macroscopic scale, the twinning deformation consists of homogeneous simple shear displacement parallel to the plane K, and in the direction 17,. The magnitude of shear can be obtained by the following equation 49 s=2cot£n1n2=2tan2n1K2=2tanzn2K1 (2 . 18) where < 17,17, is the acute angle between the two directions concerned, < 17,K, is the acute angle between 17, and the normal to the plane K,, and < 17,K, is the acute angle between the 17, and the normal to K,. AZ P = (xyz) P' = (x'y'z') “V Figure 2.18 - Lattice point displacement during twinning shear. Let us look how the specimen length changes corresponding to the twinning. We choose the orthorgonal cartesian coordinate system such that the axes x and y in twinning plane and y axis is in twinning direction, as shown in Fig. 2.18 "2’. The points P (x, y, z) and P’ (x, y+sz, z) are lattice point coordinates before and after twinning, here s is the magnitude of shear. Thus, the ratio of the lengths of the position vectors of P’ and 50 P is as follows I 2 2 2 2 .1. _1 LL=( X "LY +225yz2+(sz +1)Z ) 2=(1+Zs=l=sin7ccosA+szsin2x) 2 x +y +z (2.19) where y = L cosh, z = L sinx. Since the maximum strain (extension or contraction in length) is obtained in the plane of shear, i.e., when x = 0 or A = X, the maximum and minimum values of the extension or contraction can be obtained by differentiation of equation (2.19) and the result is expressed as / (—LL—)max=,/1+s*tanx (2.20) The maximal extension (L’/L)°max and the maximal contraction (L’lL)°M in terms of the magnitude of twinning shear (s) are expressed as follows / (_L_.)e =§+,’_§i+1 (2.21) L m 2 4 L’ s 52 (_)C =—_+ __+1 (2.22) L m 2 4 The extension and contraction of sample lengths are, therefore, conveniently represented on the stereographic projection, as shown in Fig. 2.19 "2' which is for twinning in a zinc crystal. Poles I to V1 indicate the six twinning planes {1012}. So if the sample axis falls into the triangle A, contraction in length occurs for all six twinning 51 planes; if the sample axis falls in the triangle D, extension always occurs. For the triangle B, the twinning planes II, III, V, VI result in contraction, but the planes I and IV extension. In the triangle C, however, the planes II and V cause contraction but the others extension. The twinning deformation is relatively small compared to the deformation by slip. The twinning shear is only a fraction of the lattice parameter, but the shear can be unlimited in slip. Figure 2.19 - The extension and contraction of sample length for twinning in zinc. 52 2.2.7. Prediction of Twinning Elements (Twinning Criterion) According to the analysis in the previous sections, the twinning criterion should include the following points "’3': (i) the twinning shear should be small; (ii) the shuffle mechanism should be simple, that is, q or m should be small; (iii) the shuffle magnitudes should be small; (iv) shuffles should be parallel to the twinning direction rather than perpendicular to this direction. The twinning criterion can be expressed in the matrix form as follows 170' U,U, 5 gm + 3 (2. 23) where 82.11.; is some maximum value of shear, and U,- is correspondence matrix for cubic lattice that must satisfy 8’ = UyUg' - 3 (2.24) 32 = UIyUlg ' 3 (2.25) where g is the twinning shear, and U“,- is the inverse of U,,-. So, for a given value of gm, we can determine the correspondence matrix U,,. This correspondence matrix is used to determine the twinning elements. 53 The twinning criterion can also be expressed in the notation of tensor calculus as“ «dirt/papmfl) < q2(g2,,,, + 4) (2.26) where q is the shuffle parameter described in section 2.2.1, and g2w_ is a given maximum value of shear. So, for given values of q and gum, the values of h, and vi which satisfy above inequality can be determined for a lattice defined by cij or c,,-. Thus the twinning mode with the smallest possible g may be specified. 2.2.8 Twinning in Superlattice Since the mechanical twinning in a superlattice was reported by Laves in 1952 “3', many experimental results on mechanical twinning in superlattices have been reported, such as the mechanical twinning in the superlattices L1, "9'3”, B, ”3'3“, D03 ”3"", L10 “33122341263”, D0,, m9”, DQ919132]. More recently, the review papers by Christian and Laughlin [931, Yoo ‘9“, and Yoo, Fu and Lee [25‘ have also been published for various superlattice structures. It is commonly believed that mechanical twinning makes many important contributions to mechanical properties of ordered intermetallic alloys ‘9". In this section, the mechanical twinning modes in superlattice structures will be summarized. The main difference between the twinning in superlattices and that in disordered structures is that the twinning shear of superlattice creates not only a true-twin in some 54 twinning variants but also a pseudo-twin in other twinning variants. In the pseudo-twin, the atomic sites are in twin positions but these sites are incorrectly occupied by anti—site atoms. In the true-twin, however, the twin relations of the atomic sites are not only satisfied crystallographically but also true in chemistry. True-twins may be further classified into following three types [93:95], (i) type—I/II twin, in which the direct variant gives a twin with a type-I orientation relation (equivalent to orientation relations (2. 1) and (2.4)), and the conjugate variant gives a type-II orientation relation (equivalent to orientation relations (2.2) and (2.3)), (ii) type-II/I twin, which is the reversed of type-I/II, and (iii) combined twin, in which all four classical orientation relations are equivalent. If K, plane is a mirror plane in the parent structure, the type-I/II is defined as type-I since the conjugate modes do not actually represent twins. Similarly, if 17, is a two-fold axis in the parent structure, the type-II/I becomes type-II. The twinning modes discussed in the previous sections are referred to the disordered structure. When the normal twinning mode of the disordered structure is applied to the superlattice, it frequently leads to the incorrectly ordered product, and the true-twinning mode requires a larger shear. This applies to all variants of the normal twinning mode in almost all cubic superlattices, but only to some of the variants in a non—cubic superlattice '93]. In general, for the twinning of a single lattice structure in which the primitive unit 55 cell contains only one atom, when the shuffle parameters q and q are both less than or equal to 2, the superlattice twinning leads to a true-twin; otherwise, if both q and q are larger than 2, the product of twinning is a pseudo-twin ‘93]. For the twinning of a multiple superlattice structure in which the primitive unit cell contains more than one atom, however, a structure shuffle is required. In addition, for the twinning of both single and multiple superlattice structures, an interchange shuffle (or an order shuffle) is needed to form true-twin. If we define a unit cell in the parent in a such way that the primitive lattice vectors are parallel to 17,, 17, and the positive normal to the plane of shear S, the frame of this unit cell is sheared into the cell whose frame is twin related with the parent cell. However, the further relative displacement of the atoms within the unit cell is required to complete the twinning operation and restore the original structure. This atomic displacement in the unit cell is called structure shuffle. The structure shuffle can only complete the twinning operation in crystallography but not in chemistry. 80 various atomic interchange shuffles in the unit cell are also necessary to obtain a true-twin. The structure shuffle and interchange shuffle are schematically shown in Fig. 2.20 which shows the formation of (120) twin in D03 '93]. Fig. 2.20 (a) is the parent unit cell. Fig. 2.20 (b) is the unit cell after twinning shear, which shows that the crystal structure of the parent has been changed by the simple shear. In order to finish the twinning operation, the structure shuffle is needed to restore the parent crystal structure, which is shown in Fig. 2.20 (c). Fig. 2.20 ((1) shows a possible interchange shuffle configuration to produce a true-twin as shown in Fig. 2.20 (e). 56 ./ if M? / 12M” / Figure 2. 20- The sctru ture and inte eerchang suh ffles du uring (120) twinning in D0,. 57 (d) (6) (Figure 2.20 continued) Table 2.3 [93' lists some twinning modes in cubic superlattice structures which is directly derived from the modes in disordered cubic lattice structures listed in table 2.2 and from some other references 17°96]. The shuffles of one half of the atoms are allowed in table 2.3. The twinning modes in table 2.3 are listed in the order of increment of twinning shear. 58 Table 2.3. Twinning modes in cubic lattices Mode No.(m.n) g2 m, Twin type 4.2 1/ 8 8 combined 4.2” 1/8 2 combined 4.5 3/ 8 2 II 4.5T 3/ 8 8 I 2.2 1/2 4 combined 2.2T 1/2 1 combined 2.3 1/2 4 combined 2.3T 1/2 2 combined 4.10 7/ 8 8 I 4.10T 7/ 8 2 II 1.2 1 2 combined 2.5 l 2 combined 2.8 3/2 1 I 2.8T 3/2 4 II 59 Table 2.3. Twinning modes in cubic lattices. (Continued) Mode No.(m.n) g2 m m, mF Twin type 1.3 2 1 2 1 combined 1.3T 2 1 1 2 combined 2.18 7/2 2 4 1 HI 2.18T 7/2 2 1 4 II-I 1.9 8 1 l 1 combined 1.9T 8 1 1 1 combined * m.nT is the "transposed " mode obtained by interchanging K, and 17,, K, and 17,, and changing the signs of the new K, and 17, referred to preceding twinning mode. For the cubic superlattices B2, D03 and L1,, a true-twinning mode without shuffles is predicted if a 1 appears in the shuffle columns of table 2.3 for both the disordered and the superlattice structures. For B32, however, additional examination is required in addition to the shuffle parameter equal to 1 since the primitive unit cell contains two B atoms. The possible true-twinning modes in cubic superlattices are listed in table 2.4. For non cubic superlattices, the primary and complementary twinning modes of L10 (the tetragonal), D0,2 (the tetragonal) and D0,, (the hexagonal) are listed in table 2.5 '9". In table 2.5, A is the ratio of c/a, which is based on disordered cell dimensions, 60 and e/2 is defined in equation b, = (e/2) 17 in which bz is the Burgers vector of a zonal twinning dislocation '9'". structure is 1/ 3. The numerical value of (e/2) for the twinning mode in cubic Table 2.4. Possible true-twinning modes in cubic superlattices Mode No. S K, K, 17, 17, g2 Structures (m.n) (a) Modes without Shuffles 1.3 110 111 001 112 110 2 L12 1.3T 110 112 110 111 001 2 B2 1.9 110 112 001 111 110 8 B2,B32,DO3,L1, 1.9T 110 111 110 112 001 8 B2,B32,DO,,L1, (b) Modes with 50% (non interchange) shuffles 2.3T 110 114 110 221 001 1/2 B2 1.2 001 120 100 210 010 1 B2,Ll, 2.5 0 0 1 1 3 0 1 1 0 3 1 0 1 1 0 1 B2,B32,no, 1.3 110 111 001 112 110 2 B2,B32,Do, 1.3T 110 112 110 111 001 2 L1, 61 Table 2.5. Twinning modes in non-cubic superlattices e/2 Structure q S K, K2 in '12 8 Mn (#1) ('12) type 110 P 2 1/2(110) {111} {111} 1/2(112) 1/2(112) min/(V2) (2A2-1)/(2>\2+1) C 1 1/2(110) {11 1} (001) 1/2(112) 1/2(110) 1/2/1. 1/(2>.2 +1) 1 130,, P 2 l/2<1_10) {111} {111} 1/2(1_12> 1/2(112) (2A2-l)/(X\/2) (2)3-1)/(2>.2+1) C 2 1/2(110) {111} (001) 1/2(112) (110) V2». 2/(2x2+1) 1 130,, (a) 4 i1/3(1120) {1102} {1102} 10101) i game/om) (axe/(ans) (b) 2 (1100) {1121} (0001) 1/3(1126) 1/3(1120) l/x 1/(4x2+1) 1 2.2.9. Mechanical Twin Nucleation and Propagation Mechanical twinning mechanisms can be classified into three categories, that is, (1) dislocation pole mechanism ”“0“, (2) twinning dislocation homogeneous glide mechanism [101,104], and (3) twin-matrix tilt boundary migration mechanism “05'. The main differences between them are the details in the mechanical twin nucleation and propagation. In the third model in which the twin interface is thought to be a tilt boundary, the interface dislocations glide in the planes almost normal to the twin interface. It was supposed that once such a tilt boundary was established, it could migrate easily under a very small shear stress. However, the most intensively studied model is the dislocation pole mechanism. This mechanism describes the twinning as a result of a twinning dislocation rotation around an existing pole dislocation whose screw component is perpendicular to the twinning plane and the component Burgers vector is 62 equal to the spacing of the planes parallel to the twinning plane 198:9”. Much experimental work has been done to try to validate this model. However, some experimental results had indicated that no pole dislocations existed during mechanical twinning, but mechanical twinning was found to occur by a homogeneous shearing of the twinning portion of the crystal with respect to the matrix (100,101). The dislocation pole mechanisms were first proposed by Cottrell and Bilby [98' in 1951 for bcc crystals, Thompson '99] in 1952 for hcp crystals and Venables “0‘" in 1961 for fee crystals. Since the first observation of mechanical twinning in fcc metal (in copper) was in 1957 "0", the twinning theory in fcc was developed much latter than those in bee and hop. Since the dislocation pole mechanism has been well developed, in what follows, the dislocation pole mechanism in three common crystal structures, bcc, fee and hop, will be analyzed in order to illustrate the mechanical twin nucleation and propagation phenomena. (a) Dislocation Pole Mechanism in bcc Metals The dislocation pole mechanism in bcc metals is based on the dissociation of normal dislocation b = 1/2[111] in a (112) plane into two partial dislocations: a sessile mum and a glissile 1/6[111], i.e., 1/2 [111] ---> 1/3 [112] + 1/6 [111] (2.27) 63 Since the partial dislocation 1/6[111] is also glissile in the (121) plane, it can cross-glide onto the (121) plane. The rotation of this glissile partial dislocation 1/6[111] about the normal dislocation 1/2[111] creates a monolayer twin. Since the screw part of the normal dislocation in [121] direction has a Burgers vector equal to the interplanar spacing of (121) plane in length, each complete rotation of 1/6[111] partial dislocation about 1/2[111] normal dislocation in the (121) plane will bring the partial dislocation onto the next adjacent (121) plane. Continuous rotation of this partial dislocation about the normal dislocation thickens the twin layer. So the partial dislocation 1/6[111] is called twinning dislocation and the normal dislocation 1/2[111] is called twinning pole. The twinning plane in this case is then the (121) plane. The propagation of the twin layer in a direction parallel to the twinning plane occurs by the glide of twinning dislocations on every twinning plane. The twin nucleation and propagation model in bcc is schematically illustrated in Fig. 2.21. One complete rotation of twinning dislocation 0F in Fig. 2.21 (a) about the twinning pole 0A creates a monolayer twin, i.e., a twinning nucleus. The glide of twinning dislocations forming incoherent twin boundary, as shown in Fig. 2.21 (b), results in the propagation of twin layer along the twinning plane. (b) Dislocation Pole Mechanism in hcp Metals The twinning modes in hcp structures depend on the c/a ratio (A). Table 2.6 summarizes some twinning modes in hcp metals “m‘m'. Since the (1012) twinning is a common twinning mode in hop, as shown in table 2.6, the dislocation pole mechanism Figure 2.21 - Dislocation pole mechanism for twinning in bcc lattice. in hep is illustrated based on this twinning mode. In this mechanism, the homogeneous shear occurs on every other corrugated plane while the atoms between these planes are involved in a shuffle, as shown in Fig. 2.22. The Burgers vector of the zonal twinning dislocation in this case would be (>3-3)/(>3+3) [1011] ""1. Fig. 2.22 shows two 65 Burgers Vector of Twinning Dislocation > [1011] 0 Figure 2.22 - (1012) twinning in zirconium. Projection of the lattice on the (1210) plane. Circles are in the plane of paper. Squares are a/2 above and below the paper. Solid symbols indicate atom positions in the twin. possibilities for the formation of zonal twinning dislocation: (i) a [0001] dislocation dissociates into the zonal twinning dislocation and a [1010]' pole dislocation (the asterisk indicates that the indexes refer to the twinned lattice); (ii) a [1010] dislocation dissociates into a zonal twinning dislocation and a [00011' pole dislocation. Thus, as the zonal twinning dislocation finishes a revolution about the pole dislocation, it will be on the next adjacent twinning plane. 66 Table 2.6. Twinning modes in hcp metals q S K1 K2 7’1 712 g cm 4 1210 1012 105 1011 1011 (A2-3)/(A\/3)’ all 8 1210 1011 105 1012 3052 (412-9)/(4>.V3) zl.633 6 1100 1122 11% 115 2213 (2)3-4)/3>. (1.633 2 1100 1121 0002 H26 1120 l/)\ (1.633 *A = c/a (c) Dislocation Pole Mechanism in fee Metals The dissociation of normal dislocation 1/2[110] in fee lattice is as follows: 1/2 [110] ---> 1/3 [111] + 1/6 [112] (2.28) But rotation of the 1/6[112] twinning dislocation about the nodes would be confined to the (111) plane, and after the formation of a monolayer twin by one revolution, the twinning dislocation would reunite with the sessile 1/3[111] dislocation to reform 1/2[110] dislocation. This reformed 1/2[110] dislocation would glide onto the next (111) plane and dissociate into 1/3[111] + 1/6[112] again, so that the revolution of twinning dislocation 1/6[1 12] is in a new (111) plane. Repetition of this dissociation-revolution- reunion procedure thickens the twin layer. This mechanism is schematically shown in Fig. 2.23 “°°'. At beginning, a normal dislocation AC lies in plane (a) between the nodes 67 Figure 2.23 - Dislocation pole mechanism in fee lattice. 68 N, and N,, and the poles X and Y lie on plane (b), as shown in Fig. 2.23 (a). The normal dislocation AC dissociates into A0: + aC on (a) plane as shown in Fig. 2.23 (b), and when the twinning dislocation line reaches a semicircle it is unstable and quickly extends to the configuration as shown in Fig. 2.23 (c). The partials marked "up" and "down" in Fig. 2.23 (c) reunite with A01 to form AC, and a unit jog traveling along the AC from the node N, (or N,) move the AC into the next (a) plane and repeat the revolution as shown in Fig. 2.23 (d) and (e). In this way, a lenticular twin can be built up, as shown in Fig. 2.23 (f). The twin thickness is linrited by the backward stress and by the length of the pole dislocation. 2.2.10. Twin Shape There are two models to predict the twin shape: one model considers system energy change associated with twin nucleation and propagation, and the other model considers forces acting on individual twinning dislocations. These two models have been frequently used in modelling the mechanical twinning behavior and predicting the twin Shape [112-114]. (a) A Model Concerning The Energy Change Associated With Twinning The energy associated with twinning was expressed by Cooper in 1965 “‘5‘. When we predict the equilibrium shape of a twin, we usually consider the minimum total energy change in the specimen during the twinning. This can be done by differentiation of an expression for the total energy of the twinned specimen. 69 The energy terms considered in this model are: (i) twin boundary energy, which corresponds to the twin boundary surface tension; (ii) twinning dislocation interaction energy, which corresponds to the mutual repulsive forces between the twinning dislocations making up the incoherent twin boundaries; (iii) the elastic strain energy due to the external stresses (the externally applied stress and the local stress concentration); (iv) dislocation line energy due to the dislocation line tension. The expressions of these four energy terms reported by Cooper “‘5' are briefly summarized in the following section. (1) Twin Boundary Energy The twin boundary energy is equal to the twin boundary surface energy multiplied by the total twin boundary area: B, = A7h(p-l) (2.29) where A is a twin shape constant, A = 1 when one side boundary of twin is considered, A = 2 when considering both side boundaries; h is the spacing between neighboring twinning dislocations in the incoherent twin boundary; p is the total number of twinning dislocations in one side of a twin boundary. When the twin boundary consists of a large number of twinning dislocations, the twin boundary energy term can be expressed as 70 E, = A7hp. (2.30) (2) Twinning Dislocation Interaction Energy Since the length and the width of a twin are much lager than the spacing between the twinning dislocations which form the incoherent twin boundary, the twinning dislocations can be thought to be parallelly distributed. In addition, the following two assumptions are made for the expression of dislocation interaction energy term: (i) any pair of twinning dislocations will have zero interaction energy when separated by a suitably large distance L; (ii) the twin is reasonably thin compared with its length so that the dislocations behave as if all are in the same glide plane but there is no interaction between one boundary and the other. The twinning dislocation interaction energy (E,) for one side of twin boundary is as follows, .. sz , £_ 2 ———2n(1_v) [p lnh K(p)] (2.31) where K(p) =ln(p-2) !+ 1,2— [ln(p-l) 1+ln(i—1) 1] (2.32) 1:2 G is the elastic shear modulus, b is the Burgers vector of a twinning dislocation, and v is Poisson’s ratio. (3) Strain Energy Due to The External Stress The work done by the externally applied shear stress on the formation of a twin, 71 i.e., the strain energy, is expressed as E3=thoap2f(a)h (2.33) where g is the twinning shear, To is the applied shear stress, a is the interplanar spacing of the twin composition plane, and f(a)=a‘2(e'“-1)+%. (2.34) If a = 0, 1(0) = l, and if a is large, f(oz) ---> Has. (4) Twinning Dislocation Line Energy The total dislocation line energy on one side of a twin boundary is E, = 0.5pr2. (2.35) Therefore, the total energy on one side of a twin boundary associated with twinning is the summation of all these four energy terms, that is, E = E, + E, + B, +13, (2.36) 01' 72 2 E=Yh(p-1) + Gb [pzlné-Iflp) 1 +291,ap2f(a) ling-13Gb? 211(1-v) h (2.37) It must be noted that in this model the energy due to the lattice friction on twinning dislocations is not considered. (b) A Model Based on Twinning Dislocation Interaction Forces Cooper “‘6' first proposed a model based on twinning dislocation interaction forces in 1966. Marcinkowski and Sree Harsha “”1 that used this model to predict the twin shape. This model has also been used by others to predict twin shape “mm. The forces considered in this model include (i) the force due to the surface tension of twin interface, (ii) the force due to the twinning dislocation line tension, (iii) the force due to the applied stress (iv) the force due to the repulsion between twinning dislocations. This model can give more detailed results than the model considering the system energy change. The expression of each force in this model is briefly summarized in the following. (1) The Force Due to The Surface Tension The force due to the surface tension, F”, is numerically equal to the specific surface energy, 73 Far. = "Y (2- 38) where 7 is the twin boundary energy. The negative sign in the equation indicates that the positive direction of this force points to the outside of the surface considered. The unit is dyn/cm or N/m. The twin plane energy is approximately equal to one half of the intrinsic stacking fault energy. This is because the number of mis-stacking with respect to the second neighbor planes across the intrinsic stacking fault has twice as many as that across the twin plane. (2) The Force Due to The Twinning Dislocation Line Tension The line tension of a dislocation half loop causes a force (per unit length), Fm, on the edge part of the loop, FL.T. = ‘ Zsz/l (3.39) where G = the shear modulus, b = the Burgers vector, 1 = the radius of curvature of the dislocation loop at the twin tip. (3) The Force Due to The Applied Stress This force (per unit length), F I, is equal to the resolved shear stress times the 74 Burgers vector, F, = 1b. (2.40) The applied stress can be either the externally applied stress, or the locally concentrated stress, or the residual stress in the matrix, or all of them. (4) The Force Due to The Twinning Dislocation Interaction Since the twinning dislocations in a twin boundary are all the same type of dislocations, i.e., they have the same Burgers vector, the interaction force between them is repulsive. The dislocation interaction force, F,, between two twinning dislocations, b, and b,, can be expressed as follows G = . G . . F, 2n(1_v)R[b,xC) (b,xt>1+ [(blC)*(b,C)] (2.41) 211R where 3' is the dislocation line direction, R is the distance between two dislocations, G is the shear modulus, v is Poison’s ratio. Since many twinning dislocations are involved in a twin boundary, the total interaction force on a twinning dislocation (on the mth twinning dislocation in this case) will be the sum of all interaction forces, 75 n F,,.=-—G— 11%;(b,,xt)a(b,x()+(b,,,<>*wo>uo>wn>w ((1) (Figure 3.6 continued) 99 indicate the atoms located in the plane of the paper and the shaded squares indicate the atoms in the plane just beneath the paper. The atomic arrangements are viewed in [110] direction. The right side column shows the atomic stacking sequences along [111] direction before and after twinning, in which the primed letters indicate the sheared planes. Therefore, the glide of the first twinning dislocation (the leading twinning dislocation in Fig. 3.1) creates a twin of two atomic layers, as shown in Fig. 3.6 (a). Investigating the atomic stacking sequence change after the twinning, one can see that this two atomic layer twin actually is an intrinsic stacking fault. The glide of the two twinning dislocations in two adjacent (111) planes creates a twin with three atomic layers, see Fig. 3.6 (b), which is an extrinsic stacking fault as shown in the right side column. After the glide of the third twinning dislocation on the next adjacent (111) plane with respect to the first two twinning dislocations, a four atomic layer twin is formed, and it has no single stacking fault feature, but it does have two twin planes, see Fig. 3.6 (c). Therefore, following the second twinning dislocation, the glide of any 1/6[112] twinning dislocation on the (111) twinning plane does not change the features of twin interfaces but it thickens the twin with one more atomic plane, see Fig. 3.6 (c) and (d). The thickness of the thin twin layers can be calculated based on the twin nucleation and propagation mechanism proposed above. Although the crystal structure of TiAl is anisotropic, the atomic stacking along any <111 > directions is similar. This means that the spacings of all {111} planes are identical. Therefore, the thin twin layer thickness can be calculated by counting the number of emitted twinning dislocations, assuming that the observed twin layers are perfect twins. For the longer twin A in Fig. 100 3.1 (a), the thickness is 7.2 nm, and the thickness of the shorter one B in Fig. 3.1 (a) is 5 .8 nm. For the rest of the twins in Fig. 3.1 (a), since most twinning dislocations have already propagated farther away from the region shown in Fig. 3.1 (a), the thicknesses can not be calculated. (b) Twin Nucleation at Twin Interfaces The long twin layers in Fig. 3.1 were continuously illuminated by the electron beam, and it was found that these twin layers were continuously growing into the grain interior. These growing twin layers then met an existing thin twin lath (called "vertical twin" in this case), whose twin plane was (111) plane, and crossed the vertical twin by the initiation of new twinning dislocations on the other side of the vertical twin. This is the case of twin nucleation at twin interfacies. Fig. 3.7 shows the entire sequences of twin nucleation and propagation procedure. In Fig. 3.7, the left sides are the TEM micrographs, which were taken sequentially during the twin intersection, twinning dislocation emission and twin propagation; the right side column is a schematic illustration of twin nucleation and propagation sequence corresponding to the left side pictures. Fig. 3.8 (a) is the thin twin image taken by tilting the specimen such that the electron beam was parallel to the twin layer interfaces. The diffraction pattern taken across these thin layers is shown in Fig. 3.8 (b). The four thin twin layers in Fig. 3.7 are denoted by C, D, E and F. The configuration shown in Fig. 3.7 (a) is the starting state of these sequential images, where 101 Figure 3.7 — The sequences of mechanical twin nucleation and propagation in TiAl. 102 (Figure 3.7 continued) 103 (Figure 3.7 continued) Figure 3.8 - (a) The side view of twin layers C, D, E, and F in Figure 3.7, and (b) the diffraction pattern taken across these twin layers. the thin twin layers C, D, E and F in (a) originated from the grain boundary, which was located to the right and below the picture. At the beginning, twin C was in the intermediate state of propagation and did not reach the existing vertical twin, as shown in Fig. 3.7 (a). While twin layers E and F had intersected with the vertical twin lath, and the high density twinning dislocation pile-ups had formed within these two twin layers at the intersection, which indicates that the existing twin lath was acting as a barrier for the propagation of twins E and F. However, for twin D, in addition to the high density of twinning dislocation pile-up in the twin layer at the intersection, this twin had crossed the vertical twin lath and extended to a certain length in the other side of the vertical twin. Investigating the intersection portion within the vertical twin lath, there 105 is no evidence of structure and crystal orientation changes within it. As time passed, the twin C propagated and met the vertical twin, and the twin D’ further elongated, as shown in Fig. 3.7 (b). The twinning dislocations within twin C were not densely piled up at the intersection, indicating that a stress concentration at the intersection was not formed at this moment. It was found that the propagation of twin D’ occurred as the twinning dislocations were emitted from the interface of vertical twin. So more dislocations can be seen in twin D’ in Fig. 3.7 (b). With the longer time, as shown in Fig. 3.7 (c), the twin D’ further propagated so that the twin tip was out of this picture; more and more twinning dislocations within the twin C came and piled up against the vertical twin interface resulting in the stress concentration at the intersection as indicated by an arrow in Fig. 3 .7 (c). In Fig. 3.7 (d), the stress concentration resulting from the twinning dislocation pile-up at the intersection within the twin C was so large that the two twinning dislocations were emitted on the other side of the vertical twin lath. The sequence of these two twinning dislocation emission is as follows: Two twinning dislocations were first gradually bowing out from the interface of vertical twin, and then, after the radius of dislocation lines reached a certain value, they jumped into the matrix very quickly and propagated to a certain distance. The layer formed by the glide of these two twinning dislocations is a three atomic plane twin, as shown in Fig. 3.2 (c) and (d) and Fig. 3.6 (c), and is called the twin nucleus as mentioned in the previous section. 106 As more and more twinning dislocations were emitted from the vertical twin interface, this twin nucleus grew into the matrix and became a thicker twin layer. The propagation of this twin layer is clearly seen from Fig. 3.7 (e) and (1). Taking an existing dislocation line in the matrix as a reference, the relative locations of twin layer C’ with respect to the reference dislocation line in Fig. 3.7 (e) and Fig. 3.7 (f) indicate that twin C’ in Fig. 3.7 (e) has propagated to a longer twin layer C’ as shown in Fig. 3.7 (1). However, it was found that the twin layer could not propagate continuously until more twinning dislocations were emitted from the interface. It appeared that the trailing twinning dislocations were pushing the front twinning dislocations to move forward and this push resulted in the continuous propagation of the twin layers. Once a twinning dislocation was pushed out from the source, the twin interface at the intersection in this case, it caught up to the front dislocations very quickly compared to its bowing out period. This indicates that the twin propagation might be controlled by the twinning dislocation emission procedure, not by the twinning dislocation glide in the twinning plane. Since twin D’ had extensively propagated so that the twin tip was far away from the location where the images were taken, very few twinning dislocations can be seen in twin D’ in Fig. 3.7 (e) and (1). During the investigation, the twin E and the twin F were not found to cross the vertical twin lath even though the dislocation pile-ups were seen within them at the intersections. This was probably because no more twinning dislocations were emitted 107 within these two twin layers from the original twinning dislocation source, i.e. , the grain boundary. Therefore, the stress concentration caused by the dislocation pile-up was not large enough to transfer the twinning strain across the vertical twin lath. (c) Twin Layer Morphology Near The Twin Tip According to the twin nucleation and propagation procedure analyzed above, the morphology of thin twin layers shown in Fig. 3.1 and in Fig. 3.7 can be easily determined. Fig. 3.9 shows a schematic three dimensional diagram of a thin twin layer. The curved surface near the twin tip (with darker shading) is an incoherent twin boundary, which is composed of twinning dislocations lying in the boundary. On an atomic scale, this twin boundary is not a smooth interface but has a stair-like shape with the stair height equal to the interplanar spacing of twin planes. The top and bottom flat interfaces (with diagonal line shading) are coherent with the matrix, which are normal twin planes. The lower diagram in Fig. 3.9 is a side view (cross section) of the thin twin layer. Therefore, the shape of thin twin layer near the twin tip is semi-lenticular. 3.2.2. Post-Mortem Observation of Mechanical Twinning During Creep Deformation A post-mortem investigation of a creep specimen deformed up to the end of primary creep state (the first specimen as indicated in the experimental section) indicates that fine mechanical twins nucleated at grain boundaries and propagated into the grain interior during the early stage of creep deformation in 7 grains in a similar way as described in the previous section. Fig. 3.10 is an example which shows that a large 108 f—— Twinning direction Twinnig dislocation lines Incoherent twin interface % Coherent twin interface Cross section of twin layer Coherent twin interfaces Figure 3.9 - The thin twin morphology near the twin tip. 109 number of fine mechanical twins were formed by the mechanism described above. In this case, the fine mechanical twins formed to accommodate the local stress concentration at a grain triple point [3°]. Looking at the area between the fine mechanical twins and the untwinned matrix, one can easily see that some fine mechanical twins did not propagate all the way across the grain, as indicated by an arrow in Fig. 3.10. Diffraction patterns taken within the fine twin area and within the untwinned matrix region, as shown in Fig. 3.10 (b) and ((1) respectively, show that the matrix within this fine mechanical twin region has the same orientation as the equiaxed 7 grain within which the fine mechanical twins formed. This indicates that fine mechanical twins in Fig. 3.10 (a) are formed by deforming the 7 grain. If we assume that the volume percent of twins in the fine twin region in Fig. 3.10 (a) is about 50%, the shear strain in this region is a half of the twin shear (0.707), i.e., the shear strain is equal to 0.354. The distance between two points "P" and "R" in Fig. 3.10 (a) is about 10 pm. So, the relative shear displacement of point "P" to point "R" in the direction parallel to the twin-matrix interfaces is about 3.54 um, if we assume that all fine twins sheared in the same direction. In a similar way, by counting the number of laths between points "P" and "R" and assuming that the thicknesses of all laths are roughly equal and one half of these laths are fine twins, we calculated the thickness of a fine twin lath to be typically about 50 nm '3“. In the diffraction pattern shown in Fig. 3.10 (b), an extra set of diffraction spots (dots in the figure) in addition to the spots from fine twins and the matrix indicates that 110 (a) Figure 3.10 - The configuration of fine mechanical twins resulting from the accommodation of local stress concentration at a grain triple point (a). Diffraction pattern across the interfaces (b) shows that these fine laths are twin—related with matrix and the existence of fine D0,, phase between them (0). Diffraction pattern taken in untwinned area within the same grain without tilting the sample ((1) shows that the untwinned region has the same crystal orientation as the matrix of fine mechanical twins as shown in (b). (e) is the indexes of (d). It also shows that the growth of fine mechanical twins as pointed out with an anew between the completely developed fine twins and the matrix stopped in the grain interior. 111 (Figure 3.10 continued) 112 (9) (Figure 3.10 continued) 113 there exists a third phase in this fine twin region shown in Fig. 3.10 (a). An analysis of these diffraction spots showed that these diffraction spots resulted from a D0,, type structure. 3.3. Discussion 3.3.1. Mechanical Twin Nucleation in TiAl Mechanical twinning was observed to nucleate by bowing out of one (or two) l/6[112] twinning dislocation(s) from the grain boundaries and the twin interfaces in this study. The twin nucleus can be either an intrinsic stacking fault or an extrinsic stacking fault. However, the size of twin nucleus, i.e. , the distance that leading twinning dislocation propagated into the grain interior before the second twinning dislocation was emitted, was observed to vary for each twin. Since the size of electron beam used in TEM was very small, the volume illuminated by electron beam was small too. The thermal activation in the region near the leading twinning dislocation resulting from the electron beam illumination varied from place to place. In addition, the local stress state was probably not the same along the grain boundary, and it would change with the emission of twinning dislocations from the grain boundary. Therefore, both the thermal and the stress conditions at each location along the grain boundary were different. However, a certain amount of local stress with the principal stress axis in the vicinity of [551] direction in the case of tensile stress state is necessary for the (lll)[112] twinning to nucleate. 114 As the twin nucleus grows, i.e., as the leading twinning dislocation propagates into the grain interior, the region of stacking fault behind the leading dislocation increases, and therefore a backward force on the leading dislocation develops in the direction opposite to its propagating direction (The detailed analysis on the backward force will be presented in the next chapter). When the twin nucleus reaches a certain size, the backward force will balance the driving force on the leading dislocation and the motion will cease. With the emission of the second dislocation, the twin propagates farther, but the second dislocation can not reach or overtake the leading dislocation since they repel each other due to the same type of l/6[112] dislocations "‘2'. Therefore, the second dislocation appears to push the leading dislocation forward. However, since the stacking fault behind the second dislocation is the extrinsic stacking fault (see Fig. 3.3), which possesses a similar stacking fault energy to the intrinsic stacking fault, the twinning dislocations only move slightly farther into the grain. After glide of the third twinning dislocation, the interface between the twin and the matrix is a twin plane, whose interface energy is about a half of the intrinsic stacking fault energy. The backward force on the twinning dislocations following the second one is approximately one half of that on the first two twinning dislocations. Therefore, the twin propagation is easier than the nucleation. Since the effect of a local stress concentration is limited, the local stress state could not directly provide the twinning driving force to cross the entire grain. A possible scenario for a transgranular twinning is the following: The twinning dislocations are emitted from the grain boundaries and the twin interfaces due to the local stress 115 concentrations and forced to move forward to the limited range within which the local stress is effective. Subsequent twinning dislocations push the front twinning dislocations beyond the local stress field. In this case, the twin propagation occurs under the continuous emission of twinning dislocations from the grain boundaries and the twin interfaces. Since the nucleus of mechanical twin is formed by bowing out a twinning dislocation from the dislocation source, the driving force arising from the local stress concentration must be large enough to overcome another backward force resulting from the dislocation line tension. As the twinning dislocation bows out, the radius of the twinning dislocation line decreases. This results in an increase in the backward force due to the dislocation line tension. After the radius of twinning dislocation line reaches a minimum value, the radius will increase as the twin nucleus grows, that is, the backward force due to the dislocation line tension decreases with twin nucleus growth. Therefore, there should exist a maximum backward force on leading twinning dislocation corresponding to the minimum radius of twinning dislocation line during twin nucleation. In order to form a twin nucleus, the driving force resulting from local stress state must be large enough to compensate this maximum backward force besides the backward force due to the increment of stacking fault area. So, to emit twinning dislocations continuously at the grain boundaries and the twin interfaces, the local stress must maintain being larger than a certain value, the critical stress for twinning dislocation emission from the grain boundaries and the twin interfaces. If the magnitude of local stress is lower than this critical stress, according to the twin propagation mechanism 116 proposed in the previous section, the twin could not propagate since no more twinning dislocations would be emitted. 3.3.2. Mechanical Twinning Mechanisms in TiAl Mechanical twinning mechanisms can be classified into three categories, that is, (1) dislocation pole mechanism [98"°°', (2) twinning dislocation homogeneous glide mechanism “0””, and (3) twin-matrix tilt boundary migration mechanism “03’. The main differences between them are the details in the mechanical twin nucleation and propagation. In the third model in which a twin interface is thought to be a tilt boundary, the interface dislocations glide in the planes almost normal to the twin interface. This theory assumes that once such a tilt boundary was established, it could migrate easily under a very small shear stress. However, the most intensively studied model is the dislocation pole mechanism. This mechanism describes the twinning as a result of a twinning dislocation rotating around an existing pole dislocation whose screw component is perpendicular to the twinning plane and the component Burgers vector is equal to the spacing of the planes parallel to the twinning plane ‘98'100'. Much experimental work has been done to try to validate this model. However, some experimental results indicated that mechanical twinning was found to occur by a homogeneous shear of the twinning portion of crystal with respect to the matrix “°"m‘. This homogeneous shear mechanism requires that the individual twinning dislocations glide successively on every neighboring plane parallel to the twinning plane. Comparing our observations with the twinning mechanisms described above, the 117 twinning in TiAl is consistent with the twinning dislocation homogeneous glide mechanism. Since the twin nuclei were formed by bowing out twinning dislocations at the grain boundaries, it was not necessary to form a dissociated dislocation jog in the twin plane. The normal dislocations and the superdislocations in TiAl are usually in the state of dissociated forms due to the relatively low stacking fault energy of this material. The more commonly dissociated configurations are such that the stacking faults are bounded by the Shockley partial dislocations, i.e., the Shockley partial dislocations are common and stable in TiAl '6'. Thus the emission of Shockley partial dislocations (twinning dislocations) at the grain boundaries and the twin interfaces instead of perfect dislocations (normal dislocations or superdislocations) is reasonable. It is due to the low stacking fault energy that the twin nuclei can propagate easily by the successive glide of these Shockley partial dislocations on every adjacent plane. However, a recent study by Farenc, Coujou and Couret [32‘ indicates that a dislocation pole mechanism was observed in situ in a TiAl specimen deformed at room temperature in tension. These authors identified the twinning dislocations by characterizing the stacking fault fringe configurations in the tilted twin images. The twin propagation procedure is the same as the one observed in our study, but the twinning dislocation source is within the grain interior and the twinning procwds by the rotation of an a/6<112] type partial dislocation around a perfect dislocation. In our study on a large strain specimen, which was deformed up to a tertiary state of creep deformation, an evidence of twin nucleation within the grain interior was also found, as shown in Fig. 3.11. Three individual thin twin layers, indicated by letters "A", "B" and "C", can be 118 seen in Fig. 3.11. These twin layers are completely within a 7 grain interior. So the nuclei of these twin layers were formed inside the 7 grain. The source of twinning dislocations might be some defects in the grain. Because no inclusions and second phases were found near the twin layers, the possible source of twinning dislocations might be the dislocation jogs. Thus, the twin layers probably propagated by the dislocation pole mechanism. Therefore, the mechanical twinning mechanism in TiAl may depend on the location of twin nucleation and the amount of strain. Since the large strain creates more dislocations within the 7 grains, the possibility of formation of dislocation jogs and pole dislocations within the 7 grains increases with increasing strain. So the dislocation pole mechanism is reasonable at large strain. If the twin nuclei are formed at grain boundaries, the twinning dislocation homogeneous glide mechanism is preferred, but in the case that the twin nuclei are formed in the grain interior, the dislocation pole mechanism is more probable. It must be noted that the image of dissociated perfect dislocations in a tilted microslip band is similar to that of the fine mechanical twins since the perfect dislocations in TiAl easily dissociate into the partial dislocations with the stacking faults between the partials. Therefore, it is necessary to identify whether the observed images are fine mechanical twin layers or microslip bands before further analysis. A diffraction pattern analysis is necessary to make this distinction. 119 (a) Figure 3.11 - (a) shows three thin twin layers indicated by letters "A", "B" and "C" within a 7 grain interior, which indicates that the nuclei of these twin layers were formed inside the 7 grain; (b) is a diffraction pattern taken across the twin layers. 120 0 ° illT -__ 111 111 l u o X 000 D‘ 111 ill 0 1111 0 l3 1:: O 0 (Figure 3.11 continued) 121 3.3.3. On The D0,, Phase in Fine Mechanical Twins The diffraction pattern in Fig. 3.10 (b) indicates that fine mechanical twin region at grain triple point as shown in Fig. 3.10 (a) consists of three different layers: the matrix, the fine twins and the D0,, structural phase. The formation of fine mechanical twins has been analyzed in previous section. But the formation of DOI, layers accompanying the twinning is not clear. No experimental results clearly show how the D0,, layers form. Therefore, in the following, some possibilities of D0,, diffraction spots are discussed. (a) If the stress concentration at the grain triple points and/or grain boundaries is such that it does not emit 1/6<112] twinning dislocations on every (111) plane, four atomic layers of D0,, structure, i.e. , a nucleus of the Ti3Al crystal structure, will be formed by glide of a single 1/6 <112] twinning dislocation on a (111) plane “43'. So, in the case that a large volume of 7 crystal is deformed by the heterogenous glide of 1/6[112] partial dislocations, the diffraction spots resulting from the nuclei of DO,9 crystal structure may be visible. (b) The crystal structure across a true-twin plane in TiAl is also a three atomic layer D0,, structure, see Fig. 3.6 (d). If one uses a large size of aperture in selecting area diffraction, the total volume of D0,, layers in the fine twin region will be large, and therefore, its contribution to D0,, diffraction spots is also considerable. It is worth noting that such a uniformly distributed thin lath configuration in Fig. 3.10 (a) is unlikely to be formed by phase transformations such as or, --- > a,+7 and or - -- > a,+7 due to its large interfacial energy. Such a thin 7 phase layer could not be thermodynamically stable during a phase transformation. Therefore, the thin 7 phase 122 layer should either grow and become thicker or be eliminated by the growth of adjacent 7 laths. The resultant configuration in the case of phase transformation should be the lamellae containing coarse 7 laths with different thickness. The observation of fine twin configurations only near the equiaxed 7 grain triple points in this study also provides an evidence that the fine mechanical twins are formed due to the local stress concentration at the equiaxed 7 grain triple points but not by the phase transformation. In addition, the composition of the 7 phase is the same at the creep test temperature, 760 °C, as at a higher phase transformation temperature, so the chemical driving force to decompose 7 phase is zero. 3.3.4. Twin Morphology Twin morphology has long been described as either a lenticular shape or an elliptic shape, particularly when modelling the mechanical properties, such as elastic strain energy and stress field, of twins 11‘3"“:“5'. There are two methods that have been generally used to predict twin shapes: one is based on the twinning energy calculation, and the other is based on the elastic forces acting on individual twinning dislocations. Both methods assumed that a twin has a symmetric shape so that the lenticular or elliptic shape was predicted. However, the thin twin layers observed in this study do not have a symmetric shape but a semi-lenticular shape near the twin tip, as shown in Fig. 3.9. Since the energy of the flat coherent twin plane is lower than that of the curved incoherent twin boundary and the strain energy resulting from the twinning shear can be neglected for very thin twin layer “‘3', the semi-lenticular twin shape may be energetically more stable than the fully elliptic twin shape. 123 Based on the traditional symmetric twin morphology, Marcinkowski and Sree Harsha “‘7' first calculated the twinning dislocation distributions in incoherent twin boundaries and the stress distribution surrounding the twin tip by considering the elastic forces acting on the twinning dislocations. The result shows that the twinning dislocations close to the twin tip are distributed more densely than those away from the tip. This trend of dislocation distribution is consistent with the observations in this study as shown in Fig. 3.1 and Fig. 3.7. But the twinning dislocations in the upper and the lower incoherent twin boundaries of a symmetric twin layer tend to be vertically aligned one to another “‘7‘. If this were true in this study, the overlapped stacking fault fringe images and the coupled twinning dislocation lines would be seen. But the twin layer images observed in this study show only one set of periodically changed stacking fault fringes and the single dislocation lines. This indicates that the twin layers observed in this study are thickened by the glide of twinning dislocations on only one side of the twin layer and the shape near the twin tip is semi-lenticular. 3.4. Summary The following summary can be made concerning the mechanical twin nucleation and propagation in investment cast near gamma TiAl (Ti-48Al-2Nb—2Cr) specimens creep deformed at 765 °C. 1. Mechanical twin is nucleated by the bowing out of twinning dislocations at the grain boundaries and the twin interfaces due to the local stress concentration. The 124 nucleus for true-twinning observed in this study is either a superlattice intrinsic stacking fault (SISF) or a superlattice extrinsic stacking fault (SESF). The superlattice intrinsic stacking fault is formed by bowing out one l/6[112] twinning dislocation from the grain boundaries or the twin interfaces. The superlattice extrinsic stacking fault is formed by emission of two 1/6[112] twinning dislocations from the grain boundaries or the twin interfaces on adjacent (111) planes. 2. The mechanical twin propagation mechanism in TiAl observed in this study is a homogeneous glide of twinning dislocations on every adjacent twinning plane. 3. Mechanical twin propagation mechanisms in TiAl depend on the locations of twin nucleation. If a twin nucleates at the grain boundary, it propagates by the twinning dislocation homogeneous glide mechanism. However, if a twin nucleates within a 7 grain interior, the twin propagation is controlled by the dislocation pole mechanism. 4. The locations of twin nucleation seem to be affected by strain. In low strain creep specimen (4%) investigated, we found that all fine mechanical twins resulting from the accommodation of stress concentration at grain triple points initiated at grain boundaries. But in a large strain creep specimen, twins originating in the grain interior were observed. 5. The reason of occurrence of D0,, diffraction spots is not clear at present. It seems to be the result of the heterogenous glide of 1/6[112] partial dislocations and the 125 existence of three layers of D0,, structure at each true-twin interface. Further work is needed to test these hypotheses. 6. The thin twin morphology near the twin tip observed in this study has a semi- lentucular shape rather than a lenticular or an elliptic shape. CHAPTER FOUR FORCE AND STRESS ANALYSES ON TWIN PROPAGATION The stress required for mechanical twin propagation is supposed to be such that the driving force for mechanical twin propagation must be equal to or larger than the backward force (F,,) on the twinning dislocations. This backward force acts on the twinning dislocations in an opposite direction to the twinning dislocation moving direction, i.e. , the twin propagation direction. At an equilibrium condition, i.e. , at a constant propagation rate, the backward force is equal to the driving force on each twinning dislocation. Thus, the stress necessary for mechanical twin propagation must be such that the driving force on the twinning dislocations at least equals the backward force on the same twinning dislocations. Because the twinning dislocations are the same type of 1/6[112] Shockley partial dislocations (see section 3.2.1), a repulsive force exists between the twinning dislocations. This repulsive interaction force results in a separation between the twinning dislocations. But, the stress imposed on a twin layer results in a reduction in spacing between twinning dislocations. So, we can obtain the stress distribution along a twin layer by calculating the forces on the twinning dislocations in the twin layer. 126 127 4.1. Theory 4.1.1. Forces on Each Twinning Dislocation Before the forces on each twinning dislocation are computed, the forces acting on each twinning dislocation in a twin layer are identified. The incoherent twin boundary structure shown in Fig. 3.3 is used for the force analysis. On the first (the leading) twinning dislocation, four forces are identified, as shown in Fig. 4.1 (a): (b): (i) the forward dislocation interaction force, F”, which results from the repulsive force of the twinning dislocations behind the leading dislocation; (ii) the applied force, F” (or external force F“ as defined later), which results from the externally applied stress, the local stress state (including residual stress), and is in the same direction as the twinning propagation direction; (iii) the internal friction force, Fm, which always acts in the opposite direction to the twin propagation direction; (iv) the force due to the superlattice intrinsic stacking fault behind the leading twinning dislocation, Fsm, which pulls the leading twinning dislocation backward. On the second twinning dislocation, six forces are identified, as shown in Fig. 4.1 (i) the forward dislocation interaction force, F“, which is due to the dislocations behind the second twinning dislocation within the twin layer; (ii) the applied force, F,; 128 (iii) the force due to the superlattice intrinsic stacking fault, Fs,s,.., which pulls the second twinning dislocation forward; (iv) the backward dislocation interaction force, FM, due to the leading twinning dislocation; (v) the internal friction force, ch. (vi) the force due to the superlattice extrinsic stacking fault, FSESF, which pulls the second twinning dislocation backward. 0n the third twinning dislocation, six forces are identified, as shown in Fig. 4.1 (c). (i) the forward dislocation interaction force F”; (ii) the applied force F,; (iii) the force due to superlattice extrinsic stacking fault, FSBF, which fulls the third dislocation forward; (iv) the backward dislocation interaction force, FM, due to the first two twinning dislocations; (v) the internal friction force, Fae; (vi) the force due to the twinning plane surface tension, Fm, which pulls the third twinning dislocation backward. 0n the fourth twinning dislocation, the forces F,, FM, FM, and FM are similar to those defined for the third dislocation, but there are no stacking fault forces on this dislocation. Forward and backward FM, forces exist on the fourth twinning dislocation, 129 as shown in Fig. 4.1 (d). However, since they are the same type of forces acting in the opposite direction, the net force from them is assumed to be zero, so they can be eliminated during the force calculation. 0n the rest of the twinning dislocations, the forces on each twinning dislocation are exactly the same as those on the fourth one. Twinning direction [113] ( J. (111) plane (a) Figure 4.1 - Forces on each twinning dislocation in the thin twin layer. (a) On the first twinning dislocation, (b) on the second one, (c) on the third, and (d) on the fourth. 130 Twinning direction [11.2.] .I. (111) plane 0)) Twinning direction [112] < .1. Fa ( -—'+ Ffric Ff.d. ( ——) Fb.d. FSESF ‘ ) Ftp. .I. (111) plane (C) (Figure 4.1 continued) 131 Twinning direction [112] ( J— (111) plane ((1) (Figure 4.1 continued). 4.1.2. Dislocation Interaction Forces Since all the twinning dislocations within the twin layer shown in Fig. 3.1 are the same type of dislocations, i.e., they are all the Shockley partials of l/6<112], the interaction force between any two twinning dislocations is repulsive. If we calculate the interaction force resulting from dislocations located in front of a specific dislocation, the result will be the backward dislocation interaction force (FM). Similarly, the force calculated from the dislocations behind a specific dislocation is the forward dislocation interaction force (FM). 132 Since the distance between two neighboring twinning dislocations near the twin tip are small compared to their lengths as shown in Fig. 3.1, the twinning dislocations are considered as parallel dislocations, and the dislocation end effects can be neglected. Therefore, the repulsive interaction force between two twinning dislocations can be expressed as _ G . G . . Fd- 2n(1_v)R [b,xC) (b,> t212 n — + 2 b*cosza (X,- Xm) ' . (4.2) i=m+1 2“ (Xi-Xm)2+[dt,p,*(1-m)]2 where, a is the angle between the Burgers vector and the dislocation line, d”, is the interplanar spacing of the twinning plane, x,,, is the location of the mth twinning dislocation, x,- is the location of the ith twirming dislocation that is behind the mth twinning dislocation, and the forward dislocation interaction force FM is normalized with Gb so that F,,,,,/Gb is dimensionless. For the backward dislocation interaction force FM, we can get a similar expression, but only difference is that the dislocations ahead of the mth dislocation should be considered in this case: 134 F... j“ mine, (xm-xa * 1 (xm-x.) 2- (d... * (In-1') >21 Gb 1:, 2r(1-v> [(xm-X,)2+(d,,p,*(m-i) >212 :2: “€032“ (Km-Xi) . (4 .3) 1:1 21! (Xm-x2)2+[dt.p.*(m_i)]2 4.1.3. Internal Friction Force For the internal friction force, we consider only a lattice friction force that is due to the Peierls (or Peierls-Nabarro) stress, Up, on the twinning dislocations. The stress necessary for a dislocation to pass from one Peierls potential valley to the next has been calculated by Peierls ““1 and Nabarro “‘71 as a = orp sin(21ru/a) (4.4) where a, is the Peierls stress, 11 is the displacement of a dislocation perpendicular to itself, a is the distance between one close-packed atomic row and the next. The Peierls stress is directly related to the Peierls energy, as shown in Fig. 4.2, and it is expressed as =2Wp= 29 -_4"5 4 5 ap ab (1_v)exp( ). ( . ) where Wp = the Peierls energy, 135 0' Gph ° ° Viv.) as “’b _Gp._ Figure 4.2 - (a) Variation of Peierls energy as a function of transverse displacement (u), (b) Variation of the lattice friction stress. (W o z Gb’). (After Fantozzi, Esnouf, Benoit and Ritchie “‘31). 136 r = the half-width of the dislocation characterizing its degree of delocalization, r = d / [2(l-v)] for an edge dislocation, and 5' = d / 2 for a screw dislocation, d = the spacing of the glide plane. According to the above equation for the Peierls stress, it can be concluded that the Peierls stress for a partial dislocation is much smaller than that for a perfect dislocation because the Burgers vector effect on the Peierls stress is in the negative exponential ““"49‘. The internal friction force F,ric on a dislocation per unit length then can be calculated in a normalized form Firic.= 2 ' _4“C 4 7 Gb (1_v)31n(21tu/a)exp( —b ). ( . ) 4.1.4. The Force Due to The Stacking Faults The forces due to the stacking faults, such as F5”, FSESF and F, are all P-’ numerically equal to the corresponding stacking fault energies, 75,“, 7SESF and 7”,, respectively. Since the intrinsic stacking fault energy is approximately equal to the extrinsic stacking fault “5°"5”, we have I33131? z Fsrasr- (4- 8) 137 For the force F”, exerted by the twinning plane surface tension, since the twinning plane energy is about one half of the stacking fault energy “5145”, it is numerically equal to one half of the magnitude of the force resulting from the stacking fault (FSISF or F353,): F”, = 1/2 FSISF = 1/2 FSESF. (4.9) 4.1.5. Applied Force The applied force is expressed as the following: Fa=oa(me*be+ms*bs) (4.10) or Fa=aab (cosa *cosp *cosy +cosa *cosy *C080) (4 . 11) where or, = the externally applied stress, m, = the Schmid factor for edge dislocation component, m, = the Schmid factor for screw dislocation component, bc = the edge component of the twinning dislocation, b, = the screw component of the twinning dislocation, or = the acute angle between the Burgers vector and the dislocation line, ,8 = the angle between the applied tensile axis and the direction of b,, 7 = the angle between the applied tensile axis and the twinning plane normal, 0 = the angle between the applied tensile axis and the direction of b,. 138 Thus in the case of an equilibrium condition, since the net total force imposed on a twinning dislocation is zero, the following equation is established: Fe + Ff.d. + Ffl'ic. + Fb.d. + Fsrsr + FSESF + Ft.p. = 0 (4-12) 4.1.6. Definitions of External Force, Forward Force and Backward Force For the analysis convenience, we define some new forces in terms of their acting directions and sources. At first, we define the force resulting from the externally applied stress, the locally concentrated stress and the residual stress as an external force, F“. Here, we assume that all these external stresses result in forces acting on a twinning dislocation in the same direction as the twin propagation direction, so that the external force, F“, is always parallel to the twin propagation direction in this analysis. Secondly, we define the sum of forces acting in the twin propagation direction, except for the external force, as a internal forward force or simply a forward force, F,. Finally, we define the sum of forces acting in the opposite direction to the twin propagation direction as an internal backward force or simply a backward force, F,,. The terminology "internal force" means that this force is an intrinsic character of a twin, in other wards, the internal force coexists with the twin, i.e. , no twin then no internal force. While the external force is different, whether a twin exists or not, this force still exists in the matrix. Thus we have the following equations for each twinning dislocation according to the previous analysis. For the first dislocation, 139 F,, = F,, (4.13) F, — F,,, (4.14) F, — F,,, + F,,SF (4.15) For the second dislocation, F,, = F,, (4.16) F, = F,,, + F,,,F, (4.17) F, = F,,, + F,,S, + F,,, (4.18) For the third dislocation, F,, = F,, (4.19) F, = F,,, + F539,, (4.20) F, = F,,, + F,, + F,,, (4.21) For the dislocations beyond the third one, F,,, = F,, (4.22) Fr = Fm. + Ftp.’ (4-23) F, = F,,, + F,,. + F,,, (4.24) In what follows, we will calculate all these three forces, F,, F, and F,,, within mmAdmg3un 4.2. Twinning Dislocation Distribution Within A Thin Twin Layer Let us take a thin twin layer A in Fig. 3.1 as an example to see how the twinning dislocations are distributed within the thin twin layer. For the determination of twinning 140 dislocation position, the image of this thin twin layer is further amplified to a magnitude of 120,000X, as shown in Fig. 4.3 (a) and it is schematically illustrated in Fig. 4.3 (b). The twinning dislocations in this twin layer are labeled using numbers as shown in Fig. 4.3 (b). The distances of these twinning dislocations from the leading twinning dislocation are measured and the results are listed in table 4.1. For the second twinning dislocation, the distance was determined by taking the one-third of the distance of the third twinning dislocation that was directly measured from the image, since the exact location of the second dislocation line was hard to be determined from the image. This estimation will be evaluated in section 4.6.1. The distances of twinning dislocations from the leading twinning dislocation were measured up to the sixteenth dislocation. Since the seventeenth dislocation is blocked by an extrinsic dislocation as shown in Fig. 4.3 (b), the distance of the dislocations beyond the sixteenth are not simply related to the forces analyzed before, and a force due to the dislocation reaction between the seventeenth and the extrinsic dislocations should be considered for the seventeenth dislocation. This effect can be seen in Fig. 4.3 (a) in which the spacings between the twinning dislocations behind the seventeenth are less than those just ahead of it due to the extrinsic dislocation blocking. So the extrapolated positions of dislocations beyond the sixteenth are considered when calculating the forward forces on the dislocations from 11 to 16. The directly measured dislocation locations vs. the dislocation number is plotted in Fig. 4.4, and the data fit the following equation, 141 Cobw— EE ES 2: 5 Sushi“. Samoa—fie 93:53“ 05 mo wEBfie ouufionom a mm 3v new Home. 5.5 £5 a .20 0925 353.5 2: mm 3 - m6 BEER 3 142 88588 n .v 8.9% a: 333.3% ,3 598036 ommcmuxm 143 Table 4.1. The distance of twinning dislocations from the twin tip Dislocation Distance (A) Dislocation Distance (A) l 0 12 2803 2 48 13 3247 3 145 14 3787 4 290 15 4333 5 480 16 5167 6 694 17 6000 7 916 18 7000 8 1249 19 8200 9 1582 20 9600 10 1915 21 11400 11 2359 144 25:163.04-179.69*i+66.71*iZ-4.47*i3+0.14*i4. (4.25) where xi the distance of the ith dislocation from the twin tip, i = the dislocation number. Thus the extrapolated positions of the dislocations beyond the sixteenth dislocation are obtained by extrapolating the plot in Fig. 4.4. The extrapolated distances of the dislocations beyond the sixteenth are also listed in table 4.1. 4.3. Force Calculation 4.3.1. Calculation of Forward Force F, According to the previous analysis, the forward force on each twinning dislocation can be expressed as follows. On the leading twinning dislocation, Far = Ff.d. = b*sin2a Z: (xi-X1) *1 (X,-x,)2-d2,,,,,*(i—1)2] Gb Gb 211(1-v) 1=2 [(xi-x1)2+d2(111)*(i-1))212 n + b*cosza (x,-x,) 2n 7 , (4.26) 1=2 (xi—x1)2+d2(111,*(1-1)2 0n the second twinning dislocation, 145 it 5.5 2: 2 «8&8 £5 Bounce—m6 wedge: me 8:82 25. - 4.4 .8sz c3852 8:82me r r 1 I'd coca (v) do. Utmr 1119er 991191810 146 F£.2 = Ff.d. + Fsrsp': b*sin2a Zn: (Xi-X2) * [ (Xi-X2) 2-d2(111) * (i-2) 2] G'b Gb Gb 21t(1-v) 1:3 [(Xi-X2)2+d2(111)*(i-2))2]2 b__*_2_cosza ____12:3(X (Xi-x2) + YSISF (4.27) = X1""’{2)2+d2(111)*(in?)2 Gb On the third twinning dislocation, Ff.3=Ff.d.+FSESF_ b*sin2a n 4(Xi‘x3)*[(Xi-X3)2‘d2(111,*(i-3)2] Gb 019 Gb 21t(1-V)l [(xi_x3)2+d2(111)*(i-3))2]2 n + b*cosza (Xi-x3) + YSESF _ (4 . 28) 21: 1=4 (Xi-X3)2"'d2(111)”Hi-3)2 Gb On the mth twinning dislocation, i.e., on any twinning dislocation beyond the third twinning dislocation, Em: Fm. + Ftp. = b*sin2a f (xi-x» * t (xi-x,» 2‘d2(111>* 2: Gb Gb Gb 21I(l-V) i=m+1 [(xi_xm)2+d2(111)*(i_m))ZJZ n + b*cosza (xi ‘Xm) . + YSISF . (4 . 29) 2“ i=m+1 (Xi-Xm)2+d2(111)*(l-m)2 26b where m 2 4. 147 In the previous chapter, it has been proved that the twinning dislocations in the twin layer shown in Fig. 4.3 (a) are edge Shockley partials 1/6[112], so the angle between the Burgers vector and the twinning dislocation line (a) is 90°, and the magnitude of the Burgers vector of twinning dislocations (b) is 1.633 A. The interplanar spacing of the twinning plane (dam) is 2.3166 A for the stoichiometric composition. The shear modulus G = 69620 MPa, and Poison’s ratio r = 0.265 are from [154]. The superlattice intrinsic stacking fault energy (7313,.) is equal to 70 mJ/m2 according to the work done by Hug, Loiseau and Veyssiere '6'. Therefore, the superlattice extrinsic stacking fault energy (735”) is 70 mJ/mz, and the twinning plane energy (7”,) is 35 mJ/mz. If we replace all these values and dislocation positions in table 4.1 into above equations, we have n 2 ' F x»: x--S.3666* 1-1 2 f.1=0.3537 J. [21 ( ) 1 (4.30) Gb 12:2 [Xi+5.3666*(i—1)212 Ff-2=o.3537zn: (xi-48)*[(xi_48)2-5'3666*(i-2)2] +6 1571x10'3 Gb 1:3 [(x3-48)2+5.3666*(i-2)2]2 (4.31) n 2 - 2 x--145 * x.-145 -5.3666* 1-3 ( 1 ) [( 1 ) ( )] +6.1571x10'3, F £3=o.3537 . ab 1:4 [(x3-145)2+5.3666*(1-3)2]2 (4. 32) 148 ’1 (xi-xm)*[(xi-xm)2-5.3666*(.i-m)2] h=o.3537 E . Gb i=m+1 [(Xi‘Xm)2+5.3666*(1-m)2]2 +3.0786x10‘3. (4.33) where m 2 4. When calculating the forward forces, we considered only six dislocations beyond the dislocation concerned, that is, n - m = 6 in the above equations, because the interaction force between two dislocations separated with a large distance can be neglected. The calculated results are tabulated as shown in table 4.2 and plotted in Fig. 4.5. The calculation was carried out up to the sixteenth twinning dislocation in the twin layer Shown in Fig. 4.3. For the last six dislocations, dislocations 11 to 16, we used some eXtrapolated dislocation positions in the calculation. 4-3 .2. Calculation of Backward Force F,, Before calculating the backward force, let’s first calculate the internal friction fOl‘ce Fm. The internal friction force on a unit length dislocation is Ffric.= 2 sin(2nu/a)exp(-ig—c-) . (4.34) Gb (1-v) Here the internal friction force is normalized with Gb. If we take the maximum value of sin(21ru/a), i.e., assume sin(21ru/a) = 1, the equation (4.34) can be rewritten as 149 Table 4.2 The forces on each twinning dislocation Dislocation Normalized Normalized Normalized force (F f/ Gb) force (Fb/Gb) force (F,,/Gb) 1 1.2598 x 10'2 6.1718 x 10'3 -6.4262 x 10’3 2 1.3324 x 10'2 1.3560 x 10'2 2.3550 x 10“ 3 1.1319 x 10‘2 1.2843 x 10‘2 1.5236 x 10‘3 4 7.2392 x 10‘3 8.2083 x 10‘3 9.6910 x 10“ 5 6.7571 x 10‘3 7.5630 x 10'3 8.0630 x 10“ 6 6.3764 x 10'3 7.3210 x 10'3 9.4460 x 104 7 5.6098 x 10'3 7.3130 x 10‘3 1.7032 x 10'3 8 5.5341 x 10‘3 6.2352 x 10'3 7.0110 x 10“ 9 5.3869 x 10‘3 5.9253 x 103 5.3830 x 10“ 10 4.9830 x 10‘3 5.7939 x 10'3 8.1090 x 10“ l 1 4.9235 x 10'3 5.3094 x 10'3 3.8590 x 104 12 4.8103 x 10'3 5.1606 x 10'3 3.5030 x 10“ 13 4.5376 x 10'3 5.0948 x 10‘3 5.5720 x 10“ 14 4.4683 x 10‘3 4.8440 x 10‘3 3.7570 x 10“ 15 4.0562 x 10'3 4.7521 x 10'3 6.9590 x 10“ 16 3.9493 x 10‘3 4.3423 x 10’3 3.9300 x 10“ I 150 3.3 SE 05 use? 5:335 888 25. - We 2..me As 3. 55 82m 8:55 8mm Son 83 coo». 8mm Son 83 coon 82 83 com a d......._._.d._._._.e1 88m 3:8me . a. / .o lllllll IIIII Illllllll 'l'IVOII-ll \Il'o I. 8.8m Beacon 38m uafioam 2 Na 3 .2 . (000W) 999105' PGZIWWON 151 FfIiC. = 2 _ 431‘ 4 35 Gb 1-vexp( b)' (' ) For an edge dislocation, g' = d / (2—2v), then the equation (4.35) becomes F'frric.= 2 exp[- 4nd ] (4.36) Gb 1-v 2b(1—v) We know that v = 0.265, d = 2.3166 A, b = 1.633 A, so the magnitude of normalized internal friction force (F,,,c,/Gb) is equal to 1.4723 x 10‘s. The direction of Fm, is Opposite to the twin propagation direction. This result indicates that the internal friction force is a constant for any twinning dislocation. The forces due to SISF, SESF and twin interface tension have been calculated in the previous section, that is, FSISF FSESF YSISF - = = =6.1571X10 3. (4.37) Gb Gb Gb Ft-P- = YSISF=3.O786X10‘3. (4-38) Gb 2Gb In the following, we will calculate the backward dislocation interaction force. Since the first (leading) twinning dislocation does not have any twinning dislocations ahead of it, there is no backward dislocation interaction force on the first twinning dislocation. 152 According to the equation (4.3), the backward dislocation interaction force on the second twinning dislocation is Fb.d.2 = b*sin2a (Xz—Xl)*[(X2—X1)2—d2(111)(2‘1)2] Gb 2n(1-v) [(xz-x1)2+d2(111, <2-1>212 + bcosza (X2_X1) . (4.39) 21: ("‘2'x1)2+dz(111)"‘(2‘1)2 Since 0: = 90°, b = 1.633 A, ,, = 0.265, d0“, = 2.3166 A, x, = 48 A, x, =0, equation (4.39) becomes Fb.d.2= 1.633 48*(482—2.31662) Gb 2*3.14*(1-0.265) (482+2.31662)2 =7 . 3877.3(10'3 . For the third twinning dislocation, Fb.d.3= b (X3-X2) *[(x3—x2)2_d2(1rr)] Gb 2N(l-V) [(X3_x2)2+d2(111)]2 + (x3-x1) *[(x3-x1)2-d2‘11“*22] (4.40) [(X3 _X1) 2+d2(111)*22] 2 Putting the known numerical values into (4.40), we have 153 Fb.d.3=0 353.7“4.8=1=(482-2.31662)+ 145* (1452-4*2.31662) Gb ' (482+2.31662)2 (1452+4*2.31662)2 =9.74.93x10'3 Similarly, for the fourth twinning dislocation, we have Fbg.d.4 =0 . 3537 *{ (X4'X3) * [ (X4'X3) 2'd2(111)] b [(x4-x3)2+d2(111)]2 + (X4_X2) * [ (X4'X2) 2-4""dz(111)] [(X4-X2)2+4*d2(111)]2 + (X4-x1) * [ (X4"X1)2_9*d2(111)]} (4 .41) [(x4-X1) 2+9 ’"dz(111)]2 Here x1 = 0, x2 = 48, x3 = 145, and x4 = 290. So the backward dislocation interaction force on the fourth twinning dislocation is equal to F b.d.4_ -2 —— —1 . 4461X10 Gb In general, the backward dislocation interaction force on the mth twinning dislocation (m 2 4) can be written as “7’1 (xm-xi) * [ (xm-x,) 2-5 . 3666* (m-i) 2] 1:1 [(x,,,—x,)2+5.3666=«(m—i)2]2 3E2L9=0.3537* Gb (4.42) 154 Therefore, the backward force for each twinning dislocation in the twin layer shown in Fig. 4.3 can be calculated as follows. Fb.1 ___ FSISF’+ Ffric. Gb Gb Gb =6 .1571X10’3+1.4723x10‘5 =6.1718X10‘3 Fb.2 = Fb.d.2 + FSESF+ Ffric. Gb Gb Gb Gb =7 .3877x10'3+6 .1571x10'3+1.4723x10'5 =1 . 356 OX10‘2 Fb.3 Fb.d.3 + Ft.p. + Ffric. Gb Gb G’b Gb =9 .7493X10'3+3 . 0786X10’3+1 .4723X10‘5 =1 . 284,3)(10'2 Fb.m= Fl).d.m+ Ft.p. + Ffric. Gb Gb Gb Gb (4.43) 155 m-l 2 ' 2 x -x. * x -x. -5.3666* m-1 1:1 [(x,,,-xi)2+5.3666*(In-i)2]2 +3 .O786X10‘3+1.4723X10’5 m—l x -X. * x -x. 2-5.3666* m-i 2 =0.3537 (m 1) [( m 1) ( )1 1:1 [(x,,,-xi)2+5.3666=l=(m-i)2]2 +3 .0933X10'3. where m 2 4. The calculated backward forces are tabulated in table 4.2 and plotted in Fig. 4.5. 4.3.3. Calculation of External Force F,,, According to the definition of the external force in section 4.1.6, we have g+m+n=0 MM) 01' F,, = Fb - F, (4.45) that is, the external force is numerically equal to the magnitude of backward force minus the magnitude of corresponding forward force. Thus, the magnitude of external force, F,,, on each twinning dislocation is expressed as follows, Fex.1= Fb.1'Fr.1 Gb Gb _ ._ 2_ 2 _0_ 3537231 (145 x)*[(145 x) 5. 3666*(3-1) ] 156 n 2 - 2 x * x- -5.3666* 1-1 =6.1718x10'3-O.3537 1 ( 1 ,( ) . (4.46) 1:2 [x1.2+5.3666*(.1-1)2]2 Fex.2=Fb.2'Ft.2 Gb G'b 118-xi»: 48-X1-2-5.3666* 2-1' 2 -o 35371::l( ) [( ) ( )]. (4.47) [(48-xi.)2+5 3666*(2- -i)2]2 i¢2 Fex.3 = Fb.3"Fr.3 Gb Gb -3 . 07 86X10‘3 . [(145—x,.)2+5 3666*(3-132]2 i¢3 (4.48) Fex.m_, Fb.m-Ff.m Gb Gb - 2- 2 035372 (xm MX)*[(X X1) 5.-3666*(m 1)], (4.49) [(xm -x )2+5. 3666*(m- -.z')2]2 1991!? where m 2 4. The subscripts' 'm" and " 1" represent the corresponding dislocation number. 157 The calculated results are listed in table 4.2 and plotted in Fig. 4.5. From the results, we can see that the external force on the first twinning dislocation is negative. This will be discussed later. 4.4. Stresses in The Thin Twin Layer Stresses in the thin twin layer are classified into three categories similar to the forces on the twinning dislocation: (i) forward shear stress, 1,, that produces the forward force (Ff); (ii) backward shear stress (or back stress), 1., corresponding to the backward force (F,,); and (iii) external shear stress, 1w corresponding to the external force (F,,). The external shear stress includes the externally applied stress, the locally concentrated stress and the residual stress within the matrix. Since the effect of stress normal to the twin plane on the twinning dislocation glide are negligible comparing to that of the shear stresses “5", we will not consider the normal stress in the stress calculation. Therefore, the above three shear stresses are simply called forward stress, back stress and external stress, respectively. The terminology "external" here means that the stress is from the sources outside the twin layer, that is, it exists in the matrix even before the formation of the twin layer. However, the internal stress (either forward stress or back stress) is an intrinsic stress that is related to the formation of the twin, that is, it results from the internal sources such as twinning dislocation interaction, twin interface tension, and lattice friction on twinning dislocation motion. The shear stress and the force on a dislocation per unit length, which is 158 perpendicular to the dislocation line, are correlated in the form of F=T'b (4.50) Therefore, the shear stress 1' can be easily obtained by dividing the magnitude of force (F) by the length of Burgers vector (b), i.e., 1=Wb AM) This calculated shear stress is the stress at the location of the corresponding twinning dislocation. In what follows, we will list all formulas used in the calculation of shear stresses. The calculated results will be shown in table 4.3 and Fig. 4.6. For the forward stress 1, n 2 - 2 x * x--5.3666* 1-1 c,,,=24624.29 1 [21 ( ) 1, (4,52) 1:2 [X1+5.3666*(i—1)2]2 ’1 (x,-48)*[(xg-48)2-5.3666*(1-2)21+428 66 t =24624.29 f.2 12:3 [(xi—4,8)2+5.3666=I‘(i-2)2]2 (4.53) 159 ’1 (xi-145)*[(xi-145)2-5.3666*(i-3)2] rf3=24624.29 , +428.66, ' 1:4 [(xi-145)2+5.3666*(1-3)2]2 (4.54) ’1 (x.-x ) 4 [ (x.-x )2-5.3666* (i-m)2] tfm=24624.29 2 1 m 1 m , +214.33, ' i=m+1 [(Xi‘Xm)2+5.3666*(1-m)2]2 (4.55) where m 2 4. The subscripts indicate the correspondent twinning dislocation numbers. This notation is also applicable to the following formulas. The unit of stress is MPa, and this is the same in the following equations. For the back stress 1,, =Fb.1 =429.68(MPa), (4.56) 1'61 b‘ "11 b'2=944.01(MPa), (4.57) 1"19.2 b‘ F Tb.3= b'3=894.10(MPa) , (4,53) (3‘ m-l (Km-Xi) *[(xm-xi)2-5.3666*(m-i)2] 1:1 [(Km-xi)2+5.3666=I=(m-i)2]2 tbom=24624.29 +215.36, (4.59) 160 Table 4.3 The stresses at each twinning dislocation Dislocation 1, (MPa) 1., (MPa) 1,, (MPa) 1 877.07 429.68 -447 .39 2 926.92 944.01 17.09 3 781.97 894.10 112.04 4 503.99 571.46 70.94 5 470.43 526.54 57.10 6 443.92 509.69 65.68 7 390.55 509.13 118.12 8 385.28 434.09 48.09 9 375.04 412.51 36.40 10 346.92 399.55 55.44 11 342.77 369.64 25.84 12 334.89 359.28 23.36 13 315.91 354.70 37.77 14 311.08 337.24 30.33 15 282.39 330.84 47.41 16 274.95 302.31 26.33 161 .53— 52: E5 2: E newsstamme 838: .825 - 9v DBME g E. 55 58m 852a 8mm 8m 83 89. 8mm 88 8mm 8N 82 8~ 08 c 8N7 _ . _ q _ . _ . _ . _ . _ a _ q _ . A mmobm Efiouxm g a §§§§§§° (Haw) ssans mans 162 where m 2 4. For the external stress 1“ n 2 - 2 X-* x. -5.3666* 1-1 tx1=429.68-24624.292: 1 (21 ,( 2: . 1:2 [X1 +5.3666*(1-1) ] (4. 60) 418-31. * 48-x- 2-5.3666* 2-1’ 2 44621291131311‘ 1) H 1) ( H ex 2 [(48-xi)2+5.3666*(2-i)212 1'42 I (4.61) 145-x. ... 145—x. 2-5.3666* 3-1‘ 2 =.24624 291;: ( 1) H 1) .( ) 1-214.33, 3 [(145-Xi)2+5.3666*(3-1)2]2 i? 1’-ex.3 (4.62) _ 2- 2 {24624 2912 (X, -x,-,,,)*[(x x) 5. 3666*(m- 1) ] =1 [(29,721)‘°‘+5.3666=o=(m-1‘)2]2 11m (4.63) 4.5. Simplified Equations for F,, F,, and F,,, Let’s see an example how the equations are simplified. The equation (4.26) can be rewritten as 163 _ dz<111)"‘(-7:-1)2 Ff.1_ b*sin2a n 1 (Xi-X1)2 Gb 21t(l-V)12X--X1 [1+ d2(111)*(i—1)2]2 (Xi—X1)2 n + b*cosza Z 1 1 2n 1=2 xi-Xl 1+dz(111)*(-7:-1)2 (4'64) (Xi-X1)2 Since d2(111)*(i-1)2 (xi-x1)2 the equation (4.64) can be reduced as _F__f.1_[ b*31n2a batcos2 a] (4.65) Gb 21t(1-v) 1:21: In a similar way, we can simplify all equations used for the calculation of Ff, Fb and Fe, as follows. 164 For the forward force Ff n1 .7312=:2?, 8 +6 .1571x10‘3, F n 1 455:0.35372 _ 1=3 X1 F M=0. 3537 2 .0786x10‘3. Gb i_m+1x For the backward force F,, F 111-1 1 “=0 . 3537 E +3.0933x10'3. where m 2 4. For the external force Fe, II .1_ -3 1 ——-6.l718X10 -O.3537 — Gb 12:2 X1 F13 1 - b 2 x.-145 (4.67) (4.68) (4.69) (4.70) (4.71) 165 n Fexz l —-—=o.3537 , (4.72) i¢2 n F 453:0.35372 _L_—3.0786x1o-3, (4.73) G'b 1=1 145-x,- i¢3 n F ex.m=0.3537z 1 ’ (4.74) Gb 1=1xm-xi iatm wherei _>_ 4. The results calculated using these simplified equations are shown in table 4.4. The comparison between the results calculated from the original equations and the results from the simplified equations, as shown in Fig. 4.7, indicates that the results from the simplified equations can perfectly represent the results obtained from the original equations. The deviations between the results from the simplified equations and the results from the original equations are very small, only about :I: 0.024. So we can directly use the simplified equations in the calculation of forces on the twinning dislocations. 166 Table 4.4 The forces calculated using simplified equations Dislocation F r/ Gb Fb/Gb F,,/Gb 1 1.2660 x 10'2 6.1718 x 10‘3 -6.4882 x 10'3 2 1.3039 x 10'2 1.3644 x 10'2 6.0500 x 10“ 3 1.1322 x 10'2 1.2901 x 10'2 1.5790 x 10“ 4 7.2409 x 10'3 8.2139 x 10'3 9.7300 x 104 5 6.7583 x 10‘3 7.5663 x 10‘3 8.0800 x 10“ 6 6.3773 x 10'3 7.3230 x 10'3 9.4570 x 10‘4 7 5.6102 x 10'3 7.3151 x 10‘3 1.7049 x 10'3 8 5.5345 x 10'3 6.2365 x 10'3 7.0200 x 10“ 9 5.3871 x 10‘3 5.9257 x 10'3 5.3860 x 104 10 4.9831 x 10'3 5.7944 x 10‘3 8.1130 x 10“ 11 4.9235 x 10'3 5.3096 x 10‘3 3.8610 x 10“ 12 4.8104 x 10'3 5.1607 x 10‘3 3.5030 x 10“ 13 4.5376 x 10‘3 5.0950 x 10'3 5.5740 x 10“ 14 4.3936 x 10'3 4.8442 x 10‘3 4.5060 x 104 15 4.0562 x 10'3 4.752 x 10'3 6.9580 x 104 16 3.9493 x 10‘3 4.3423 x 10'3 3.9300 x 10“ 167 .8823 woman—5m 2: So: confine 3:52 .23 823:3 REES 2: 80¢ 3538 3:58 .«o anemia—coo .. 56 0.5me 3 as. 55. 68m Dogma 8mm Son 89. 83 8mm 89.. 8mm 88 82 88 9% c 88m @855 1 8.8m cafioam 35mg x Sufism o (0001):) 39010:} pezneuuo N i 168 4.6. Discussion 4.6.1. Stress and Force Distributions Within The Twin Layer From the previous results, we can see that the force distributions along the twin layer is similar to the stress distributions. So we will analyze only the stress distributions in this section, and this analysis is also applicable to the force distributions. (3) External Stress 7,, External stress has been defined as the stress resulting from the externally applied load, the local stress concentration and the residual stress within the matrix. Since the specimen was free from external load during the TEM investigation, the externally applied stress term can be eliminated from the external stress. Investigating the external stress line in Fig. 4.6, we can see that the external stress is more or less evenly distributed along the thin twin layer, which indicates that no stress concentration exists within the region of the twin layer investigated. Also, no stress concentration would be expected in the middle of a grain. This means that the external stress calculated in this study may be purely a residual stress within the matrix where the twin layer was growing. It is possible that the twin layer investigated has propagated long enough so that the front portion of twin layer, which we are studying, has been out of the range of local stress concentration that is at the grain boundary, as shown in Chapter Three. The negative value of external stress at the twin tip may be due to the wrong values of back stress and forward stress, which will be analyzed in the following section. So taking an average of the calculated external stresses with neglect of the stress at the fist twinning 169 dislocation, the residual stress (1,.) in the matrix is 7,, = [Tex],,,_ = 51 MPa (4.75) where [1”],v, means the average value of external stresses except for the stress at the first twinning dislocation. We assume now that the external stress, equal to the residual stress in this case, is distributed evenly along the thin twin layer including the first twinning dislocation. Then the driving force resulting from the external stress is the same for all the twinning dislocations in the thin twin layer of Fig. 4.3 (a). (b) Forward Stress r, The forward stress, which is an internal stress resulting from the dislocation interaction and the twin interface tension, decreases as the distance from the twin tip increases, as shown in Fig. 4.6. This decrement of forward stress is remarkable near the twin tip up to about 1000 A, which corresponds to the location of the seventh twinning dislocation, and the decrease becomes gentle beyond about 1000 A. The forward stress tends toward a constant value far from the twin tip. This indicates that the driving force for the twinning dislocation glide resulting from the forward stress is different for the dislocations located at different places in the twin layer. The driving force necessary for the twinning dislocation glide is very large near the twin tip, and it drops quickly as the distance increases. Therefore, the twinning dislocations having large distance from the twin tip glide forward more easily than the twinning dislocations near the twin tip. 170 The forward stress at the first twinning dislocation is smaller than that at the second twinning dislocation, as seen in Fig 4.6 and in table 4.3. This is possibly due to the incorrect approximation of location of the second twinning dislocation. So if we extrapolate the plot of forward stress using the data from the third dislocation to the sixth dislocation, we have the forward stress at the first and the second twinning dislocations equal to 1245.6 MPa and 1059.4 MPa, respectively. Thus the modified forward stress distribution along the thin twin layer is as shown in Fig 4.8. The corresponding forward force distribution is shown in Fig. 4.9. (c) Back Stress 1,, The back stress is an internal stress, which results from the dislocation interaction, the twin interface tension and the internal friction on the twinning dislocations. The change of back stress with the distance is similar to that of forward stress, that is, back stress drops very quickly with the distance near the twin tip, and decreases gently with the distance when the distance is large. Since the back stress results in a force (backward force) opposite to the twin propagation direction, the back stress resists the twinning dislocation glide, so it must be overcome in order for twinning dislocations to move forward during the twin propagation. So the resistance to dislocation glide in the twin layer is very large near the twin tip and relatively small at a location far away from the twin tip. In other words, the trailing twinning dislocations have less resistance for glide than the leading twinning dislocations, so the trailing twinning dislocations more easily glide comparing to the leading twinning dislocations. This conclusion is the same as that drawn from the forward stress analysis. 171 .323 5:33 gnazacflom mam: 6823—8 32% £56m a mm a .o define—been“ SEER 35590 8a 83 9.05. .n .x 9223. a 2.2%: + _x 23. - 2.32 n 4.: :38ch 05 E 5% 32F .m ”882 2-225; £2332; 19mg .8; an: ”-225; “.2493 ".2422 3% 2K- .2? m «-2232 2-22312 "-2483 2:: $2 .43: N ”-0303 482%; ”beam: 0.82 $3 .932 _ 81.6 8 if geese 3.64.: @334: 3:64.: 1.88225 823856 mafia}: 8:: “we 2: HE 888 ES 388% 85968 23. We 2nd. 172 3.3 :35 £5 05 wee? ecu—.586 3.8:. 35.58 25. - w .v Bawfi 2V 5 55. 58m 83:5 8mm coon 89. 83 8mm 08.... 8mm coon con“ 83 com o m . _ q . 1 4 _ . . J _ q . 4 - . q 1 _ m 8N7 m. .m 82- 8.- ..,.. 4.. cs- .. 386 352.3% ..... 84. m o m f 84 . 80 385 Beacon \ mmobm Egfioam §§§§§ (9cm) $89118 mm{S .53 E: as 2: was Susfima 85. 8588 2F - 3 «as A3 a? EBB 88m 8835 173 8mm coon oomv gov comm ooom comm coo“ 83 83 con c u — d - - d - u - d 1—‘ d d a _ d - d o u 0 O O O. O O O O O O O O 08v . a x . u + a ._x. m a x x .. m X . + + + x x . u + + “A. X . . + . .. - S H 305 3:28“ 8 5 88m 0 x” .. +- . 088 Baaxoam + . 2 n 088 Eanm x _- . — p — _ n — . _ . — _ . _ p — . (000 I x) 890105 WIIBUUON 174 Comparing the back stress with the forward stress on the first twinning dislocation in the twin layer as shown in Fig. 4.6 and in table 4.3, we can see that the back stress is smaller than the forward stress. If this were true, the first twinning dislocation should glide further forward until the forward stress at the first twinning dislocation was equal to the back stress. But this is not what we observed since the twin tip was stationary. Therefore, there may exist a very large internal friction on the first twinning dislocation (or say for the twinning dislocations near the twin tip). This internal friction on the dislocations near the twin tip should be much larger than that on the dislocations located far away from the twin tip. The magnitude of internal friction stress at the dislocations near the twin tip are equal to the difference between the forward stress, the residual stress and the stress due to the stacking fault (1”), the back stress (1“) i.e., Tfricm'p = 7-f + Trs - TS.F. - Tb.d.’ (476) where mm, is the internal friction stress on the twinning dislocations near the twin tip. The corresponding internal friction force on the dislocations near the twin tip is equal to the difference between the corresponding driving force (the forward force plus the force due to the residual stress) and the backward forces due to the stacking fault and the dislocation interaction, Ffric,tip = Ff + Frs - FS.F. ' Fb.d., (477) where PM,” is the internal friction force on the twinning dislocations at the twin tip, F,, 175 is the force resulting from the residual stress 1“,. So the modified back stress and the modified backward force should be Tb,tip = Twp + Tex = TfJip + Tm (4.78) Fb,tip = Fftip + F", (4.79) 01' Fb.tip = b Tb.lip' (4.80) Here the subscripts "f.tip" indicates the modified forward stress and modified forward force at the twin tip. The modified forward stresses at the twin tip have been calculated for the first three twinning dislocations in the previous section, so the modified back stresses and the modified backward forces for the first three twinning dislocations can be easily calculated using equations (4.78) and (4.80). The results of the modified stresses and the modified forces for the first three twinning dislocations are listed in table 4.5. The plots for modified stresses are shown in Fig. 4.8 and for modified forces in Fig. 4.9. 4.6.2. Dislocation Distribution in The Case of Equal Backward Force or Equal Driving Force on Each Twinning Dislocation At an equilibrium condition, the backward force on a twinning dislocation should be equal to the driving force on the same twinning dislocation. Thus the equal backward force condition is equivalent to the situation of equal driving force when we consider the 176 twinning dislocation distribution in a twin layer. If we assume the backward force on any twinning dislocation is constant, then the twinning dislocation distribution within a twin layer will be uniform, i.e., the spacing between any two adjacent twinning dislocations will be constant, as shown in Fig. 4.10. Here we selected a normalized backward force equal to 1.18 x 102. The result shows that if the backward force on every twinning dislocation is equal to 1.18x10F2 Gb, the spacing between any two adjacent twinning dislocations, except for the that two twinning dislocations, is equal to 164 A. This result is inconsistent with the experimental observation shown in Fig. 3.1 and Fig. 3.5. So this result indirectly proves that the backward force (and the forward force) along a twin layer is not constant. 4.6.3. Effect of Dislocation Location on The Stress Distribution Fig. 4.11 (a) shows how much the deviation of the second twinning dislocation location from the measured position affects the back stress on this dislocation. We selected :1: 10 A deviation from the measured value 48 A. This deviation of the dislocation location results in either 131 MPa stress increment when the dislocation is at the left side of the measured position or 87 MPa stress decrement when the dislocation is located at the right side of the measured position. This indicates that a small deviation of the dislocation location results in a large variation in stress on the second dislocation. Location of the third twinning dislocation has a similar effect on the stress as the second one, as shown in Fig. 4.11 (b), however, the effect is less than the second one. In Fig. 4.11 (b), the deviation of j: 10 from the measured value, 145 A, results in either 42 MPa 177 f: x w: n 620 8:825 39:3 85 8 85... 3303 :58 a segues 888:5 - 2 .4 2&5 20 ea 55 see 8525 00m_. _ 000.. 00m 0 . kind: I No.0 H N r 0 I .v0.0 w - . m. N 80082 50803:. maniac. H W H. I 00.0 m. . . O - . 9 I 1 00.0 H < a: H r . p . p _ h p p . _ . h p . .. F.o 178 increase in stress on the right side of the measured position or 34 MPa decrease in stress on the right side of the measured position. If we look at the tenth dislocation, as shown in Fig. 11 (c), the effect of deviation of dislocation location from the measured position on the stress is very small. :th deviation results in only +4 or -3 variation in stress. Even at the deviation of $50 for the tenth twinning dislocation location, the stress change is only +20 or -15 MPa that is still smaller than those for the second and the third twinning dislocations at the deviation of j; 10. Therefore, the determination of twinning dislocation positions is very important for the dislocations near the twin tip. For the dislocations far away from the twin tip, a slight deviation from the correct position does not change the stress much. This is reason why the stress curves in Fig. 4.6 are rough near the origin of the distance axis (near the twin tip). 4.7. Summary The distributions of forces and stresses along a thin twin layer was numerically calculated based on the dislocation theory and the morphology of twin nucleation and propagation analyzed in previous chapter. The forces acting on a twinning dislocation are classified into three categories: forward force, backward force and external force. The corresponding stresses are forward stress, back stress and external stress. The calculation was carried out based on the twinning dislocation locations in an experimentally observed thin twin layer. The results show that both forward stress and back stress are very large at the twin tip and drop quickly as the distance from the twin tip increases. The external stress is the residual stress in the matrix, which is uniformly 179 dogging 5:8 05 .80 60 new .5080ng .55 05 .80 80 623858 2.88 2: 8m 3 .393 3.563 2: :0 5:82 serge—fie wage?» 05 we Hootm - :é DEER 3 g .5. 53... see 83am 02 8“ mm 0n n . 0 q . . a a . 1 . a _J q . . . . a a a — . . 1 a a . J . q 25“! .0 H H m . «mm 0 I - V4. I .. 82 S . .x w 1 . 0.2: S I l S r n ) 1 .. o8~ W 1 . .d . . B 1 n ( I I 08m r . P . p b . . h w _ . . L p b n p p p — p . . h p . u h p r g 180 (BdW) 559118 51398 IIIIIITTTITIIIIIT I ITTT )— _- l- b mi In . Ft- m1 . W. H .4..:. ..1. ml . 4.L O § § §gv § 0 m N fig“ '7 300 250 200 150 100 50 Distance From twin tip (A) (b) (Figure 4.11 continued) 181 826:8 :4 Bare 3 .20 .5 55. 62m 88:5 000m 00mm 00.3 00mm 00mm 003 08m 003 0000 00: 003 009 60¢ é (8cm) sssns pmmxava 182 distributed along the twin layer. Comparison between the original equations and the simplified equations for the calculation of forces and stresses shows that there is a very small difference between them. CHAPTERFIVE NIECHANICAL TWINNING DURING CREEP DEFORMATION IN TiAl 5 .1. Introduction In recent years, much work has been done to understand deformation mechanisms of TiAl in short term properties both at room temperature and at high temperature """55‘. But for long term creep deformation, much less information is available in the literature. Also, there are two differing results concerning mechanical twinning in creep deformation of TiAl. Loiseau and Lasalmonie 92' investigated creep deformation of equiaxed single phase 7 Ti.,.,A154 and found that mechanical twinning was an important creep deformation mechanism at deformation temperatures up to 800 °C. Huang and Kim [39‘ studied creep behavior of two phase 7+oz2 alloy with composition of Ti—47.0Al— 1.0Cr-1 .0V-2.5Nb at 900 °C and observed no evidence of mechanical twinning in creep deformation. Since one goal of TiAl components will be to replace nickel and cobalt base superalloys in aircraft applications, understanding creep deformation at high temperatures is necessary. However, creep deformation mechanisms are not clearly identified: (1) The activation energies for creep are much larger than those for self- diffusion and interdiffusion in TiAl [37""”', which suggests that the creep rate may be controlled by processes other than the usual lattice diffusion mechanism; (2) The reported 183 184 value of the stress exponent varies widely from about 2 to 8 “3'44"56'157'. This indicates that several deformation mechanisms are involved in the creep of TiAl. The details of deformation mechanisms are not known and the limited results in the literature are not consistent with the creep theory. Mechanical twinning has been found to be an important component of creep deformation in TiAl ”2'2””. It is possible that a parallel creep deformation mechanism exists in the case when the mechanical twinning and some other creep deformation mechanisms such as dislocation movement, diffusion, recovery, and so on, occur independently. Therefore, it is important to understand the mechanical twinning behavior during creep deformation and its contribution to the creep deformation in order to understand and therefore improve our ability to optimize this material. In this chapter, observations of mechanical twinning behavior during creep and some contributions of mechanical twinning to creep deformation in near-y TiAl are presented. The results are analyzed based on the twin nucleation and propagation theory proposed in the previous chapters and in terms of a maximum resolved shear stress criterion for mechanical twinning proposed in this chapter. 5.2. Material and Experimental Procedure The creep specimens were produced and machined at Howmet Corp. , Whitehall, Michigan. Investment cast test bars, 16 mm in diameter and 125 mm long, were HIP’ed before specimens with 25 mm gage length and 5 mm diameter were machined from the bars. After HIP’ing, the test bars were heat treated at 1300 °C for 20 hrs in Ar 185 atmosphere, and cooled at 65 °C/min in argon to produce the equiaxed + lamellar microstructure. The nominal composition was Ti-48Al-2Nb—2Cr atomic percent. A constant stress test and a multi-stress drop test were conducted at 765 °C in air. A constant stress creep test at 176 MPa was interrupted at 4% strain, near the end of primary creep. The multi-stress drop test started at 276 MPa and the stress was dropped in increments to a final stress of 103 MPa at about 20% strain. The specimens were fumace cooled to room temperature while maintaining the final stresses. The temperature difference between the two ends of the specimens was less than 3 oC. Microstructural investigation on the specimens before and after deformation was carried out using optical and transmission electron microscopy. Specimens for optical microstructure were prepared by making longitudinal sections from the deformed specimens and the original bars. The specimens were etched using Kroll’s reagent to view grain structures after the specimens were ground and polished. For TEM investigation, however, 0.7 mm thick slices were cut in both longitudinal and transverse directions from deformed and undeformed samples. 3 mm diameter disks were cut from the slices using an ultrasonic cutting machine and ground to about 0.1 mm thick. The disks were finally thinned in a double jets electropolishing system using a 10% sulfuric acid + methanol solution at 20 °C. The TEM investigation was performed on a HITACHI H800 transmission electron microscope with an accelerating voltage of 200 kV. 186 5.3. Results 5.3.1. Optical Microstructure The initial microstructure after heat treatment was a duplex structure with (7+a2) lamellar colonies and equiaxed 7 grains, as shown in Fig. 5.1 (a). After creep Figure 5.1 - Optical microstructures (a) before creep deformation and (b) after creep deformation. deformation, cross twinning configurations were observed within lamellar colonies as indicated by the arrows in Fig. 5.1 (b). One possible form of mechanical twinning occurs parallel to the existing lamellar interfaces, which is denoted as "parallel twinning" 187 hereafter. But it is difficult to distinguish parallel twins from original lamellae from their images since both have the same image characteristics. 5.3.2. Cross Twinning and Parallel Twinning in Lamellar Grains The investigation in lamellar grains exhibited many cross twinning configurations. The cross twinning was found to occur when two differently oriented lamellae meet and either or both of the lamellae are oriented in such directions that twinning occurs parallel to the lamellar interfaces, as shown in Fig. 5.2 (a). Lamellar orientations with respect to the tensile axis were determined. For the inclined lamellae in Fig. 5 .2 (a), the interfaces were tilted 42° from the tensile axis. For the vertical lamellae, lamellar interfaces are tilted 5° away from the tensile axis. Many similar configurations were found in the multi-stress—jump creep specimen. At the intersection region of two lamellae in Fig. 5.2, many relatively fine twins were formed. Investigating the coarse lamellae that are located away from the intersection region, the orientation relationship between the two lamellae has true-twin relationship, as shown in Fig. 5.2 (b). For convenience, we denote "Lamellae 1" to the inclined lamellae and "Lamellae 2" to the vertical lamellae in the following. These two sets of lamellae are indexed in Fig. 5.2 (c) by taking the 7 phase matrix as a reference orientation and the indices are referred to 7 laths in lamellae. Therefore, the interfaces of Lamellae l are (111)l (subscription 1 indicating Lamellae 1); the interfaces of lamellae 2 are (l ll)2; and the intersection axis of two lamellar sets is perpendicular to the page and in the [110] direction, as shown in Fig. 5 .2 (c). 188 (a) Figure 5.2 - (a) Cross twinning configuration in a lamellar grain, (b) diffraction pattern of the original lamellae, and (c) orientations of the original lamellae. 189 (Figure 5.2 continued) 190 Lamellae 2 F [11111 Lamellae l [1'i'0] (C) (Figure 5 .2 continued) The cross twinning configurations shown in Fig. 5.3 is a different situation from that described above. The cross twinning in Fig. 5.3 were formed in an equiaxed 7 grain, and both approximately vertically oriented twins and slightly inclined twins were formed during creep deformation, which will be verified later. 5.3.3. Fine Mechanical Twins at Grain Triple Points In the specimen deformed to a strain close to the end of primary creep, we found that very fine mechanical twins frequently formed at equiaxed 7 grain triple points. Fig. 5.4 shows that these fine mechanical twins formed at the grain triple points within an equiaxed 7 grain, as indicated by letters A, B and C. The thickness of the fine twins in Fig. 5.5 (a) is about 50 nm. These fine mechanical twins are much thinner than typical Figure 5.3 - Cross twinning configuration within a large equiaxed 7 grain. 192 Figure 5.4 — Fine mechanical twin configurations at equiaxed 7 grain triple points. 193 Ia: 0=Tw|n 2 : (C) (b) Figure 5.5 — Fine mechanical twins (a), diffraction pattern across several fine twin interfaces (b), and twin relation in the diffraction pattern (c). 194 lamellar laths that form in phase transformation ‘22]. The diffraction pattern across these fine mechanical twins is shown in Fig. 5.5 (b) and (c). Fig. 5.5 (c) is a schematic drawing of the diffraction pattern of Fig. 5.5 (b). The open circles in Fig. 5 .5 (c) are the diffraction spots from the 7 matrix, and the rhombuses are the twin spots. In addition to these two spot sets, there also exist very faint spots in Fig. 5.5 (b), which are designated by dots in Fig. 5.5 (c). This set of diffraction spots represents the D019 crystal structure. This result indicates that two crystal structures exist in Fig. 5.5 (a): one is twin related 7 phase layers (the fine mechanical twins and the untwinned matrix), and the other is fine DO19 structure layers. The locations of these D0,, structure layers are not known in this case. These fine lath configurations will be called fine mechanical twins in this study. 5.4. Analysis and Discussion 5.4.1. Cross Twinning in Lamellar Grains If we look at a 7/7 lamellar interface, there exist six possible orientation relation- ships between the two adjacent 7 lamellae that can be described in terms of < 110> directions “53””. Among these six relations there are two kinds of twin relationships: if [110] directions in two 7 laths are anti-parallel to one another in the lamellar interface, these two adjacent 7 lamellae have a {111} < 112] true-twin relationship; if [110] direction in one 7 lath is anti-parallel to [011] or [101] direction in an adjacent 7 lath, a pseudo-twin will form ”‘3'”9'. 195 In a cast and heat treated specimen, the 7/7 lamellar interfaces with a {111} < 112] true-twin relationship are more favored than those with a pseudo-twin relationship, since true-twin relation has lower interface energy and is thermodynamically preferred during solidification and phase transformation. Fig. 5.6 shows interface structures of a true-twin in (a) and a pseudo-twin in (b). In Fig. 5.6 (a), the interface is a 23 boundary with a (111) boundary plane, and two laths are fully symmetric in the boundary plane (111). The interface of pseudo-twin in Fig. 5 .6 (b) is a 226 boundary, that is, only atoms of every second (111) plane are in completely symmetrical locations with respect to the boundary plane (111), while atoms of other (111) planes are in anti- symmetric positions, in which perfect symmetric sites are occupied by atoms of different type. This even 2 boundary contains antiphase boundary elements “6‘“. Therefore, the energy of lamellar interfaces with pseudo-twin relation is higher than those with true-twin relation. This is consistent with our observations. We found that original lamellae had true-twin relationship in the specimens investigated, as shown in Fig. 5.2 (b). For mechanical twinning, there are the same kinds of configurations as analyzed above. General twinning systems in 7 phase are in the form of {111} < ll2> , the same as for disordered fcc crystals. However, because of its anisotropic crystal structure, the different twinning systems result in either true-twins or pseudo-twins during deformation. If the twinning shear occurs in the direction of [112] on a (111) plane, the deformed crystal will be exactly symmetric to the undeformed crystal with respect to the (111) twin plane, which is a true-twin, as in Fig. 5.6 (a). If a crystal is deformed by either of (111x511) or (111)[1§1], the result will be a pseudo-twin, as shown in Fig. 5.6 (b). In 196 (b) Figure 5.6 - Atomic arrangement of true twinning (a) and pseudo twinning (b). 197 order to understand the twinning behavior in the larger strain creep specimen , the possible twinning systems in Fig. 5.2 (a) are plotted in a stereographic projection along [110] direction by considering the tensile axis orientation, as shown in Fig. 5.7, where the twinning system is assumed to be the {111} < 112> type. A resolved shear stress on each twinning system in Fig. 5 .7 is considered to determine the probable twinning systems in this particular orientation. Table 5 .1 shows the computed Schmid factors for these possible twinning systems, where d) is the angle between the tensile axis and the normal to the twin plane, A is the angle between the tensile axis and the twinning direction. In Lamellae 1, both (lll)[211] and (lll)[l2l] have relatively large Schmid factors, and are more likely to operate during creep deformation. In lamellae 2, all Schmid factors are small, so twinning is less likely. Therefore, the mechanical twins in Fig. 5.2 (a) are formed by operating either (lll)[2ll] or (lll)[l2l] or both twinning systems in lamellae 1. The mechanical twins in this case are pseudo-twins, and they were clearly formed during creep deformation, since this kind of twinning is not observed in undeformed specimens. Since these pseudo-twins are not thermodynamically favored, an interfacial energy criterion of mechanical twinning is unsuitable in this case. A maximum resolved shear stress criterion for mechanical twinning is appropriate. This criterion is that mechanical twinning operates in the twinning system with the highest resolved shear stress. However, if the maximum resolved shear stress is less than a certain critical value, the interfacial energy criterion may play a role. 198 Figure 5.7 - Stereographic projection of possible twinning systems along [110] direction. 199 Table 5.1. Schmid Factors for Mechanical Twinning Systems in Lamellae 1 and Lamellae 2. Lamellae l (lll)<112> <15 A COM (1 1 1)[1 12] 48° 84° 0.070 (lll)[211] 48° 45° 0473* (lll)[l21] 48° 52° 0412* (lll)[112] 48° 96° 0.070 lamellae 2 (111)<112> <15 A cosmos); (lll)[fi] 85° 53° 0.052 (lll)[fii] 85° 20° 0.082 (lll)[fil] 85° 66° 0.035 (lll)[112] 85° 127° 0052 * Probable twinning system Fig. 5.8 schematically shows the formation of cross twins in Fig. 5.2 (a) according to the maximum resolved shear stress criterion. Fig. 5 .8 (a) is a starting situation before twinning. Fig. 5.8 (b) is a metastable situation showing growth of mechanical twins in lamellae 1 into lamellae 2. To accommodate this growth, mechanical twinning on the other twinning system with a high Schmid factor is needed as shown in Fig. 5.8 (c). Most thick twins can easily traverse through existing thin 200 Lamella I n o 4 \\\\ (a) — (b) Figure 5.8 - Schematic mode of cross twinning: (a) initial condition of two lamellae; (b) metastable condition of twinning in lamellae l; (c) final configuration of cross twins as seen in Fig. 5.2 (a). (c) (Figure 5.8 continued) vertical twins. Therefore, mechanical twinning in Fig. 5.2 (a) occurred in the inclined lamellar colony parallel to the lamellar interfaces and across the vertical lamellar colony by twinning shear in (111)[2i_1] and (lll)[12l] systems. Therefore, this is a parallel twinning with respect to the inclined lamellae, but it is also a cross twinning with respect to the vertical lamellar colony. In Fig. 5.2 (a) the inclined lamellar laths become finer after twinning. This indicates that extensive mechanical twinning occurred during creep deformation in this lamellae. But vertical lamellar laths are unchanged. This is because the inclined 202 lamellar laths of lamellae 1 contain two possible twinning systems having the highest Schmid factor and the vertical lamellar laths of lamellae 2 do not have high Schmid factor twinning systems. This is consistent with the maximum resolved shear stress criterion. 5.4.2. Cross Twinning in Equiaxed 7 Grains The cross twinning configuration in Fig. 5.3 shows a different situation than that in Fig. 5.2 (a). In the case of Fig. 5.3, the cross twins were formed in an equiaxed 7 grain, and both approximately vertically oriented twins and slightly inclined twins are formed during creep deformation, which is easily verified by carefully investigating the twinning sequence of these two sets of mechanical twins. The inclined twins are sheared by the formation of vertical twins at their intersections. This indicates that the inclined twins formed before the vertical twins. However, the inclined twins are tapered as they grow into the grain interior, which means that these twins are formed by the twinning mechanism described in chapter three. A possible scenario for this sequence is as follows: The inclined twins were first formed due to a local stress concentration, and the vertical twins were generated by the operation of twinning system with a maximum resolved shear stress resulting from the external tensile creep stress ”8'. Comparing our result with others ”2'3” indicates that mechanical twinning is playing an important role in creep deformation at the combination of stress and temperature. Mechanical twinning reduces a stress concentration so as to maintain strain continuity. There may be a twinning transition temperature, below which mechanical 203 twinning occurs. In Loiseau and Lasalmonie’s study 1221, the transition temperature is about 800 °C for a complete equiaxed single 7 phase. But this transition temperature may change depending upon the composition and microstructure. Our observations of mechanical twinning in creep specimens are consistent with Loiseau and Lasalmonie’s result where the equiaxed 7 grain structure was investigated. In Huang and Kim’s study ‘39], the material was in the two phase 7+oz2 lamellar condition; the mechanical twinning did not operate, possibly due to the deformation temperature higher than the twinning transition temperature. At high temperature, since dislocations are much more mobile and some cube <100> dislocations become mobile, mechanical twinning is less favorable than that at temperatures below the twinning transition temperature. 5.4.3. Fine Mechanical Twins at Equiaxed 7 Grain Triple Points (a) Fine Mechanical Twins Formed by Accommodation of Stress Concentration at Equiaxed 7 Grain Triple Points Comparing a crystal orientation in the untwinned area with that in the fine twin region, we found that the matrix orientation in the fine mechanical twin region in Fig. 5.4 was exactly the orientation of the untwinned area. This suggests that the fine mechanical twins are formed by deforming the 7 phase. This is also evident in the configuration of the fine mechanical twins growing toward the 7 phase grain interior, which is located between the fine twin region and the untwinned region as indicated by arrows in Fig. 5.4. 204 The configurations of twinning accommodation of stress concentrations at equiaxed 7 grain triple points were also occasionally found in the specimens before creep deformation. However, these accommodation twins are widely spaced compared to the fine mechanical twins in crept specimens. These accommodation twins at the grain triple points in the specimens before creep deformation are probably formed either by the previous HIP’ing process or by the heat treatment or both. The deformation mechanisms in HIP’ing process are similar to the creep deformation mechanisms, for example, the grain boundary sliding and the diffusion accommodated deformation “6”. Therefore, the stress concentration at the grain triple points during HIP’ing process is probable. Since the cooling rate after heat treatment is relatively fast (65 °C/ min), the thermal stress concentration at equiaxed 7 grain triple points is also possible due to the anisotropic thermal contraction in 7 phase. However, the stress concentration formed during the HIP’ing or the cooling could not result in so intensive twinning as observed in the crept specimens. The shear displacement along the twin interfaces between two points "M" and "N" in Fig. 5.5 (a) is calculated to be 1.18 pm. Such a large displacement could not be caused by the thermal stress arising from the thermal contraction mismatch. The observed fine mechanical twins at equiaxed 7 grain triple points in this study are probably formed either by refining the largely spaced accommodation twins existing in the specimens before creep deformation, or by twinning in untwinned grains due to local stress concentrations at equiaxed 7 grain triple points during creep deformation. (b) Formation of Fine Mechanical Twins Prevents Grain Boundary Sliding Fig. 5.9 is an another example of fine mechanical twins formed at an equiaxed 205 7 grain triple point. These fine mechanical twins formed to accommodate the stress concentration caused by grain boundary sliding and/or the deformation of adjacent equiaxed 7 grains at the grain triple point (Fig. 5.9 (a)). However, this accommodation twinning caused the grain boundaries to become zigzagged instead of smooth (Fig. 5.9 (b)). The presence of curled dislocations in the untwinned neighbor equiaxed 7 grain suggests that subsequent creep deformation occurred by dislocation slip in the untwinned equiaxed 7 grain interior near the zigzagged boundaries (Fig. 5.9 (c)). All three of the grain boundaries that form the equiaxed 7 grain triple point in Fig. 5.9 (a) are large angle grain boundaries. The grain orientation relationships and boundary structures of these three grains (designated as grain I, II, III in Fig. 5 .9 (a)) were determined and are shown in table 5.2. The nearest coincidence boundary and the deviation from this boundary are indicated. The fact that these boundaries are high angle, and not low energy special boundaries, indicates that grain boundary sliding would have been relatively easy "‘2‘. These grain boundaries were smooth before twinning since they were equiaxed 7 grain boundaries, which can be seen if one looks at the same grain boundary in untwinned region. Table 5.2. Misorientation and Grain Boundary Structure of Grains at the Grain Triple Point. Grain boundary I/II II/III III/I Misorientation [120]/45° [164]/150° [in/132° Boundary structure 3° from 215 7° from 233,, 7° from 2331,, 206 Figure 5.9 - Fine mechanical twins from accommodation of local stress concentration (a), twin end configuration at equiaxed 7 grain boundary (b), and dislocation slip near the zigzagged boundary (c). 207 The zigzagged grain boundary resulting from mechanical twinning is much different from the serrated lamellar grain boundary in their morphology. The " serrations" of the zigzagged grain boundary formed by mechanical twinning are regularly distributed along the trace of the previous equiaxed 7 grain boundary, and the " serrations" are small, as shown in Fig. 5.9 (b). However, the typical lamellar grain boundary, the boundary between an equiaxed 7 grain and a lamellar grain or the boundary between two lamellar grains, formed from phase transformation, are roughly serrated, and the "serrations" are large and irregular, as shown in Fig. 5 .10 and 5.11. (c) On D0,, Crystal Structure in Fine Mechanical Twins The twinning mechanism in TiAl observed in this study is a homogeneous shear of a crystal by the glide of 1/6<112] twinning dislocations on every (111) close packed plane. However, if the stress concentration at the grain triple points and/or grain boundaries was such that it did not create 1/ 6 < 112] twinning dislocations on every (111) plane, four atomic layers of D0,,, structure, i.e. , a nucleus of the Ti3Al crystal structure, could be formed by glide of a 1/6< 112] twinning dislocation on a (111) plane “‘5'. In addition to this, the crystal structure across the true—twin plane in TiAl is also a three atomic layer DO19 structure. Therefore, the occurrence of DO19 diffraction spots with the twin related TiAl diffraction spots in the diffraction pattern in Fig. 5.5 (b) is reasonable. It is worth noting that such a uniformly distributed thin lath configuration in Fig. 5.5 (a) is very difficult to form through the phase transformation such as 012 --- > a2+7 or a --- > a2+7. Such a thin new phase layer, 7 phase in this case, could not be thermodynamically stable in the case of phase transformation, and therefore, the thin 7 208 Figure 5.11 - Optical microstructure showing the roughness of lamellar grain boundary. 209 phase layer should either grow and become thicker or be eliminated by growth of the adjacent 7 laths. The resultant configuration should be the lamellae containing coarse 7 laths with different thickness. The observation of fine twin configurations only near the equiaxed 7 grain triple points in this study also provides an evidence that the fine mechanical twins are formed due to the local stress concentration at the equiaxed 7 grain triple points but not by the phase transformation. 5 .5. Conclusions 1. Mechanical twinning is an important deformation mechanism in creep deformation of TiAl intermetach compound. 2. An energetic criterion is not suitable for mechanical twinning in this study. Above a certain value of stress (at constant temperature), the twinning transition stress is apparently temperature dependent. 3. Mechanical twins are formed by either true-twinning or pseudo-twinning. The formation of mechanical twins during the creep deformation of TiAl follows the maximum resolved shear stress criterion. 4. The mechanical twins observed in this study are propagated by homogeneous glide of 1/6<112] twinning dislocations on every close packed (111) plane, that is, by a homogeneous shearing mechanism. 210 5. The stress concentration at the grain triple points can be accommodated by the formation of fine mechanical twins. The zigzagged grain boundaries formed by mechanical twinning can inhibit grain boundary sliding. The formation of fine mechanical twins is probably controlled more by the local stress concentration than by an externally applied stress. CHAPTERSIX CONCLUDING REMARKS AND RECOMMENDED FUTURE WORK The in situ investigations of mechanical twin nucleation and propagation presented here have demonstrated that mechanical twin nucleates by bowing out of twinning dislocations at grain boundaries due to the local stress concentration and propagates by continuous emission of twinning dislocations from the grain boundaries and glide of twinning dislocations on every adjacent twinning plane. The nucleus of true-twinning is found to be either a superlattice intrinsic stacking fault (SISF) in the case that one 1/6[112] twinning dislocation is emitted from the grain boundaries in a (111) plane, or a superlattice extrinsic stacking fault (SESF) when two 1/6[112] twinning dislocations are emitted from the grain boundaries in two adjacent (111) planes. The mechanical twin propagation observed in this study occurs by trailing twinnning dislocations pushing leading twinning dislocations in the twin layer. The stress analysis on a thin twin layer has shown that the stress distribution along the thin twin layer is not even. The changes of forward stress and back stress are similar. The magnitudes of forward stress and back stress are very large near the twin tip and drop very quickly as the distance from the twin tip increases. At very large 211 212 distance from the twin tip, the stress changes are very small and both stresses tend to become constant. The calculation of external stress has shown that the magnitude of external stress is small and evenly distributed along the thin twin layer investigated. The external stress has been identified as a residual stress within the grain where the thin twin layer is located. The result of stress analysis indicates that the trailing twinning dislocations glide forward more easily in the twin layer than the leading twinning dislocations. This result is consistent with the experimental observation of the trailing dislocation pushing the leading dislocation phenomenon. The formation of various mechanical twin configurations observed in creep deformed specimens has been well understood, and interpreted using the twinning mechanism identified and the maximum resolved shear stress criterion proposed in this study. The cross twinning occurs when two differently oriented lamellar colonies meet in such a way that one lamellar colony with a larger Schmid factor generates twins across the other one. However, the cross twinning configurations are also observed in the large equiaxed 7 grains: one set of twins, which initiate at grain boundaries, are formed due to the local stress concentration and the other set of twins, which cross the first set of twins, are formed by the external loading. Analysis of the cross twinning in the lamellar colonies indicates that an energetic criterion for mechanical twinning is not suitable in this study, and the pseudo-twinning obeys the maximum resolved shear stress criterion. The fine mechanical twins at the equiaxed 7 grain triple points are due to the accommodation of local stress concentration during the creep deformation. 213 to fully understand the mechanical twinning and creep deformation in TiAl: 1. Correlations between mechanical twinning and creep deformation in TiAl, or the role of mechanical twinning in the creep deformation of TiAl. 2. What type of grain boundary structure is preferred for mechanical twin nucleation at the grain boundary. 3. Investigation of creep deformed single crystal TiAl in which the Schmid law can be precisely applied. LIST OF REFERENCES LIST OF REFERENCES [1] H.A. Lipsitt, D. Schechtman, and RE. Schafik, Metall. Trans. A, 6A (1975), 1991. [2] T. Kawabata, T. Kanai, and O. Izurni, Acta metall., 33 (1985), 1355. [3] SC. Huang, Scripta Metall., 22 (1987), 1985. [4] D. Schechtman, M.J. Blackburn, and H.A. Lipsitt, Metall. Trans., 5 (1974), 1373. [5] G. Hug, A. Loisseau, and A. Lasalmone, Phil. Mag. A, 54 (1986), 47. [6] G. Hug, A. Loiseau, and P. Veyssiere, Phil. Mag, 57 (1988), 499. [7] SA. Court, V.K. Vasudevan, and H.L. Fraser, Phil. Mag. A, 61 (1990), 141. [8] S.H. Whang and Y.D. Hahn, Scripta Metall., 24 (1990), 485. [9] B.A. Greenberg, O.V. Antonova, V.N. Indenbaum, L.A. Karkina, A.B. Notkin, M.V. Ponomarev, and L.V. Smirnov, Acta Metall., 39 (1991), 233. [10] B.A. Greenberg, O.V. Antonova, L.A. Karkina, A.B. Notkin, and M.V. Ponomarev, Acta Metall., 40 (1992), 815. [11] SC. Huang and EL. Hall, Metall. Trans. A, 22 (1991), 427. [12] SC. Huang and EL. Hall, Scripta Metall., 25 (1991), 1805. [13] V.K. Vasudevan, M.A. Stucke, S.A. Court, and H.L. Fraser, Phil. Mag. Lett., 59 (1989), 299. 214 215 [14] H. Inui, A. Nakamura, M.H. Oh, and M. Yamaguchi, Phil. Mag. A, 66 (1992), 557. [15] M.A. Morris, Phil. Mag. A, 68 (1993), 237. [16] M.A. Morris, Phil. Mag. A, 68 (1993), 259. [17] M. Hirabayashi and S. Weissmann, Acta Metall., 10 (1962), 25. [18] B. Hanssom and RS. Barnes, Acta Metall., 12 (1965), 315. [19] V. Krasevec, P. Delavignette, and S. Amelinckx, Mater. Res. Bull., 2 (1967), 1029. [20] V.I. Syutkin and ES. Yakovleva, Physica Status Solidi, 21 (1967), 465. [21] D.W. Pashley, J.L. Robertson, and MJ. Stowell, Phil. Mag., A19 (1969), 83. [22] A. Loiseau and A. Lasalmonie, Mater. Sci. Eng., 67 (1984), 163. [23] C.R. Feng, D.J. Michel, and C.R. Crowe, Scripta Metall., 22 (1988), 1481. [24] M.H. Yoo, C.L. Fu, and J.K. Lee, DE89-014779, 1989, [25] M.H. Yoo, C.L. Fu, and J.K. Lee, J. Phys. 111, l (1991), 1065-1084. [26] G.T. Gray III, Microstructure/Property Relationships in Titanium Aluminides and Alloys, ed. by Young-Won (Y -W) Kim and Rodney R. Boyer (TMS Publications, 1991), p263. [27] H. Inui, A. Nakamura, M.H. Oh, and M. Yamaguchi, Phil. Mag. A, 66 (1992), 557. [28] Z. Jin and T.R. Bieler, Scripta Metall., 27 (1992), 1301. [29] Z. Jin and T.R. Bieler, High-Temperature Ordered Intermetallic Alloys V, ed. by I. Baker, R. Darolia, J.D. Whittenberger, and M.H. Yoo (MRS, Pittsburgh, PA, 1993), Vol. 288, p. 775. 216 [30] Z. Jin, R. Beals, and T.R. Bielar, Structural Interrnetallics, ed. by R. Daralia et al. (A Publication of TMS, Warrendale, PA, 1993), p. 275. [31] Z. Jin and T.R. Bieler, Twinning in Advanced Materials, ed. by M.H. Yoo and M. Wuttig (A Publication of TMS, Warrendale, PA, 1994), in press. [32] S. Farenc, A. Coujou, and A. Couret, Phil. Mag. A, 67 (1993), 127-142. [33] S. Wardle, I. Phan, and G. Hug, Phil. Mag. A, 67 (1993), 497-514. [34] Y.Q. Sun, P.M. Hazzledine, and J.W. Christian, Phil. Mag. A, 68 (1993), 471- 494. [35] Y.Q. Sun, P.M. Hazzledine, and J.W. Christian, Phil. Mag. A, 68 (1993), 495- 516. [36] M. Yamaguchi and H. Inui, Structural Interrnetallics, eds. by R. Darolia, J .J . Lewandowski, C.T. Liu, P.L. Martin, D.B. Miracle, and M.V. Nathal (A Publication of TMS, Warrendale, PA, 1993), p127. [37] DA. Wheeler, B. London, and DE. Larsen, Jr., Scripta Metall., 26 (1992), 939. [38] FL. Martin, M.G. Mendiratta, and H.A. Lipsitt, Metall. Trans., 14A (1983), 2170. [39] 1.8. Huang and Y-W. Kim, Scripta Metall., 25 (1991), 1901. [40] M. Yamaguchi, S.R. Nishitani and Y. Shirai, in Proc. on "High Temperature Aluminides & Intermetallics", eds by S.H. Whang, C.T. Liu, D.P. Pope and 1.0. Stiegler, TMS, 1990, p63. [41] R.W. Hayes and B. London, Acta Metall., 40 (1992), 2167. [42] K.I. Ouchi, Y. Iijirna, and KL Hirano, in the Proc. of Titanium ’80 - Science and Technology, ed. H. Kirnura and O. Izumi (The Metallurgical Society of AIME, 1980), 217 559. [43] S.L. Kampe, J.D. Bryant, and L. Christodoulou, Met. Trans., 22A (1991), 447. [44] PL. Martin et al., in the Proc. of Creep And Fracture of Engineering Materials And Structures, ed. B. Wilshire and R.W. Evans (The Institute of Metals, London, 1990), 265. [45] S.C. Cheng, J. Wolfenstine, and OD. Sherby, Met. Trans., 23A (1992), 1509. [46] H. Oikawa, in the Proc. of High Temperature Aluminides & Intermetallics, ed. S.H. Whang, C.T. Liu, D.P. Pope and 1.0. Stiegler, TMS, 1990, p353. [47] H.R. Ogden, D.J. Maykuth, W.L.Finlay and RI. Jaffee, Trans. AIME, 191 (1951), 1150. [48] D. Clark et al., J. Institute of Metals, 91 (1962-1963), 197. [49] LL. Murray, Binary Alloy Phase Diagrams, ed. T.B. Massalski, ASM, Metals Park, OH, 1986, p173. [50] ES. Bumps, H.D. Kessler and M. Hansen, Trans. AIME, 194 (1952), 609. [51] P. Duwez and J .L.Taylor, JOM, January, 1952, p70. [52] S.C. Huang, E.L. Hall and M.F.X. Gigliotti, High Temperature Ordered Intermetallic Alloys 11, ed. N.S. Stoloff, C.C.Koch, C.T. Liu and O. Izurni, Mrs, Pittsburgh, PA, 1987, p481. [53] R.P.Elliott and W. Rostoker, Acta. Met., 2 (1954), 884. [54] B.A. Greenberg, Phys. Status 80]., B55 (1973), 59. [55] T. Kawabata and O. Izurni, Scripta Metall., 21 (1987), 433. [56] S.H. Whang and Y.D. Hahn, Scripta Metall., 24 (1990), 1679. [57] B.A. Greenberg, Phys. Stat. Sol., 42 (1970), 459. 218 [58] B.A. Greenberg and L.I.Yakovenkova, Phys. Stat. 80]., 18A (1973), K129. [59] V.N. Indenbaum, et al., Physics Metals Metallogr., 59 (1985), 52. [60] B.A. Greenberg and Yu.N. Gomostirev, Scripta Metall., 22 (1988), 853. [61] G. Hug. A. Loiseau and P. Veyssiere, Revue Phys. Appl., 23 (1988), 673. [62] C. McCullough, el at., Scripta Metall., 22 (1988), 1131. [63] 1.]. Valencia, el at., Scripta Metall., 21 (1987), 1341. [64] Y.S. Yang and S.K. Wu, Scripta Metall., 24 (1990), 1801. [65] MJ. Blackburn, The Science, Technology and Application of Titanium, ed. R.T. Jaffee and N.E. Promise], Pergamon Press, London, 1970, p633. [66] P. Mukhopadhyay, Metallography, 12 (1979), 119. [67] R.W. Cahn, Advanc. Phys., 3 (1954), 363. [68] B.A. Bilby and A.G. Crocker, Proc. Roy. Soc.A, 288 (1965), 240-255. [69] M. Bevis and A.G. Crocker, Proc. Roy. Soc. A, 313 (1969), 509-529. [70] M. Bevis and A.G. Crocker, Proc. Roy. Soc. A, 304 (1968), 123-134. [71] E. Schmid and W. Boas, Plasticity of Crystals, (London: Hughes, 1950). [72] ED. Hall, Twinning and Diffusionless Transformations in Metals, (London: Butterworths, 1954). [73] H. Kiho, J. Phys. Soc. Japan, 9 (1954), 739. [74] H. Kiho, J. Phys. Soc. Japan, 13 (1958), 269. [75] M.A. Jaswon and DB. Dove, Acta Crystallogr., 9 (1956), 621. [76] M.A. Jaswon and DB. Dove, Acta Crystallogr., 10 (1957), 14. [77] M.A. Jaswon and DB. Dove, Acta Crystallogr., 13 (1960), 232. [78] F. Laves, Naturwissenschaften, 39 (1952), 546. 219 [79] DE. Mikkola and J.B. Cohen, Acta Metall., 14 (1966), 105. [80] A. Guirnier and LL. Strudel, in Proc. 2nd Int. Conf. Strength of Metals and Alloys (ASM,Metals Part, OH, 1970), Vol. 3, p1145. [81] B.H. Kear, J.M. Oblak, and A.F. Giamei, ibid., p1155. [82] SB. Chakraborty and B.A. Starke, Jr., Acta Metall., 23 (1975), 63. [83] GP. Bolling and RH. Richman, Acta Metall., 13 (1965), 709. [84] J.P. Kubin and J.V. Rieu, Scripta Metall., 12 (1978), 927. [85] M.L. Green and M. Cohen, Acta Metall., 27 (1979), 1523. [86] E. Goo, T. Duerig, K. Melton, and R. Sinclair, Acta Metall., 33 (1985), 1725. [87] R.W. Cahn and J.A. Coll, Acta Metall., 9 (1961), 138. [88] D. Shechtman and LA. Jacobson, Metall. Trans. A, 6 (1975), 1325. [89] G. Vanderschaeve and T. Sarrazin, Phys. Status Solidi, 43a (1977), 459. [90] M. Yamaguchi, Y. Umakoshi, and T. Yamane, Phil. Mag., 55 (1987), 301. [91] A. Faress and G. Vanderschaeve, Acta Metall., 35 (1987), 691. [92] G. Vanderschaeve and B. Escaig, Phys. Status Solidi, 20a (1973), 309. [93] J.W. Christian and DE. Laughlin, Acta Metall., 36 (1988), 1617. [94] M.H. Yoo, J. Mater. Res., 4 (1989), 50. [95] RC. Rowlands, E.O. Fearon and M. Bevis, Trans. Metall. Soc. A.I.M.E., 242 (1968), 1559. [96] M. Bevis, P.C. Rowlands, and F. Acton, Trans. Metall. Soc. A.I.M.E., 242 (1968), 1555. [97] M.H. Yoo and ET. Loh, Fundamental Aspects of Dislocation Theory, ed. by J .A. Simmons, R.deWit, and R. Bullough (US-NI Spec. Publ. 317, 1970), Vol. I, p479. 220 [98] A.H. Cottrell and B.A. Bilby, Phil. Mag., 42 (1951), 573. [99] N. Thompson and DJ. Millard, Phil. Mag., 43 (1952), 422. [100] J.A. Venables, Phil. Mag., 6 (1961), 379. [101] J.T. Fourie, F. Weinberg, and F.W.C. Boswell, Acta Metall., 8 (1960), 851. [102] PB. Price, Proc. Roy. Soc. (a) (London), 260 (1961), 251. [103] E. Orowan, Dislocation in Metals, A.I.M.E., New York, 1954, p116. [104] R.L. Bell and R.W. Cahn, Proc. Roy. Soc. (a) (London), 239 (1957), 494. [105] R. Bullough, Proc. Roy. Soc. (a), 241 (1957), 568. [106] TH. Blewitt, R.R. Coltman and J.K. redman, J. Appl. Phys., 28 (1957), 651. [107] D.G. Westlake, Deformation Twinning, ed. by RE. Reed-Hill, J.P. Hirth, and H.C. Rogers (Gordon and Breach Science Publishers, 1964), p29. [108] H.S. Rosenbaum, ibid., p43. [109] RE. Reed-Hill, ibid., p295. [110] M.H. Yoo, Metall. Trans. A, 12A (1981), 409. [111] D.G. Westlake, Acta Metall., 9 (1961), 327. [112] K. Yasutake, S. Shimizu, and H. Kawabe, J. Appl. Phys., 61 (1987), 947. [113] J.K. Lee and M.H. Yoo, Metall. Trans. A, 21A (1990), 2521. [114] RA. Lebensohn and C.N. Tome, Phil. Mag. A, 67 (1993), 187. [115] RE. Cooper, Acta Metall., 13 (1965), 46. [116] RE. Cooper, Acta Metall., 14(1966), 78. [117] M.J. Marcinkowski and KS. Sree Harsha, Trans. Metall. Sco. AIME, 242 (1968), 1405. [118] S. Takeuchi and AS Argon, J. Mater. Sci., 11 (1976), 1542. 221 [119] A.H. Clauer, B.A. Wilcox and J.P. Hirth, Acta Metall., 18 (1970),381. [120] D. McLean, J. Inst. Metals, 80 (1951-52), 507. [121] JD. Parker and B. Wilshire, Phil. Mag., 41A (1980), 665. [122] M. Henderson and KP. Hale, Proc. 3rd Int. Conf. on High Voltage Electron Microscopy, eds by PR. Swann, C.J. Humphreys and MJ. Goringe, Academic Press, New York, 1974, p206. [123] T.G. Langdon and RB. Vastava, Mechanical Testing for Deformation Model Development, eds. by R.W. Rohde and J .C. Swearengen, ASTM, 1982, p435. [124] D. Hanson and M.A. Wheeler, J. Inst. Metals, 45 (1931-32), 229. [125] W.A. Wood and GR. Wilms, J. Inst. Metals, 75 (1948-49), 693. [126] C.R. Barrett, J.L. Lytton and OD. Sherby, Trans. AIME, 239 (1967), 170. [127] J.B. Bird, A.K. Makherjee, and J.F. Dorn, Proc. Int. Conf. on Quantitative Relation Between Properties and Microstructure, eds. by D.G. Brandon and A Rosen, Israel University Press, Jerusalem, 1969, p225. [128] F.R.N. Nabarro, Conference Proceedings on ”Strenght of Solids", The Physical Scoiety, London, 1948, p78. [129] C. Herring, J. Appl. Phys., 21 (1950), 437. [130] R.L. Coble, J. Appl. Phys., 34 (1963), 1679. [131] J. Harper and J.B. Dom, Acta Metall., 5 (1957), 654. [132] OD. Sherby and RM. Burke, Prog. Mater. Sci., 13 (1967), 325. [133] J. Weertman, Trans. AIME, 218 (1960), 207. [134] J. Weertman, Acta Metall., 25 (1977), 1393. [135] D. McLean, Rep. Prog. Phys., 29 (1966), l. 222 [136] R. Lagneborg, Met. Sci. J., 3 (1969), 161. [137] NE Mott, Proc. Conf. on Creep and Fracture of Metals at High Temperatures, H.M.S.O., London, 1956, p21. [138] PB. Hirsch and D. Warrington, Phil. Mag., 6 (1961), 735. [139] C.R. Barrett and W.D. Nix, Acta Metall., 13 (1965), 1247. [140] M. Yamaguchi and H. Inui, Structrual Intermetallics, ed. by R. Daralia et al. (A Publication of TMS, Warrendale, PA, 1993), p127—142. [141] M. Yamaguchi, and Y. Umakoshi, Prog. in Mater. Sci., 34 (1990), l. [142] J .P. Hirth, and J. Lothe, Theory of Dislocations, Second Edition, John Wiley & Sons, Inc., 1982. [143] M.J. Darkbum, The Science, Technology and Application of Titanium, edited by RT. Jaffee and N.E. Promise] (Pergamon Press, London, 1970), p633. [144] H.E. Deve, A.G. Evans, and D.S. Shih, Acta Metall., 40 (1992), 1259. [145] TE. Mitchell and J.P. Hirth, Acta Metall., 39 (1991), 1711-1717. [146] RE. Peierls, Proc. Phys. Soc., 52 (1940), 23. [147] F.R.N. Nabarro, Proc. Phys. Soc., 59 (1947), 256. [148] G. Fantozzi, C. Esnouf, W. Benoit, and LG. Ritchie, Prog. Mater. Sci., 27 (1982), 311. [149] R. Bullough and V.K. Tewary, Dislocations in Solids, ed. by F.R.N. Nabarro (North-Holland Publishing Company, 1979), Vol. 2, pl. [150] I.L.F. Ray and D.J.H. Cockayne, Proc. Roy. Soc. London A, 325 (1971), 543. [151] C. Weigel, H. Alexander, and J.W. Corbett, Phys. Status Solidi. B, 71 (1975), 701. 223 [152] C.L. Fu and M.H. Yoo, Phil. Mag. Letter, 62 (1990), 159-165. [153] C.L. Fu and M.H. Yoo, Phil. Mag. Letter, 58 (1988), 199. [154] M.H. Yoo, C.L. Fu, and J.K. Lee, Twinning in Advanced Materials, ed. by M.H. Yoo, M. Wuttig (TMS Publication, 1994), in press. [155] T. Fujiwara, A. Nakamura, M. Hosomi, S.R. Nishitani, Y. Shirai, and M. Yamaguchi, Phil. Mag. A, 61 (1990), 591. [156] S.C. Cheng, J. Wolfenstine, and OD. Sherby, Met. Trans., 23A (1992), 1509. [157] H. Oikawa, in the Proc. of High Temperature Aluminides & Intermetallics, ed. S.H. Whang, C.T. Liu, D.P. Pope and 1.0. Stiegler, TMS, 1990, p353. [158] M. Yamaguchi, Metals and Technology, 60 (1990), 34. [159] Y. Yarnabe, K. Sugawara and M. Kikuchi, Proceedings - 1990 Tokyo Meeting, Japan Inst. of Metals, Sendai, the Japan Inst. of Metals, 1988, p409. [160] V.Yu. Gertsman, et a], Scripta Metal]. 24 (1990), 1027. [161] W.A. Kaysser, in the Proc. of Second International Conference on Hot Isostatic Pressing: Theory and Applications (Gaithersburg, MD, USA, June 7-9, 1989), pl. [162] K. Matsuki et al., 2. Metallkde, 79 (1988), 231. "1111111111111111“