I OF 2 ORNLP 1722 met - 1. v '. } Alli 3 6 ES 140 AWO 1.1.4 1.6 MICROCOPY RESOLUTION TEST CHART NATIONAL BUREAU OF STANDARDS - 1963 oroord-1922 Conf. 65/031-2 NOV 18 BSS ORNI - AIC - OFFICIAL RESLEASID FOR ANNOUNCEMENT I NUCLEAR SCIENCE ABSTRACTS 1L - AEC - OFFICIAL GRAIN BOUNDARY MIGRATION by privately owned rights; or with the Commission, or his employment wits such contractor. dissemiantes, or provides accon to, kay In:ormation pursuant to his employment or contract such employee or contractor of the Commission, or employee of such contractor prepares. ployee or contractor of the Commission, or employee of such contractor, to the extent that As usod in the above, "person acting on boball of the Commission" includes any on- use of any information, apparitus, method, or process disclosed in this report. B. Assumos any ilabilities with respoct to the use of, or for damages resulting from the of any information, apparatus, method, or process disclosed in this report may not infringe racy, completeness, or urefulness of the information contained in this report, or that the we A, Makos any warranty or representatioa, expressed or implod, with respect to the accu- States, por the Commission, nor any person acting on behalf of the Commission: This report was preparad as an account of Goverament sponsored work. Neither the United LEGAL NOTICE Most metallic materials in their ultimate use form are polycrystalline. The presence of the crystallite - or grain - boundaries has profound ef- fects on the properties of the materials, and boundary migration may alter these properties significantly. In attempts to understand these effects and the underlying nature of grain boundaries, boundary migration studies have played an increasingly important part. It is the purpose of this paper to delineate the main results of these migration studies. This will be done in two parts; in Part I, the effect of orientation and boundary structure factors and their interdependence with impurity effects will be considered. In Part II, investigations aimed at the quantitative understanding of the drastic effect dissolved impurities have on boundary migration will be examined; the results of these investigations will be compared with the predictions of migration rate theories designed to account quantifatively for the impurity effect. In view of the facts that there have been a number of excellent re- . views on various aspects of grain aggregates and grain boundaries (1-8) and that there are several new,ones in the present volume, no attempt will be made to cover the literature exhaustively, but rather only those works considered to have the most direct bearing on the subjects under discussion will be referenced. . . . . 7 . . --- ORNL-AEC-OFFICIAL * Professor, Department of Metallurgical Engineering, Illinois Institute of Technologyä Chicago, Illinois. # Metallurgist at the Oak Ridge National Laboratory, Oak Ridge, Tennessee, which is operated by Union Carbide Coiporation for the U. S. Atomic Energy Commission. ORNI - AEC - OFFICIAL Part I. Orientation and Structure Low-angle and high-angle boundaries. A grain boundary may be defined as a layer of distorted material created by the atomic mismatch be- tween two differently oriented, adjacent crystals of the same phase. It is customary to categorize grain boundaries as either low-angle or high-angle boundaries. The dividing line is usually somewhat indefinitely placed at a relative grain mismatch angle below which the boundary is in many respects equivalent to a wall of discreet dislocations and above which the disloca- tions are so close together as to lose their individual identities. This generally means at mismatch angles of a few degrees. Thus, it has been demonstrated (9,10) that low-angle - or sub-grain — boundaries can be caused to migrate, presumably by dislocation glide, under the impetus of uniaxial stress, reversing migration direction when the direction of the stress is reversed, Figure 1. The migration rates of such boundaries are found to be dependent nd only on the stress level but also on the temperature and the angle of misfit between the bounding grains. In particular, the rates in- crease with temperature and decrease with increasing misfit angle (11). The latter effect, shown in Figure 2, is attributed to the mutual interference of the dislocations when they are too close together. In contrast to the relative unanimity of opinion as to the basic nature of low-angle boundaries, there has been no single predominating concept concerning the structure of high-angle boundaries; it is with this more con- troversial field of high-angle boundaries that the remainder of this paper is ". . .. -- . . . concerned, . . .: -.: . : .. :'. . . . .. .... . : ci Little direct evidence has as yet been obtained as to the detailed nature of high-angle grain boundary structures. This is because such evidence requires "seeing" individual atoms, and even the powerful electron micro- scope has not been able to do that. The recent invention by E. W. Muller of the field-ion microscope (12), however, with its capability of resolutions down to 2-3 A along with readily obtained magnifications of 100 or more, has now brought individual atoms within view. Investigations on grain boundaries with the instrument are still fragmentary, but it is not unduly optimistic to expect that the field-ion microscope will in the next few years do for grain boundary structure ( and for other atomic-scale phenomena as well) what the electron microscope has done for the understanding of dis- locations in the past decade or so. An example of a grain boundary in tungsten taken from the field-ion microscopy work of Brenner (13) is given in Figure 3. It should be noted that each bright spot in this photograph is the image of a single atom at a magnification of about 1,600,000 X! Grain boundary models. Present day views of grain boundary structure date from abc ut 1929 when Hargreaves and Hills (14) first visualized the boundary as a region one to two atoms wide in which the atoms of the adjoin- ing grains assume compromise positions between the two lattices. The models which have since been proposed in attempts to describe in some detail the references 1, 5, 6 and 7). Those which have been given the most attention are: (a) Mott's (15) "island" model. The boundary is considered to be ..----.. . ..... .. ---. made up of islands of good fit between the adjoining lattices surrounded by regions of bad fit. Migration is viewed as a process involving the disordering, or "melting," of a group of .n atoms from the matrix grain into the bad fit boundary regions and the "crystallization" of these atoms onto the lattice of the growing grain. -na (b) Ke's (16) viscous boundary model. The boundary is looked upon as a viscous region of disorder which can absorb disordered groups of atoms - presumably resulting from relaxed vacancies -- from a grain, thereby lowering the boundary viscosity and facilitating its migration. Im- purity atoms are suggested to have an effect opposi te to that of vacancies. (c) The dislocation model. The boundary is considered to be made up of an array of closely spaced dislocations similar in many respects to low- angle boundaries; the mod. has been proposed in various forms by many authors, e.g., references 17, 18, 19 and 20. As pointed out by Lucke and Stuwe (6), it has been difficult to make choices between these models. In essence, the reason for this is that both the theories and the available experimental evidence are still too limited to give the kind of comprehensive description of grain boundaries necessary to ex- plain all their major attributes. With respect to grain boundary migration, these models have had some modest success. It is reasonably well establish- ed that grain boundary diffusion rates are higher for boundaries between grains with large misorientations (with the exception of twin orientations) than for those where the misorientations are less than 15-20° (see, for example, references 2 and 3). On the assumption that boundary migration is controlled by boundary diffusion the above models then predict (21) that boundary n een kanto.am 44% binn . . .' . 1 .. . migration should be more rapid for high-angle boundaries with large mis- orientations than for those with smaller misorientations. There is much in- direct evidence that this is true. However, recent measurements of the migration rates of single boundaries have produced direct evidence that the orientation dependence of high-angle boundary migration is a more specif- ically selective effect -- i.e., that within the category of high-angle boundaries with large misorientations there are certain special orientations which under some conditions have much higher mobilities than others. The three models for boundary structure listed above cannot account for such specific selectivity. These problems are discussed in detail in the next two sections. Orientation dependence of migration. It has been known for many years (7,8) that when deformed polycrystalline materials are annealed, marked preferred orientation textures may be developed. These textures may be similar to the deformation textures from which they form or may be quite different. On further annealing, secondary recrystallization may take place, and again marked, but different, textures can develop. It has been argued (22,23), on the one hand, that oriented nucleation - i.e., the predominant presence of growth nuclei in only certain orientation relations with the matrix — plays a significant role in forming the annealing textures; on the other hand, it has been maintained (21) that oriented growth -- high growth rates associated with nuclei of certain orientations.: elative to the matrix alone is sufficient to account for the annealing textures. This controversy is, perhaps, not yet unambiguously resolved; be that as it may, however, during the past decade studies of the growth of new grains in single crystals have demonstrated quite convincingly that the orientation dependence of growth is both real and marked. For example, Kohara, Parthasarathi and Beck (24) found that in a single crystal of 99.994% purity aluminum cold roll- ed to 80% reduction in area, rubbing with a fine abrasive on one side of the sample produced many fine, randomly oriented grains on this side after short subsequent annealing, but that on further anneciing only those grains having orientation relationships with the matrix corresponding to approximately 40° 'rotations around a common (111) pole survived to grow completely across the rolled crystal. A number of other studies have been made on single crystal samples provided with a growth-supporting driving force either by deformation or by a striation (or mosaic) structure resulting from solidification. Usually the driving forces and annealing temperatures were selected in such a way that spontaneous nucleation of new grains throughout the crystal was avoided. Heavy local deformation introduced at one end of the crystal would then cause new grains to nucleate "artificially" in this local region during annealing. Most often, one or two of the grains so nucleared in a given crystal were found to crowd out the others early in the competition to consume the matrix crystal. One, or both, of two types of observation were then made: (1) As in the work of Kohara, Parthasarathi and Beck (24), the orientation relation- ships between the dominant new grains and the inatrix were determined and expressed by giving for each dominant grain the rotation angle and the crystal- lographic direction common to both new grain and matrix around which the minimum rotation was necessary for the lattices of the two to coincide; (2) The migration rates of the boundaries between the dominant grains and the matrix were measured. Most frequently only type (1) measurements were made (as examples, see references 24-33, 38). Both orientation and migration rate measurements (i.e., types (1) and (2)) have been made only by Graham and Cahn (34), by Aust and Rutter (35-37) and by Rath and Gordon (39). In some cases (e.g., 31 and 40) orientation relation measurements were made on new grains formed by spontaneous nucleation, that is, without the help of localized heavy deformation. Finally, in two instances, (Graham and Cahn (34) and Aust and Rutter (41)) migration rate experiments were carried out on boundaries in bicrystals prepared with predetermined misorientations. The single crystal measurements in which type (1) — orientation -- measurements were made established quite clearly that there is an ubiquitous and strong tendency for grains with certain relative orientations with respect to the matrix crystal to be preferred during annealing just as there is for polycrystalline material. Further, many of these investigations (e.g., Kohara, Parthasarathi and Beck (24), Liebmann, Lucke and Masing (25), Yoshida, Liebmann and Lucke (26), Lucke and Ibe (33) and Aust and Rutter (29)) succeeded in proving that growth selection alone can produce these preferred orientations*. For face-centered cubic crystals, 30° and 40° rotations about (111), and for body-centered cubic crystals 25-50° rotations around (110), have been most frequently, but not exclusively, found. An example from the work of Yoshida, Liebmann and Lucke (26) is given in Figure 4. It is im- portant to note that findings such as these do not mean that differences in mobilities between various boundaries are necessarily responsible for growth * Note that these experiments have not proved the general ab oriented nucleation effect. Direct evidence on the latter, however, is difficult to obtain if for no other reason than that it is difficult to define the end of a nucleation event and the beginning of nucleus growth. selection in single crystals or preferred orientation in general. As has been demonstrated by Aust and Rutter (27-29) and by Lucke and I kve (33), twinning during growth can result in the replacement of a migrating boundary by a second boundary and a coherent twin boundary whenever thereby the total interfacial energy is lowered -i.e., whenever the second boundary has a specific interfacial energy lower than that of the first by an amount larger than the specific interfacial energy of the coherent twin boundary. The interfacial energy of grain boundaries varies with orientation factors and, therefore, giains oriented so that they have low energy boundaries may well Le favored by the interfacial energy factor during the nucleation or the growth stage of any development of new grains. Thus, the observation of growth selectivity based only upon orientation measurements means strictly only that there is a strong orientation dependence of either -- or both — boundary mobility or energy. It seems probable, however, that whenever there is a strong mobility dependence this will be the dominating influence in growth selection. This is because a new boundary produced to lower interfacial energy would not be likely to survive unless its mobility is among the highest. These orientation relation measurements on dominant grains growing in single crystal matrices have been made largely on impure metals or on alloys. In a few instances, zone-refined metals were used, however, and these have made it evident that very high purity materials may behave dif- ferently than less pure material.* In particular, it has been found that in Here high purity is taken to mean carefully zone-refirad, for it has be- come clear that dissolved impurities in amounts as small as 0.1 to 1 atomic ppm can have very large effects on boundary migration behavior (for ex- amples, see references 29, 30, 35-39 and 42-47). zone-refined lead (35-37) and zone-refined copper (30) the preferred orienta- tions do not appear, at least at high temperatures. Some of the Aust-Rutter results on zone-refined lead are reproduced in Figure 5 (35). It is apparent from this figure that the rotation axes which relate the new grains to the matrix crystals are grouped around the (771) and (001) poles in lead containing 0.0005 to 0.004 wgt % tin (Figure 5b), but that in the purer material (Figure 5a) no such grouping takes place. Aust and Rutter have described the arrange- ment of rotation axes in Figure 5a as random; however, careful reexamination of the figure engenders the impression that the rotation axes in Figure 5a are random except that they have avoided the regions around which rotation axes have clustered in Figure 5b. This impression is strengthened by a study of all the Aust-Rutter data for materials in which they did not find preferred group- ings such as those in Figure 5b. A composite plot of this data is shown in Figure 6; included are not only the data on zone-refined lead (references 29 and 35) but also data from the alloys of this lead containing silver or gold in amounts up to 0.83 at. ppm (the entire range studied )(37). The arcs around the principle poles of the stereogram in Figure 6 enclose regions each of which should contain approximately 5% of the rotation oxes if these are distributed at random (48). Since the total number of rotation axes shown is 62, each region should contain 3-4 axes but in fact contains only one (1-1/2 near the (110) pole), and none of these is closer than about 4-5° to the principle pole of concern. It may well be, then, that under certain conditions in par- ticular, extremely high purity or with the presence of specific impurities - a reverse kind of preferred orientation takes place in which certain orientations are avoided. Grahams and Cahn (34) noted a similar phenomenon in quite impure aluminum (99.6% pure — principle impurities 0.19% Fe and 0.12% Si). This point needs some further experimental attention. In view of the indicated importance of impurity effects in the develop- ment of selective growth, a number of investigations based on zone-refined materials have been carried out in which actual migration rate measurements were made on single boundaries for which the misorientations of the bounding grains were also measured. In each of these the investigators found definite evidence that boundary mobility is orientation dependent.* The extensive studies of Aust and Rutter (35-37, 41) huve dominated this area; they studied boundary migration in striated crystals of zone-refined lead and very dilute alloys of tin, silver and gold in this lead. Again, localized deformation at one end of the striated crystals was used to provide, upon subsequent annealing, many striation-free nuclei for growth into the striated crystals. Aust and Rutter, found, as indicated above, that the dominant new grains had nearly random orientation relationships with the matrix when grown in gold and silver alloys (up to 0.83 at. ppm) and in unalloyed zone-refined lead at high tem- peratures (sele, for example, Figure 5a); the boundaries between new grains and matrix were in this case designated "random" boundaries. In alloys containing 9-80 at. ppm tin and in the unalloyed zone-refined lead at low temperatures (e.g., Figure 5b), however, the dominant grains had certain unique relationships -- rotations about low-index axes such as (111). (110) and (100) -- with that It should be noted that occasional observations have been made/the boundary mobility may depend on the orientation of the boundary itself. For example, it was pointed out by Kohara, Parthasarathi and Beck (24) that for the same relative grain orientations, tilt boundaries appeared to be more mobile than twist boundaries. This dependence on boundary orientation has been given relatively little attention, however; most often it has been taken into account only by making efforts to hold the effect constant. the matrix crystals; the associated boundaries in this case were called "special" boundaries. These unique relationships were not found for tin con- tents above 80 at. ppm. Measurements of the migration rates of the boundaries revealed that the rates of special and random boundaries were approximately the same at high temperatures, but those of random boundaries were signifi- cantly the lower at low temperatures. Increasing tin content markedly lowered the migration rates and raised the apparent activation energies of random boundaries (as did Ag and Au) but had only minor effects on the migration rates, and none on the activation energies, of special boundaries. The latter activation energies were the lowest of all the boundaries studied. These effects are illustrated in Figures 7 (35) and 8 (36). Aust and Rutter also found (41) that for tilt boundaries produced at predetermined misorientations in bicrystals of zone-refined lead, those between crystals having the special orientations found before showed higher migration rates and lower activation energies than other boundaries. Similar, but more limited, results have been obtained in very dilute alloys of copp er in zone-refined aluminum by Frois and Dimitrov (38) and by Rath and Gordon (39). An explanation of these results in terms of the differences in impurity segregation at boundaries of different structures has been advocated by Aust and Rutter. They concluded, however, that none of the models for boundary W very specific orientation selectivity of growth mobility found experimentally. Instead, they based their interpretation on a concept of boundary structure first advanced by Kronberg and Wilson (49) in a study of secondary recrystallization in copper Table 1. Kronberg-Wilson Relationships (from Ref. 50) Axio or rotation Reciprocal of Least. Angle of rotation A Α Σ Φ Α Σ Φ Α Σ ω - 100 - - - - - 180° 120.7° 10.1° - - - 100 36.0° 22.09 28.1° 100 210 3 210 6 210 , 7 210 210 16 211 .. 3 211 6 211 7 211 10 211 10 181.8° 180° 73.40 96.49, : 48.26 180ο. 101.00 136.6° 630 78.7° 70.6° 38.0°. 50.5o 86.09 20.5o ------ 310 310 310 310 310 311 311 311 311 311 180 116.4° 141.0° . 76.7• 03 . 146.40 164.2° 67.1 7 11 136 10o 3 6 9 11 18 321 421 321 322 322 922 41ο 410 410 18 ο 13 176 Φ. 136 17% io 152.7 107.0° 180». 152.70 107.0° 180 10 10 . 10 311 180° 50.7° 117.00 411 Ο 180° 11 600 11 17A 106 120.0" 14° 111 153.50 11 38.2° 27.8° 46.8° - 221 221 221 221 221 6 9 135 176 143.1° 90 180° 112.6° 61.0° Οι 320 320 320 320 320 411 140° 411 411 100.6° - 180° 331 121.0° 331 1.80 381 331 7 11 13o 176 105 5 06.7° 881 10 17» 196 density of common points 13η 17η 17» 10η 13ο ΔΩ!» 110.00 82.1 63.00 180° Kronberg and Wilson noted that the textures obtained in their samples - rotations about common (111) and (100) axes -- were such that the lattices of the new and old grains had a common sublattice of a special kind. The type of sublattice referred to — labeled a coincidence lattice — consists in general of an array of points which lie on the lattices of both grains (see Figure 9a) and has the same symmetry as, but greater unit cell than, the basic lattices. Coincidence lattices can be found for any orientation relationship between two adjacent grains, but they have special significance only when, as in the Kronberg-Wilson case, the density of coinciding sites is relatively high. The orientation relationships leading to these special coincidence lattices are now frequeritly referred to as Kronberg-Wilson (K-W) orientations. A listing of some of these, where the coincidence-site density is 1-in-19 or greater, is given in Table ! (50). The boundaries between grains of K-W relationship are boundaries of relatively good fit, since some of the coinci- dence sites lie in the boundary; the degree of good fit increases with the density of coincidence sites. According to Kronberg and Wilson, the presence of coincidence atoms may endow boundaries with high mobility and thus account for the development of the particular textures they found. This view can be applied to much of the preferred orientation work in both polycrystalline and single crystal material, for the dominant grains observed usually can be shown to have orientation relationships with the matrix which approximate K-W rela- tionships. For example, the 30° and 40° (111) rotations frequently observed inf.c.c. materials can be identified with the 27.80 and 38.20 (111) K-W rotations shown in Table I, and the 25-30° (110) relation often found for b.c.c. materials with the 26.50 (110) K-W relation in Table 1. The Aust and Rutter "special" orientations were largely rotations of approximately 36-42° and 23° about (111) and 230, 28° and 37° about (100); the Frois and Dimitrov preferred grains were approximately 40° about (111); and those of Rath and Gordon were 35-37º about (100) and 39-44º about (111). All of these may be seen from Table 1 to be roughly K-W relationships. Aust and Rutter argued that K-W -- or special boundaries are good-fit boundaries and, therefore, absorb considerably less impurity than ! bad-fit--- random — boundaries. The tendency to absorb impurities is also affected by temperature; at low temperatures impurities segregate at boundaries strongly, but at high temperatures the impurity atmosphere "evaporates." Thus, at high temperatures there was little difference between the migration rates of :: their special and random boundaries since neither absorbed significant amounts of impurity. At low temperatures, however, the random boundaries become impurity laden and the migration rates therefore fell well below those of the relatively impurity-free special boundaries. The absence of special boundaries in the Ag and Au alloys (and in the alloys with Sn above 80 ppm) was difficult to explain. Evidence was found, however, that both Ag and Au had an especially strong tendency to segregate to boundaries in lead; it was, there- fore, suggested that the absence of special boundaries in these cases might mean that the segregation tendency was strong enough so that even good-fit boundaries were highly impurity-laden. This interpretation is undoubtedly a reasonable one. It seems appro- priate at this point, however, to extend it somewhat and to make a modifica- tion in the viewpoint. The modification is indicated by a reinterpretation of some of the experimental evidence and by a conceptual difficulty. Kronberg and Wilson (49) originally suggested that high-density co- - .. . 14 incidence-site boundaries - i.e., good fit boundaries — are intrinsically high mobility boundaries. They reasoned that this may be so because the atom movements necessary for migration of a coincidence type boundary may be only fractions of an atomic distance – at least near the coincidence atoms – and may also be partly cooperative in nature, almost martensite-like with- out the macroscopic distortion, thus, perhaps, having low activation energies. This view has also been expressed by others (e.g., ref. 7). In this connection, however, it should be noted that though some of the atom movements for the migration of a coincidence boundary may be cooperative in a given atom plane, the sense of the ring-like movements in alternate planes would be in opposing directions (49); this could conceivably reduce the effectiveness of any co- operative component in the atom movements. More important, however, the concept of lower activation energies in relatively "tight" boundaries is dif- ficult to understand — both intuitively and because it appears to be contrary to the well established fact of lower activation energies for diffusion in grain boundaries as compared with grain bulk; the latter is presumably due to the "looseness", or porosity, of the grain boundary structure as compared to the "tight" bulk structure. It is not clear whether Aust and Rutter explicitly advocated the Kronberg-Wilson idea that good-fit boundaries are intrinsically high mobility boundaries; their interpretation, however, certainly appears to imply that such boundaries unimpeded by impurity atmospheres have mobilities which are at least as high as - if not higher than those of more random boundaries. It seems more natural: to adopt the view advanced by Li (20, 51) that boundaries between grains having ideal high coincidence-site density relations are in . fact low mobility boundaries because their porosity is low and that only when the grains deviate appreciably from the ideal K-W relations is the boundary mobility high. This view is supported by the fact that essentially all of the preferred growth relationships which have been observed are at best only approximately K-W relationships. The rotation axes and rotation amounts found are usually several degrees from the ideal values. The very thorough statistical investigation by Lucke and Ibe (33) established this lack of ideality quite clearly. Whereas previous studies were each limited to relatively small numbers of samples, Lucke and I be examined the orientation relationships be- ; tween some: 1200 recrystallized grains and the deformed single crystal matrices in aluminum (99.998%) and in alloys of this aluminum with iron, silicon and manganese in the range of about .2 to .5 wgt %. They found that although certain boundaries, roughly of the K-W type, were indeed favored in the growth process, there was a considerable scatter from ideal K-W relations. The rota- tion axes of the dominant grains deviated by 12° on the average from the ideal coincidence-site positions. As a matter of fact, Lucke and Ibe, along with Lucke and Stuwe (6), questioned whether with such large deviations the K-W concept could still have physical significance. It is believed, however, that this objection is largely obviated by a model put forward by Brandon, Ralph, Ranganathan and Wald (50). They proposed that deviations of several de- grees in coincidence relationship could be accommodated by supposing disloca- tions to be present in the coincidence sublattice along the boundary plane. This is illustrated schematically for a two-dimensional, simple square lattice in Figure 96. It may be seen in the figure that the introduction of disloca- tions not only rotates the two lattices with respect to one another, but also . .. brives : widens the boundary (to perhaps 3 or 4 atom diameters) and introduces consider- able porosity. Some experimental corroboration for the idea that boundaries may still exhibit appreciable atomic fit even when they deviate several degrees from ideal K-W relations has recently been found with the field-ion micro- scope (13, 52). These same studies have also indicated that near-ideal K-W boundaries have widths of no more than 1-2 atom diameters, whereas more random high-angle boundaries — which still exhibit some good-fit regions - have widths of 3 or 4 atom diameters. The idea that good-fit boundaries are low mobility boundaries derives further experimental support from the fact that, the most frequently observed high mobility bound-ry types — 30° and 40° about (111) and 25-30° about (110) - are not the highest coincidence-site density boundaries (see Table 1). In addition, a coherent twin boundary may be considered to be a K-W boundary of the highest coincidence-site density; yet, coherent twin boundaries are known to be virtually immobile (see, for example, Graham and Cahn (34) and Rathenau (53)). Hypotheses for orientation-impurity effect on boundary mobility. In the light of these considerations it seems worthwhile to consider modification of the Aust-Rutter view on the orientation dependence of grain boundary migration. This modification will be based on the following postulates: (1) The mobility, M, of a grain boundary in a pure material depends on the porosity — or vacancy concentration in the boundary; the mobility is the higher the greater is the porosity because the energy barriers to atomic diffusion are low when the porosity is high. (2) The mobility in an impure material depends also on the segrega- tion of solute impurities at the boundary, the greater the segregation the 17 lower the mobility tends to be. (3) The boundary mobility in general may, thus, be represented by the equation M(p,c) = fy(p) x fg(1/cp) where f, is an appropriate function of the boundary porosity, p, in the pure material, and fy of the solute concentration at the boundary, Cg. Both f, and fy may be temperature dependent. (4) The porosity, and thus, the porosity effect on mobility, for varies with the density of coincidence 'sites in the boundary and, consequently, with the type of coincidence-site relationship, the perfection with which this relationship is attained, and the orientation of the boundary with respect to the planes of closest coincidence-site packing in a given coincidence lattice. The latter two factors can be represented by an angular increment, 40, such that 40 and porosity increase as the departure from ideal orientation, both relative grain and boundary, increases. (5) Solute segregation at a boundary follows the principles enunciated by Cahn (54) and others (6, 42, 43, 55, 56) and discussed in detail later in this paper*. Therefore, whether the solute tends to adsorb or to desorb at a stationary boundary, for a moving boundary there will always be a solute build-up at or near the boundary which can be represented by some effective average boundary concentration, Cg: (6) Because low porosity boundary structures are more like the bulk grain structures than are high porosity boundary structures, for a given bulk * See page 48 * 18 concentration c, the value of CR will be the higher the greater the boundary porosity* . On the basis of these six postulates, the curves in Figure 10 can be constructed. Here, for convenience in plotting, the normalized functions fo = fy(p) /f,(p)., f. = fg(1/cp)/f2(1/cp. and M/M, rather than the abso- lute functions, are plotted against the incremental misfit angle 40. The subscript o in each case indicates the value of the function in a completely pure, randomt boundary. The curves can be understood most readily by considering an absolutely pure material first. In this case the value of the impurity function is by definition unity, the mobilities of various boundaries are determined by the porosity function alone, and, thus, the curves in Figure 100 essentially give these mobilities. For a high-density coincidence-site boundary with ideal relative grain orientation and boundary position, i.e., with 40 = 0), the porosity function f, has the low value given at point c in Figure 10a, and the mobility is very low. This would correspond to, say, a coherent twin boundary or, perhaps, a K-W boundary with a l-in-5 coincidence-site density. For some other K-W boundary, such as the commonly observed 1-in-7 boundary represented by a 38° rotation around (111) the porosity function and the mobility are higher, as indicated by the point b in Figure 10a. As the porosity Electronic effects, as well as porosity, determine the amount of boundary segregation, but for a given impurity these effects are undoubtedly secondary to boundary porosity: Random with respect to a boundary is taken in the Aust-Rutter sense to mean a boundary with the lowest percent of good-fit regions possible. in any boundary increases due to increase in 40, all boundaries increase their mobilities as shown by curves A, B and C in Figure 10a. The shapes of these curves have been approximated by assuming a sinesoidal variation of porosity with 40 as derived by Li (20) for porosity versus misfit angle in tilt boundaries. These curves indicate, thus, that for very pure materials random or near-random boundaries, having the highest mobilities, should be preferred rather than special boundaries of the K-W type; further, they indicate that boundaries of the latter type should actually be avoided. Both of these con- ditions have been observed as discussed in connection with Figures 5 and 6*. In contrast, on the basis of the K-W concept of good-fit boundaries as high mobilities boundaries, there appears to be no reason why such boundaries should not still be preferred in very high purity materials. The prediction of preference for non-K-W boundaries in high purity materials appears at first glance to be in contradiction to the results of Aust and Rutter for tilt boundaries in bicrystals of zone-refined lead (41). There is, however, an alternative interpretation of these results to that given by Aust and Rutter. The migration rates at 200° and 300°C and the temperature de- pendence parameters, Q, for these boundaries are reproduced in Figures 11 and 12. The Q's particularly appear to indicate that K-W boundaries ought to be preferred boundaries. Aust and Rutter have chosen to interpret these Q's as real activation energies, that is, as the energy barrier to the basic atomic process in the migration. If this is in fact true, then for each boundary the * It should be pointed out that if the predicted preference for highly porous boundaries in pure materials is sharp enough, it could conceivably lead to a marked preferred orientation in which the observed rotation amounts and axes are only those corresponding to boundaries with structures of the poorest fit possible. migration rate, G, should be related to Q by an equation of the form G = G, explained A plot of the data in the manner shown in Figure 13 indicates that this equation very probably does not hold for these boundaries. In Figure 13 the loga- rithm of the various migration rates have been plotted at fixed temperature against the corresponding Q's. In addition, the quantity exp ) has also been plotted against Q. It may be seen that at each temperature the two resulting lines, have widely different slopes -- G appearing to be relatively :: insensitive to Q. For example, G at 300°C decreases by only a feictor of about 4 when Q changes from 5 to 20 kcals/mole, whereas the exponential quantity decreases by almost six orders of magnitude. This means either the measuied Q's have little physical significance or there is some temperature- independent terin in G which in every case almost exactly compensates for the change in Q. The latter interpretation leads to the presumption of very high entropies of activation (for a discussion of this point, see page 29 and references 36 and 57) for which there seems to be no satisfactory ex® planation. On the other hand, theories of impurity effects do indicate that impurities may readily lead to apparent activation energies of migration which have no real relation to the underlying physical processes (see Part II). Thus, it seems probable that at least many of the Q's in Figure 11 are only apparent activation energies, and little physical significance should then be attached directly to the variations shown.* The migration rates themselves, on the other hand, are undoubtedly physically significant (although they may in some cases be average rates). In interpreting these, however, it should be kept in mind that variations in the striation structure probably introduced considerable variations in driving force - Aust and Rutter did not give the magnitude of such variations, but a factor of 2 or more would not seem un- likely. In addition, because of the 3-4º mutual misorientations amongst the striation subgrains, and other experimental limitations, the actual boundary misorientations cannot be known very accurately. As a result, it becomes .. apparent that a much larger number of rate measurements than shown in Figure 12 is required to decide conclusively whether the data indicate higher mobilities for K-W, near K-W or non-K-W boundaries. : In materials containing dissolved impurities the effect of porosity on mi- gration is considerably more complex than in pure materials. Here a change in boundary porosity has an indirect effect on boundary mobility as well as the direct effect already described. In a material of a specific bulk impurity concentration, a boundary with a high porosity supports a higher impurity segregation than one with low porosity; the boundary concentration CR is therefore higher in the former, the impurity drag on the boundary is greater and, thus, the mobility tends to be lower. As a result, increased 40, which corresponds to increased porosity, *ends through this impurity drag effect to produce lower mobility. However, the * It is worth noting that plots such as those in Figure 13 show similar results for all of the single boundary migration data based on zone-refined material, in- cluding other data of Aust and Rutter and those of Rath and Gordon. Thus, it 9. Chi seems likely that all of the associated Q's are only apparent activation anti energies, with the probable exception of the lowest values found by Aust and Rutter. . ... . . . . - 22 shapes of concentration function versus 40 curves and their relative positions for different boundaries depend in a complex way on the magnitudes of bulk impurity concentration, driving force and specific impurity-boundary inter- action energy and also on the temperature, as discussed in detail later in Part II of this paper). Briefly, in the mobility - Cg relationship there will be a region at low co and high mobility where the mobility is almost unaffect- ed by og because the segregated solute cannot keep up with the rapidly mov- ing boundary. On the other hand, at high cg and low mobility the mobility will depend more strongly on ce because the segregated solute now virtually travels with the boundary. The region between the high and low mobility segments will consist of a transition zone where (a) for boundaries subjected to low driving forces the mobility changes smoothly from the high to the low mobility region, but (b) for high driving forces the transition represents con- ditions for which the boundary migrates part of the time in the high, and part in the low, mobility regions, leading to an observed average apparent mobility which drops very rapidly from one region to the other. Since boundary concentration, GR, is a function of porosity, and porosity of 40, the effects of concentration on mobility may be translated into curves of concentration function versus 4,0, as has been done in Fig. 10b. The schematic curves shown have been drawn for conditions such that for the relatively low porosity boundaries, B and C, the virtually impurity-independent region of high mobility is observed at low A0, but in the almost random boundary, A, there is presumed to be enough porosity andthus, enough bility, impurity-dependent region. At high 40 where all the boundaries 23 tend toward similar porosities, it has been assumed that there is enough segre- gation at all these boundaries to place each of them in the impurity-dependent region. These conditions have been chosen for illustrative purposes because, according to the model, they are the conditions which produce preference for the special K-W boundaries of the type which have been most frequently ob- served. This is shown in Figure 10c, in which the mobility of the near-random boundary, A, and those of the medium-density and high-density coincidence- site boundaries, B and C, have been obtained as appropriate products of the two sets of curves in Figures 10a and b. It is seen that the K-W boundary of intermediate coincidence-site density is indicated to have a higher mobility than other boundaries when the orientations of the grains and the boundaries are ideal, that is, when AO is zero. Furthermore, the mobilities of high and medium coincidence-site density boundaries go through maxima as 40 in- creases, the highest overall mobility corresponding to that of some intermediate coincidence-site density boundary for which the relative grain and boundary orientations are not quite ideal. Other conditions of bulk concentration, driving force, impurity- boundary interaction energy or temperature may be chosen to give different concentration function -- A0 curves. For example, it may be supposed that a given impurity has such a strong affinity for the boundary that even very small absolute values of bulk concentration would effectively bring all the boundaries for all values of 40 into the low mobility region where the impur- ity atmosphere travels with the boundary. In this case the corresponding concentration functions would be very nearly alike for all boundaries, as shown in Figure 14a; multiplication by the porosity functions in Figure 10a 24 would then produce the mobility curves of Figure 146. It is seen that such conditions could again bring about a preference for near-random boundaries and the avoidance of ideal, high coincidence-site density boundaries just as was indicated for pure materials (Fig. 10a); this might then account for the Aust-Rutter observations on silver and gold alloys in zone-refined lead shown in Figure 6. Since high values of bulk concentration for impurities with lesser specific impurity-boundary interaction energies could presumably also bring about the same condition, a possible explanation of the results of Graham and Cahn (34) is suggested. The considerable quantities (0.4 wgt %) and the particular nature of the impurities present in their aluminum may well have been responsible for the avoidance of (111) rotation axes and the random distribution of rotation axes observed. 31 . Part II. Quantitative Impurity Effects, Theory and Experiment It is evident from the discussion in Part I that small amounts of soluble impurities can have a profound influence on the recrystalli- zation of relatively pure metals. As early as 1940, Beck (58) observed that the addition of 0.0083 wt % silver to lead greatly decreased the rate of recrystallization of this metal. Subsequently, the work of Smart and Smith (59) on copper and Lücke (60)(61) and co-workers on aluminium showed that the addition of as little as 0.01 at. % of a number of different impurities could raise the recrystallization tem- perature of these metals by as much as 220°c. These experiments, Shuninte!! however, were only exploratory and were not complete enough to separate the two component processes of nucleation and grain-boundary migration. It was, therefore, not possible to determine whether the impurity effect was due to a decrease in the rate of nucleation, in the rate of boundary migration, or ab seemed more probable, in both. Thus, it became clear that emphasis in subsequent work should be placed on obtaining quantita- tive data which dealt with the nucleation and boundary migration processes separately. Furthermore, the purity level of the parent metals used in these early investigations was probably not high enough to insure that residual impurity elements in the starting materials may not have masked to some extent the full influence of the added solute element. With the advent of zone-refining and the subsequent availability of extremely pure materials, it became possible to make more careful studies of the role of minute amounts of defined impurity additions. Thus it was that Bolling and Winegard (44) (62) studied post-recrystallization b - - ---- . . 32 graln growth in polycrystalline, zone-refined lead as influenced by the solutes tin, silver, and gold. Holmes and Winegard (45)(46) (63)(64) conducted similar experiments on zone-refined tin to which they added controlled quantities of either lead, bismuth, silver, or antimony. A different approach to the impurity problem was carried out by Aust and Rutter (35) (36) (37) who investigated the motion of individual grain boundaries in bicrystals of zone-refined leed containing very small additions of either tin, silver, or gold. In this case the driving force for migration was the melt-grown lineage substructure which remained stable during the experiments. This may be contrasted to the ever-decreasing driving force, derived from the self-energy of the grain boundaries, in the grain growth work. Gordon and Vandermeer (42) studied the impurity problem from yet another experimental point of view, one in which the driving forces for boundary motion were several orders of magnitude larger than in either the grain growth or the single boundary work. They measured the rates of growth of new grains during recrystallization after moderate deformation in polycrystalline, zone- refined aluminum and in this aluminum containing small amounts of copper. Along this line also, Frois and Dimitrov (65) (66) Investigated the influence of the solutes copper, magnesium, and silver on the recrystallization of zone-refined aluminum after heavy deformation, The one striking common feature in all of this work with zone- refined metals was the large retarding effect the addition of only small quantities of impurities exerted on boundary-migration rates. For example, 0.006 wt% tin dēcreased the speed of boundary migration of random boundaries at 300°C in lead by a factor of 1000 (35) while as little as 0.0017 at. % copper reduced the mobility of boundaries in zone-refined aluminum by something greater than 1000 at 139°c (42). It was also found that in a given zone-refined metal, one solute would retard boundary motion to a different degree than another. In lead, for example, silver and gold showed a much larger effect per solute atom than tin (37), while bismuth was more effective than lead in retarding grain growth in tin (44). Early Theories of Boundary Migration. The first theories of grain-boundary migration rates were put forth before the quantitative measurements on zone-refined metals were available. Mott (15) and Turnbull (67) derived expressions for the rate of motion of high angle grain boundaries based on the formalism of absolute reaction rate theory. Turnbull (67), defining the rate of boundary migration as the product of the mobility of the rate determining step and the gradient of the chemical potential in the direction of movement, assumed that the atoms are transferred singly across the boundary during migration. He obtained the expression G - rate of grain boundary migration, e = 2.718 (base of natural logarithms), , d = distance of boundary advance locally when an atom 18 transferred from one grain to the other, . 3 4 h = Planck's constant, N = Avogadro's number, AP= driving free energy per mole, as = entropy of activation for boundary motion, R = wiversal gas constant, Q = activation energy for migration, T = absolute temperature (°K). Expressed in terms of a diffusion coefficient for atom transport, Eq. 1 can be rewritten in the equivalent form (1-a) where a lattice parameter, D. - preexponential term, and Q = the activation energy for the diffusion proce88 responsible for boundary migration. Thus G 18 of the form and therefore Eqs. 1 and 1-a predict that a plot of log G against 1/T should give a straight line for which the slope 18 related to Q and the intercept at 1/T = 0 18 C. Most, though not all, experimental results are in good agreement with this straight line relationship, at least for the rather limited temperature ranges over which migration rate measurements can be made." - --.--veronder eine nove program.. Turnbull (67) further suggested that there is reason to expect the activation free-energy barriers which the atoms must surmount in the elementary act of grain-boundary migration to be similar to those surmounted in the elementary act of grain boundary self-diffusion. Thus, be supposed, as did Beck, Sperry, and Hu (68), that the rate determining step for grain-boundary mobility ought to be related to the diffusion coefficient for atom transport along the grain' boundary. Turnbull (67) offered some support for this point of view. He calcu- lated the free energy of activation for boundary migration, AF, from data obtained on metals of ordinary purity using Eq. 1.* On comparing these AF, 's to the activation free energies, of lattice self-diffusion, Turnbull concluded that the atomic mobility in grain-boundary migration is many orders of magnitude larger than in lattice self-diffusion. In spite of this, the very disconcerting fact remained that the activation energy, Q, in these materials was frequently equal to or greater than the activation energy for lattice self-diffusion (see Burke and Turnbull) (57). Also, the As, '8, which, based on theoretical considerations, should be nearly zero, were often very high. Turnbull (67) hypothesized that these anomalies might be attributed to inclusions which, retarding boundary migration, become less effective at elevated temperatures due to their solution coalescence. Mott (15) also attempted to account for the atoms are activated in group. He envisioned the atom transfer process to be one where a group of atoms belonging to one crystal melt * AFA 18 related to Q and A8A by the equation AFA - Q - TASA - RT. - 36 and resolidify as a group onto the other crystal. On this basis, an equation of the following form results: where a = the number of atoms in a group, T = the absolute melting temperature of the material, Q = ni, L being the latent heat of fusion of the material. There 18, however, no theoretical reason to suppose that atom transfer ia boundary migration occurs in groups rather than singly. Boundary Migration in Zone-Refined Metals. It seems to be clear now that the early attempts at comparing experimental rates to theoretical estimates proved unsatisfactory because the early theories took no account of impurity effects, whereas the data used for compari- son was obtained on "impure" materials. This has been brought to light by the more recent work carried out on zone-refined metals. Aust and Rutter (36) compared their experimental boundary migration rates in zone-refined lead containing tin with the predictions of the theories of Mott (15) and Turnbull (67). They found that with impurities present · neither theory predicted the correct order of magnitude for boundary . velocity. However, when Impurity effects were absent, the single pro- cess théory of Turnbull (67) gave order of magnitude agreement with experiment whereas the group process theory of Mott (15) was off by about three orders of magnitude. A similar test of the theories was 10 ..... .................. . ....... .. starten Pusthave made by Holmes and winegard (69) for migration rates observed during grain growth in zone-refined lead and tin. The experimental data were In excellent agreement with the single process theory predictions but not those of the group process theory, and here again impurity effects had been minimized. In a recent recrystallization study of zone-refined nickel, Detert and Dressler (70) were, led to interpret their experi- mental growth rates also in terms of the single process atom transport - - V .. . T .. .' A. . . . *Vane 1. Idea. } *.*..* It should be pointed out that the activation energy in the single process theory cannot be calculated theoretically with any precision and therefore when Eq. 1 is used to obtain theoretical estimates of G, the value used for Q must be taken from experiment. However, the magnitude of the preexponential factor, a Dor in Eg. 1-a can be calculated sufficiently well to permit a meaningful comparison with experiment provided the driving energy, i.e. AF, 18 known. Such a comparison was made by Gordon and Vandermeer (42) for their boundary AT I ... -. . - . -" ... come .. . . migration rate data on zone-refined aluminum. Table II shows this . . ., . . . . . . . wwtated . comparison and similar comparisons for experimental data obtained in other recent studies on zone-refined metals. The comparisons between calculated G'8 and experimental G's are given in terms of log Q. since in view of the estimations of D. and Af in the calculated quantities and the errors in the experimental quantities, only order of magnitude agreement can be expected (42). The theoretical and experi... mental values of Q, are seen to agree within approximately one order . . . .-. . . . haberini istening this time.commons . . . . . . . . n . . . . . . - tamba - - 38 Table II. Comparison of Theoretical and Experimental Preexponential Factors for Grain-Boundary Motion in Zone-Refined Metals Type Metal Driving Force (Cal/mole) Log lo Log G Theoretical Experimental Reference Experiment Aluminum 4.66 6.23 Aluminum . Recryst. Recryst. 28 Single boundary 2 x 10-3* Recryst. 10* . Single boundary 2 x 10-3 (42) (65) (71) Aluminum 4.58 0.96 4.89 1.09 Nickel . (70) 6.55 - 1.57. Lead random Lead Single boundary 2 x 10°3 special 1.09 - 0.15 : (36) ; Lead 0.99 (62) . . Grain growth Grain growth 1.8 x 10°3* 1.6 x 10^3* - 0.78 - 0.47 Tin 0.95 -- - - - - - - - - u us. -- --- Notes to Table I . D. assumed = 0.5 cm/sec d = lattice parameter a * driving force estimated . . . . . 1 .. ... 39 . . .. ........ . .. Hina S., of magnitude. This check is considered to be reasonably good, partic- ularly in view of the huge discrepancies--up to 10 or more orders of magnitude as indicated by the high ASA's mentioned earlier--found for similar comparisons on ordinary purity metals. Thus, there seems to be à mounting body of evidence which strongly indicates that the elementary process in grain-boundary migration involves the transfer of single stoms across the boundary. In all instances where the single process theory of Turnbull gives reasonable agreement with experiment, the activation energies of boundary migration are also much less than those of lattice self-diffusion. As shown in Table III, they are much closer to the activation energies found for grain boundary self-diffusion. Since the grain boundary is a transition region where the atoms do not fit together very well, it may, as pointed out in Part I, bė looked upon as a region containing excess porosity or vacancies. · The boundary could then be thought to move by means of an atom-vacancy Interchange. The activation energy of boundary - - - -- - . ... . . .. .. migration would in this case be approximately equal to the activation energy for the motion of vacancies in the boundary and should correspond to that for boundary self-diffusion. This has been found to be the case in zone-refined aluminum and nickel as Table III 11lustrates. The agreement 18 poor, however, for tin. Holmes and winegard (63) pointed out that the activation free energies of boundary migration in zone- refined lead and tin are also comparable to the activation free energies of liquid self-diffusion in these retals. Thus, while there is some . Let u Table III. Comparison of the Activation Energies for Grain Boundary Migration with those of Lattice and Grain Boundary Self-Diffusion QUB Metal Measured (Kcal/mole) for Lattice Self-Diffusion (Kcal/mole) Doponenden Em for Vacancy MI- gration (Kcal/mole) for Grain Boundary Self-Diffusion (Kcal/mole) Aluminum 34.08 16.0* 13.8 8 (42) 15.0 13.1 1.3.0 Nickel 30.1 66.86 26.6f 34.52 Lead iesire (Random bdy's) 6.0 (special bdy's) 25.7€ . 15.7° 6.1 - Lead 6.7 mer (62) Tin 6.0 24.50 9.55€ 12.21 (63) See Reference (72) (73) (74) 8 " h " 1 " ," *Est'd assuming (75) (76) (77) (78)(79) (80) (81) = 0.47, see Ref. (42) hii PMI. TRE U experimental support for the concept that the activation energy barriers which the atoms must overcome are similar in grain-boundary , migration and grain boundary self-diffusion, the evidence 18 not entirely in agreement on this point. : Impurity Drag Theories of Boundary Migration. When the predictions of Turnbull's single process theory are tested with data obtained on boundaries migrating in the presence of dissolved impurities, hugie discrepancies are noted. This is, for example, shown in Table IV, which compares the experimental G's for zone- refined aluminum containing additions of copper with the Go predicted by Eq. 1-a. Aust and Rutter (36) also found similar discrepancies for grain boundaries migrating in lead containing tin impurity atoms. It is evident from this and what has been said earlier that impurities play an extremely important role in the migration of grain boundaries--a role not explicitly taken into account by the single process theory of Turnbull. Theory of Lücke and Detert. Lücke and Detert (55) were the first to attempt to formulate a quantitative atomistic theory treating the impurity effect. The basic hypothesis of their theory was that impurity atoms in solid solution tend to segregate to grain boundaries because of interaction forces which exist between these atoms and the boundary (1.e., the boundary 18 an energy sink for ... Impurity atoms). They further assumed that the concentration of impurity atoms at the grain boundary, whether stationary or moving, 1. ot ri - ... - .:,: . . Table IV. Comparison of G, from Turnbull Theory with Measured G's of Dilute Aluminum-Copper Alloys Log G. Material (ppm copper) Log G. (single process theory) (measured) 6.23 Zone-Refined Aluminum 4.66 2.1 14.3 4.66 4.66 4.3 17.5 11.3 4.66 17.0 34.0 10.5 4.66 68.0 4.66 4.66 124.0 258.0 4.66 18 given by the approximate equilibrium expression for a stationary boundary . E Co Cexpre where Ce grain boundary impurity concentration, C = bulk impurity concentration, E -= interaction potential between impurity atoms and grain boundary. E is taken to be negative if impurities are attracted to the boundary. When a grain boundary laden with impurity atoms is caused to move, the impurity at sphere creates a drag which, if not overcome by the driving force, compels the boundary to carry the impurity atmosphere along with it. In this situation the speed of the boundary is con- trolled by the rate at which impurity atoms can diffuse along behind the boundary. Ilicke and Detert obtained for the rate of migration under these solute-controlled conditions an equation of the form com- where the quantities are as defined previously. Since the impurity atoms are envisioned as moving behind the boundary and making their diffusion jumps towards the boundary, the diffusion constants D. and my were assumed to refer to lattice diffusion of the Impurity. 44 According to this theory, if the driving force 1s high enough to overcome the dragging effect of the impurities, then the boundary breaks away from the impurity atmosphere. This will occur at low impurity concentrations or at high temperatures where accord- ing to Eq. 3, Co will be small. In this "broken away" state the rate of boundary motion 18 assumed to be independent of impurities and the activation energy for migration will be that for grain- boundary diffusion, 1.e. Quapro Figure 15 shows schematically how G should vary with temperature for a given alloy composition based on the Ilcke-Detert theory. At high temperatures G 1s solute-independent and determined by Eq. 1 (1-2), whereas at low temperatures G depends on composition in the manner. indicated by Eq. 4. According to the Lücke-Detert impurity drag theory, the breakaway process should be very abrupt and should occur at certain critical values of composition, driving force, and temper- ature. It was recognized, however, that in reality the breakaway would not be sharp but rather that a transition region, in which the boundaries would be only partially broken away, should be expected, as depicted in Fig. 15. The extent of this transition region can not be ascertained from the theory, however. Figure 16 illustrates schematically the temperature dependence of boundary-migration rates predicted by the Lücke-Detert theory for a series of impurity compositions. Experimentally it is not feasible to study migration rates over as wide a temperature range as 18 4 5 . MN. -- --- .. . - -.- -. . . . . necessary to obtain any one complete curve in Fig. 16. As suggested by this figure, however, it is possible to change C sufficiently 80 that at some composicions the measurable range of G lies in the impurity controlled range (for example Cy, Cz, and Cz), at others in the transition region (C. and Col, and at still others in the impu- rity independent region (C). Thus, from theory the experimental velocities would be expected to fall on those segments of the indi- vidual curves which lie in the cross-batched area of Fig. 16. AB may be seen in Fig. 17, the grain-boundary migration rates for alloys of copper in zone-refined aluminum from the studies of Gordon and Vandermeer (42) follow the pattern suggested by Fig. 16. The data of Frois and Dimitrov (65) for copper in zone-refined aluminum also show this same trend, though their data 18 not as extensive as that of Gordon and Vandermeer in the high concentration region. Further- more, it may be deduced from Fig. 16 that in the experimentally measurable range of G, the "activation energy" (1.e. -RIIG I *se a 1/1 ought to be low for low compositions and with increasing c go through - Rd In G a maximim and eventually level off. For low compositions ....... . .... . . . .. . . . . . .... . ... .rar . . should equal Que and at high compositions Q-E. It should be pointed out that all other "activation energies" are only apparent and do not represent an actual energy barrier which must be overcome during the movement of the boundary. The trend of the measured "activation energies" for the Al-Cu alloys 18 just that predicted as can be seen in Fig. 18. A similar trend 18 also noted in the data of Frois and Dimitrov (65) for magnesium in zone-refined aluminum. . . The excellent qualitative agreement between the Lucke-Detert theory and experiment already noted for zone-refined aluminum was not found in other investigations on impurity doped, zone-refined metals. Figure 19 shows the composition variation of the activation energy for grain growth in zone-refined tin to which lead has been added, from the work of Holmes and Winegard (46) (64). At the higher concen- trations of lead the activation energy, to be sure, is constant in agreement with the Lücke-Detert idea. In addition, the activation energy is low for zone-refined tin with no lead, which 18 as expected. At intermediate lead levels, however, the activation energy curve does not go through a maxima as Llcke and Detert would predict, a behavior which Holmes and Winegard, (46) suggested was in contradic- tion to the breakaway concept. Mkewise, Aust and Rutter (36) found that the measured activation energies of random grain boundaries in zone-refined lead alloyed with tin did not behave in a manner con- sistent with the Ilcke-Detert breakaway process either. They observed, as Fig. 8, page indicates, a steadily increasing activation energy as impurities were added, with no obvious tendency to level off at least up to tin concentrations of 0.0013 wt %. It would be interest- ing to know whether or not at higher tin contents the upper curve of Fig. 8 would go through a maximum and begin decreasing again. Another aspect of boundary-migration behavior predicted by the Jücke-Detert theory has been tested with experiment; this is the . variation of G with C at constant temperature and driving force. The . : -- :. expected behavior is 1llustrated schematically in Fig. 20. At low . impurity concentrations where Eq. 1 (1-a) holds,, the migration rate : $should be high and independent of C as shown. The results from Frois. and Dimitrov (65) for impurities in zone-refined aluminum are good examples of this type of behavior as Fig. 21 indicates. At high C, the rate is supposed to vary inversely with composition in accordance with Eq. 4. The corresponding experimental data of Gordon and Vandermeer (42) on migration rates in aluminum-copper alloys at several different temperatures are plotted in Fig. 22. It is seen that the data give curves of the predicted shape. Here also the results of Frois and Dimitrov (65) are in qualitative agreement with the predicted behavior. When the Lücke-Detert equation for impurity controlled boundary migration 18 expressed in terms of the rate of grain growth, it . becomes 11/2 Q. 20 Yan RT RT . where A = average grain diameter, t = annealing time, a = grain growth exponent, usually = 2, 0 = specific grain-boundary energy per unit area, ... Ve molar volume. "Holmes and Winegard (46) found that in the composition range where the activation energy 18 constant, the rate of graia growth for lead ,-:. - - . wa watu wote : in & in tin is proportional to cau/. According to Eq. 5 this aspect of grain growth is in agreement with the Lücke-Detert theory. Aust and Rutter (35) (36), on the other hand, had little success in finding agreement between the Lucke-Detert theory and their experi- mental results on the migration of single boundaries in zone-refined tin alloys. They calculated that on the basis of the Lücke-Detert theory, tin contents greater than 1 x 10° atom fraction and velocities less than 11 mm/sec should give rise to impurity controlled boundary migration behavior in zone-refined lead. Their data, shown in Fig.23 1llustrates that the variation of measured velocities with tin content at 300°C for random boundaries does not in fact follow the expected pattern. For example, at the higher tin contents, the velocities are proportional to c-5/2 rather than cot. It 18 necessary when considering only the role of impurities in boundary migration as the Lücke -Detert theory does that all orienta. tion variables be held constant. For single-boundary measurements this would require that identical boundaries are studied in each of the different alloys. Gordon (43) has suggested that the lack of Agreement between the A ist and Rutter data and the Lücke-Detert theory may be a result of the tremendous difficulty involved in 1solating orientation variables from impurity variables. Thus while Aust and Rutter managed to separate their boundaries into three groups with respect to orientation variables, there are probably differences in orientation effects within each group. The . • -...- www 1 49 Come . . i in acest modern mit vero experimental techniques used in polycrystalline boundary-migration measurements, however, incorporate a statistical assurance that only boundaries with approximately the same environment present themselves for measurement. As a result, polycrystalline samples, properly prepared and tested, automatically maintain the orientation -.. -- variables approximately constant and may offer a readier means for ---- - - ... .. . ........... . ... . . . testing the Lucke-Detert theory quantitatively. When further attempts were made to test the Lücke-Detert theory quantitatively with those experiments where the predicted trends were observed, areas of disagreement were uncovered. Theory suggests that in the impurity-dependent range, the activation energy for migration should be equal to @-E. For copper in aluminum @ze 31, 120 cal/mole (82). Assuming E to be about -2280 cal/mole (1), Q4 - E 18 33,400 cal/mole. Gordon and Vandermeer (42), however, observed an activation energy for migration of only 28,000 - 29,000 cal/mole for copper in zone refined--about 5 to 6000 cal/mole too low. A similar effect was found by Holmes and Winegard (46) in grain growth. Thus it was that when Holmes and Winegard calculated solute diffusion coefficients from their growth-rate data, using the Illcke-Detert theory, the calculated rates appeared to be about two orders of magnitude too large. . . . . . . . . . . .. ..-_.,...... Soinin il serience der mir contra os indication Landing * . : cal . . . Modifications to the Lucke-Detert Theory. Gordon and Vandermeer (42) : tried to account for these discrepancies. They took issue with the Lücke-Detert assumption that in impurity-controlled migration the atoms i. . . . . . . in Chem . . . . artawan dan mening . . . w .. . . may be treated as diffusing in undistorted bulk material, which led to the insertion of D. and Q, into Eq. 4. It was pointed out by Gordon and Vandermeer that, strictly speaking, the impurity atoms can only exert a dragging force on the boundary in a region where the lattice is still distorted to some extent by the presence of the grain boundary (42). They argued that since the distortion of the lattice can be expected to lower diffusion activation energies, Quy should be substituted for Q in Eq. 4, where Qy may be significantly less than QzIt was not possible, however, to decide how much lower Qz should be than , but it was suggested that it could never be less than Qce and that it was probably much closer to Q than QGB One further modification in the Lücke-Detert theory was also introduced by Gordon and Vandermeer (42). They suggested as McLean (1) nad pointed out, that the boundary concentration given by Eq. 3 above. is only an approximation and did not take into account alterations in the vibrational frequencies of the impurity atoms when they are i placed into a grain boundary. They proposed that the Ilicke-Detert formula be modified to .. -. . -.. AF 1 com ( RT AC RT where A 18 a vibrational frequency factor approximately equal to 1/4 and D. and af are impurity diffusion constants referring to impurity diffusion in the distorted grain-boundary region. With these modifications in the theory, the data on copper-doped zone-refined NEX VO 1 . . T .. . .. . . . . .. . simon.eran.. n ä. E , ........ ir prisimine t e le this band and cansations are presenti n renant -! aluminum were found to be in excellent quantitative agreement with theory. For example, when the preexponential portion of Eq. 6, i.e. Go, was calculated from theory and compared with experiment, quite good checks were obtained except for alloys in the transition region where agreement 18 not expected. Table V shows this comparison. Subsequently, Gordon (43), also showed that the Winegard and Holmes data on grain growth in alloys of zone-refined tin contain- ing lead and bismuth could also be satisfactorily explained by the modified theory. Gordon pointed out that because of the low-driving forces present in grain-growth studies, the "breakaway". behavior will manifest itself in a way different from that in recrystallization studies where the driving forces are much higher. This is schemati- cally illustrated in Fig. 2A for log G vs 1/T. If the breakaway temperature, T, 18 much less than Tes defined as the temperature where Eqs. l-a and 6 cross, then the behavior shown in Fig. 24-a is expected. If, however, T, is approximately equal to T., then the behavior depicted in Fig. 24-18 likely. Gordon (43) showed that in the Holmes and Winegard grain-growth data T. 4 T. and the latter behavior is expected. It can also be seen from Fig. 24 that in going from the solute independent to solute dependent regions, in the first case the "activation energy" increases, goes through a maximum and decreases again before leveling off. In the second case, however, the "activation energy" increases and gradually levels off without going through a maximm. The experimental behaviors shown in Figs. 18 and 19 are therefore both consistent with the modified Ilicke-Detert impurity drag theory. L W .. . "::.. . . . . 52 Table V. Comparison of Theoretical G's with Experiment for Impurity-Controlled Boundary Migration in Zone-Refined Aluminum Material pom cu Log G. Meas: Log G. Calc from°EC. 6 2.2 4.3 17.0 34.0 68.0 124.0 258.0 (14.3) (17.5) 11.3 10.5 9.6 11.0 10.7 10.1 9.8 9.4 9.3 9.2 8.8 9.1 Concerning the relative effects of different solute atoms on boundary-migration rates, some qualitative comparisons have been made with the predictions of the Lücke-Detert theory. Arct and Rutter ---. ---rend the windy..'- found that per impurity atom, gold and silver were more effective in retarding single-boundary migration in lead than was tin. This they claimed was in contradiction to the predictions of the theory because gold and silver being faster diffusers in lead than tin should, according to Eq. 4, not have as large a retarding effect as tin. This claim is certainly valid if all three solutes have about the same interaction potential with the grain boundary. It is probable, however, that solutes which diffuse rapidly and thus have low QE will also have large, negative interaction energies, E. Since the quantity Q - E, rather than Qy alone, 18 controlling in impurity dependent boundary migration, it follows that Qe - E may be larger for fast diffusing, than for slow diffusing solutes. Thus, the Aust and Rutter results could be consistent with the Lücke-Detert theory as modified by Gordon and Vandermeer. The need to modify the Ilcke-Detert theory to obtain satis- factory quantitative agreement with experiment has emphasized the desirability of a more sophisticated theory. The main assumptions, approximations and limitations of the Illcke-Detert theory were: .: a) The prediction that breakaway from an impurity atmosphere 18 an abrupt process. .- . this thin chambre ja teises staan en diensten toiminnan toiminnasta mielenkiin section les entre basement des Formation on 54 b) The assumption that the velocity of the grain boundary 18, equal to the velocity of the impurity atoms. c) The implicit assumption that Cg 18 given by the stationary equilibrium value and that it does not change when the boundary moves. a) The consideration only of segregation in which impurity atoms are absorbed at boundaries, but not that in which they are desorbed from boundaries, 1.e., E. positive. e) The assumption that diffusion of the impurity atoms in undisturbed bulk material controls the solute-dependent migration behavior. 1) The lack of consideration that the boundary-impurity interaction energy and the diffusion coefficient of the impurity vary as a function of distance across the grain boundary. 3. The Impurity-Drag Theory of Cahn. Recently Cahn (54) attempted & more rigorous formulation of the impurity drag problem in grain- boundary migration. He sought to remove some of the limitations of the Ilcke-Detert theory retaining, however, their fundamental assumption that it is the force exerted by an impurity atmosphere which is responsible for the observed phenomena. Simultaneously and independently of Cahn, Lücke and Stuwe (6) albo treated the Impurity drag problem in more detail. Their treatment was in many respects identical to that of Cahn. A brief review of Cahn's . . formulation is as follows. He supposed that as far as the impurity atoms were concerned, the grain boundary could be represented as a planar discontinuity characterized by an interaction energy, E(x), and by a diffusion coefficient, D(x), for motion normal to the boundary. Both E(x) and D(x) are functions only of the distance x normal to an arbitrarily chosen center plane of the boundary. By . --dinner .warw.rairera considering in detail the diffusion or the impurity atoms in the . mm --- .-. - ..-.--. potential field of the boundary, Cahn was able to derive an expres- sion for the composition profile across a boundary moving with a constant velocity, G, explicitly in terms of integrals involving E(x) and D(x). Due to the complexity of this expression and because of a lack of knowledge about the functions E(x) and D(x), Cahn was unable to obtain a general solution for the composition profile and was forced to consider only certain limiting cases. He was able to solve for the total force exerted on the boundary by all the impurity atoms under conditions such that the velocity of the grain boundary was either much higher or much lower than the drift velocity of the impurity atoms. Cahn then constructed an approximate equation for the impurity drag force, Pe, which fit both these high and low- velocity extremes. Finally, by suming all the forces acting on the grain boundary, Cahn obtained the following expression relating the velocity of boundary migration to the driving force, composition, . and indirectly to temperature: -". ini а са AF = NG 4 56 In Eq. 7, 18 the reciprocal of the boundary mobility in pure specimens. Alpha and beta are referred to as impurity drag parameters and are related to E(x), D(x) and temperature through the integrals Sinha E(x) RT 2 RT et dx aux (8) RT D(x) dx , and D(x) dx , () where R = universal gas constant, T = absolute temperature, Ve molar volume of parent material, D(x) = diffusivity profile of the impurity atoms, E(x) = boundary-impurity atom Interaction energy profile. The drag parameter a can be thought of as the reciprocal of the boundary mobility at unit concentration of impurity while B represents the reciprocal of the average drift velocity of the impurity atoms across the grain boundary. In very pure materials, i.e. C20, Eq. 7 becomes AF » G (10) which 18 of the form of the single process theory of Turnbull, i.e., Eq. 1(1-2). In view of the fact that this theory has been shown to yield satisfactory agreement with experiments on zone-refined metals (see earlier discussion), it is possible to evaluate , by comparing . : 57 demers Eq. 10 with Eq. 1-a. Thus ^ can be written as nibüste naar het Na GB RT expan RT . (11) eren " t . --.. '-. .... : nis in i i -ni-. merkintäm .. -- . . . , . - . ...: "" Since the important basic factors comprising the impurity : drag parameters a and B are D(x) and E(x) and these are involved as integrals, it le far more difficult to obtain a and B in terms of fundamental quantities. This is brought about mainly because the profile shapes of D(x) and E(x) across grain boundaries are not known. Furthermore, the magnitudes of D(x) and E(x) are known, if at all, only approximately. Nevertheless, it 18 possible to devise some simple, albeit crude, profiles for D(x) and E(x) which allow integration of Eqs. 8 and 9; this permits the derivation of relation- ships between the theoretical drag parameters, a and a/b', on the one hand, and the impurity diffusion coefficients, Dr. and DER, the . . grain boundary half-width, 6, and an interaction potential at the boundary center, E., on the other. This has been done for the assumed profiles shown in Fig. 25. The profiles are defined only for x positive, but it should be kept in mind that they are symmetric about the grain boundary center. The interaction potential was arbitrarily chosen as negative indicating absorption of impurity atoms at the grain boundary. The solutions of Eqs. 8 and 9 for various combinations of D(x) and E(x) are listed in Appendix A. The combinations of E(x) and D(x) chosen were ones which approximate the ..... . ... ... ... . .. .. .... ...... . .. . T mim . -- Ir... erw .. . . . . . . . . . . . 58 types of boundaries distinguished by Cahn (54) (1.e., flat or pointed extremms in E(x) at the boundary center, and different relative ranges over which E(x) 18 altered compared to D(x)). In Cahn's theory, the expression which relates the velocity to driving force, composition and temperature, 1.6., Eq. 7, 18 unfortunately a cubic algebraic expression in G. Thus it is diffi- cult to determine from it the velocity for a given set of conditions. The situation becomes somewhat less complicated when certain extreme cases are examined. If, for example, the velocity of the boundary 18 less than the drift velocity of the impurity atoms in the boundary, 1.e., G < 1/8, then Eq. 7 simplifies to . AF = NG + a CG . (12) a + QC AF This 16 considered a low velocity in the sense of the Cahn theory and 18 attained either by a low-driving force, 1.e., AF/A < 1/8, or a high-impurity content - . Equation 12 is equivalent to the older Lücke-Detert formula 1f QC >> A, D(x) 16 constant and equal to a lattice diffusion coefficient and E(x) 18 large and ac B negative (absorption of impurities). Thus the Lücke-Detert theory appears as a limiting case of the much more general theory of Cahn. .. V . .. 1.. Another limiting cabe occurs when the velocity of the boundary ,18 much greater than the drift velocity of the impurity atoms, i.e., G>> 1/B. In this high-velocity region + RU f (13) closer-menomis Equation 13 reveals that in the high-velocity range the velocity 18 not strictly proportional to driving force except under condi- tions for which the second term is negligible, 1.e., for low C or high AF. Also, since a/82 18 proportional to a diffusivity (see : Eqs. in Appendix A), Eq. 13 indicates that impurities having greater diffusivity will show a greater retarding effect, provided the other terms that make up a/B2 are the same. There are, according to Cann's theory (Eq. 7), three types of transition to be expected in going from pure to impure material. These three types of transition behavior are 1llustrated in Fig. 26, which is essentially a plot of boundary velocity versus impurity concentration on a log-log scale. As indicated, it is the magnitude of the driving force which determines which one of these transitions ought to be observed. At low driving forces the transition 18 smooth and continuous suggesting a gradual capture of the grain boundary by impurity atoms. For intermediate driving forces the transition 18 still continuous but there is an inflection in the variation of velocity with composition. The range of driving forces over which this second kind of transition occurs is, however; quite limited and thus it would be observed only infrequently. At high driving forces a third type of transition behavior 18 predicted. In this case there appears to be, for a certain composition range, three velocities .. ... M --- - tier . - mana mes.com. - mana Se mere e . S omein tantummod den n risi which satisfy Eq. 7. Cahn pointed out that the middle velocity represents a physically unstable situation and that only the highest and lowest velocities in this composition range can be expected to persist for finite periods. Thus, for high driving forces there is a range of composition where the theory predicts that there are two possible velocities for grain-boundary migration, with a discontinuous transition between them. This raises the question as to what kind of experimental behavior is expected for boundaries falling in the composition range where these two branches of the velocity-composition curve overlap. It should be remembered in this connection that when the velocity of a grain boundary is measured experimentally, the macroscopic, and hence only the average, behavior of the boundary is really observed. In fact, boundary velocities can be measured metallo- graphically only after the boundaries have moved tens to hundreds of thousands of atom distances. As a result, localized fluctuations in migration rates would go unnoticed and only an average velocity would be observed. If it may be assumed that there are perturbations present which can alter the boundary velocity to and from the extremums in the two-velocity range, then it can be imagined that a given boundary segment might oscillate between these extremas. The perturbations could undoubtedly arise from local fluctuations in impurity concentration and/or in driving force. Macroscopically, therefore, an apparent velocity at a value between the two actual extremum velocities would be measured. This situation is schematically ressions to interp weermin www **.com 1:2lustrated in Fig. 27-2 where a growth curve representing the plots made during recrystallization studies of grain-boundary migration 16 shown. The stepwise behavior represents the Oscillations between actual velocities, whereas the macroscopic resultant is indicated by the heavy line through the steps. This bebavior may even be more complicated. For example, suppose a local segment of a grain boundary moving with a low velocity encounters a region free of impurities in which it 18 temporarily able to move at a high velocity. The fast moving segment then loops out ahead of the slower moving portions of the boundary, introducing local curvatures in the boundary. If these curvatures are high enough, the driving force on the boundary will be altered by a term involving the product of the grain boundary self energy and the curvature. This surface tension contribution would act to slow down the fast moving portion of the loop; buit, since inverse curva- tures would be generated at positions where the loop joins the slower- moving boundary segments, thua augmenting the driving force there, . these segments would move at a slightly increased speed. On the average, however, the overall boundary velocity will be increased to a value above that given by the low-velocity branch. Thus, on a local scale, boundary energy considerations may well be important in deciding what average velocity the grain boundary assumes in the two- velocity regions. In any event, the average velocity should be intermediate in value between the high- and low-velocity extremes predicted by the theory. Wu... CS VI www wani : 62 The consequences of this rationlization of the behavior of boundaries in the two-velocity region are 1llustrated in Fig. 20-6. The heavy line indicates the sort of behavior expected of the experimental rates. The similarity between this and the curve predicted by Gordon and Vandermeer from the Lücke-Detert theory (see Fig. 20) should be noted. Thus, for high-driving forces, the ndivision measured velocities can be regarded as falling in one of three regions corresponding to either the high-velocity branch, low-velocity branch, or a transition region (two-velocity region) with velocities inter- mediate to these. Apparently the two-velocity region in the Cann theory corresponds to the partial breakaway (transition) regions in the Lücke-Detert theory. There was, however, no way to decide from from the Lücke-Detert theory how wide the transition region should be. In this respect the Cam theory offers improvement for it specifies that the transition region should be as wide as the overlapping of the high and low branches of the velocity versus composition curves. Thus, with reference to Fig. 29-the limits of the transition region can be defined by the outermost extremity, Cy, of the high-velocity branch and the innermost extremity, Cg of the low-velocity branch. These extremities can be estimated in terms of AF, 1, a, and B. The procedure 18 given in Appendix B. The result 18 that alloys with compositions in the range ntissimenticiniame stomasacinio teikiamavaraistai turnirwani Vn . . . : (14) can be expected to show transition region behavior. . Sam 63: There are two other aspects of the Cana theory which differ from the Llcke-Detert theory. First, Cahn's theory predicts that impurities which avoid the grain boundary, 1.e., E(x) positive, will produce the same drag effect as those which are absorbed, i.e., E(x) negative. The former are pushed ahead of the boundary, the latter dragged along by the boundary. The Lücke - Detert theory did not explicitly consider the case of impurities which avoid the boundary. Second, Cahn's theory specifies which diffusion coefficient is appropriate 1f D(x) 18 not constant and equal to Dp: Thug, one of.. the modifications Introduced by Gordon and Vandermeer (42) 18 automaa: tically incorporated into the Cahn theory. While Cahn's theory has taken care of many of the limitations of the Lücke --Detert theory, it still does not explicitly take orientation effects into account. It should be pointed out, though, that there are provisions in the Cahn theory for crientation effects to be included. This could be accomplished 1.f the profiles of E(x) and D(x) were known for all the different orientation relationships between grains. However, this does not appear to be possible at the present time due to insufficient knowledge about the atomic structure of grain boundaries. Although the Cahn theory makes certain predictions with regard to boundary-migration behavior, it is restrictive in the sense that it still does not allow a decision from first principles as to whether a measured velocity 18 a high velocity or a low velocity, or whether a given driving force 16 a high one or a low one. This 18 2 OF 2 ORNL P 1722 . hot 3 . : . . H4 5 50 ** 5 6 ENYA - 11:25 | 1.4 1.6 MICROCOPY RESOLUTION TEST CHART NATIONAL BUREAU OF STANDARDS - 1963 . r. fr grounds. They also pointed out that in the composition range where G 18 proportional to 1/23, the activation energy for migration should be constant, whereas Aust and Autter observed an ever- increasing activation energy. This led them to suggest that the apparent agreement mentioned above may be only fortuitous. (20) viewed the impurity effect in another way. He assumed that impurity atoms segregate to the boundary in order to reduce the porosity of the boundary. His approximate treatment led to the following expression for the velocity. G (C = 0) (Avº - AV*) KC ln - (15) RT 1 Kc Where r 18 a proportionality constant, avo 18 the porosity of the boundary with no impurity atoms, AV* 18 the porosity of the boundary when it is completely saturated with impurity atoms, K is the equili- brium constant for segregation and C is the composition. This theory leads to a saturation effect in the velocity at high compositions, an effect which is not considered in either the Tücke - Detert or Cahn theories. Li (51) compared the predictions of Eq. 15 with the experimental data of Aust and Rutter (35) for random boundaries in alloys of lead containing tin at 300°c. He pointed out that the variation of in hon with concentration for this data followed the trend predicted by Eq. 15. He was also able to calcula te a standard free energy of absorption for tin to the grain boundaries of-lead and found a value of 11,500 cal/mole. However, in order to make a significant quantitative check of the theory of L (20) more Information 18 required regarding the porosity of the boundary with and without impurity atoms. and comprehensive Correlation of Cahn's Theory with Experiment. No one has yet tried to correlate Cahn's theory quantitatively with experiment. In this section of the paper an attempt is made to analyze some of the available experimental data on grain-boundary migration rates in terms of the theory with the aim of calculating the parameters a, B, and a. It is hoped this analysis will indicate just how detailed experimuntal rate studies must be to determine the se parameters. At the present time, only two sets of experimental data are extensive enough to make these calculations: the recrystallization study of Gordon and Vandermeer (42) on copper in aluminum and the grain growth study of Holmes and winegard (46) on lead in tin. To a limited extent, the data of Frois and Dimitrov (65) can be used to supplement that of Gordon and Vandermeer. Since for grain growth work, the equations of Cahn must be modified, it is most convenient to discuss these two sets of experiments separately. 1. i Boundary Migration in Recrystallization; Copper in Zone-Refined Aluminum. If the driving force is known, I can be calculated very simply from Eq. 10 and the measured migration velocities of the zone-refined metal. This assumes, of course, that the zone-refining has produced metal pure enough so that residual impurity effects are negligible. Figure 28 summarizes the temperature dependence of 1/4 for zone-refined alminum using the velocity measurements of several different investigators (42)(65)(71) (83) and the driving forces in Table II. The 1/2's calculated from the Gordon and Vandermeer (42), Dimitrov and Frois (65), and the Gordon and El Babayonni (83) recrystallization data agrer quite well and can be represented by a single equation of the form 1 = 4, exp (16) with the constants 1. and L equal to 1.5 x 10°3 cal/mole/cm/sec and 14,000 cal/mole respectively. The 1/2's estimated from data of Aust (71) cannot be fitted to Eq. 15 with these same constants; they seem to yield the same activation energy but the preexponential factor falls about 1 to 2 orders of magnitude too low. Since the Aust data was obtained on single boundaries, this discrepancy may reflect an unknown inherent difference between single-boundary migration and the growth of grains during bulk recrystallization. In succeeding calculations it will be assumed that , is given by Eq. 16 with the constants listed above. The calculation of a requires that some of the experiinental velocities fall entirely within the low velocity region of the Cahn theory. Such conditions will prevail at high-impurity levels; in this case Eq. 12, which predicts that the reciprocal of velocity be. a linear function of solute composition at constant driving force s , and temperature, should apply. As Fig. 29 shows, again from the data of Gordon and Vandermeer, such a dependency of velocity on composition does in fact exist for zone-refined aluminum alloys containing copper in excess of 17 x 10*° atomic fraction. The vertical lines through some of the data points represent the range ما م انی of estimated uncertainties in the measured velocities. Within these uncertainties, straight lines can fairly well describe the data. Equation 12 also predicts that the intercept at zero composition for the plots shown in Fig. 29 should equal 1/AF. However, it appears experimentally that this intercept is zero, a result implying that ^< - Calculation of the Extremities of the High and Low Velocity Branches in the Cahn Theory for High Driving Forces Equation 7 expressing the interrelationship between G, C, T, etc. 18 rewritten in terms of the composition OY coate Srper G - Am B - (BI.) For the experimental results discussed in this paper the last term (1/a) 18 always negligible. For high velocities the first term of this equation can also be neglected. mbus :c- G - A e B . (Be) The high velocity extremum Cz can be calculated by differentiating Eq., with respect to G and settiug the result equal to zero. We obtain (B3) at this extremity. Subøtituting (B3) Into (B2) gives (BH) The other extremity Ca defining the macroscopic transition region can be arrived at by considering Eq. Bl in the 11ght of low velocity behavior. In this case, the third term in Eq. Bl becomes negligible Bo that Sande (B5). Again differentiating with espect to G and setting the result to zero yields (B6) at Ca. Substituting (B6) back into (B5) gives Ca - Part 3 (B7) Thus, alloys whose compositions fall in the range | << ple (B8) may be expected to show the macroscopic transition behavior. BIBLIOGRAPHY 1) D. McLean, Grain Boundaries in Metals, Oxford University Press, London, 1957. 2) M. C. Inman and H. R. Tipler, Interfacial Energy and Composition in Merals and Alloys, Metallurgical Reviews, 8, 30 (1963), 105. 3) F. Weinberg, Grain Boundaries in Metals, Progress in Metal Physics, 8, (1959), 105. 4) J. H. Westbrook, Segregation at Grain Boundaries, Metallurgical Reviews, 2, 30 (1964) 415. 5) K. T. Aust and J. W. Rutter, Grain Boundary Migration, in Recovery and Recrystallizatior of Metals, Interscience Publishers, New York, 1963, 131. 6) K. Lucke and H. Stuwe, On the Theory of Grain Boundary Motion, ibid, 171. 7) C. S. Barrett, Structure of Metals, McGraw Hill Book Company, New York, Chapter XIX. 8) 1. L. Dillamore and W. T. Roberts, Preferred Orientation in Wrought and Annealed Metals, Dept. of Metallurgy, University of Birmingham, England, unpublished report. 9) S. Washburn and E. R. Parker, Kinking in Zinc Single-Crystal Tension Specimens, Trans. AIME, 194 (1952) 1076. 10) C. H. Li, E. H. Edwards, J. Washburn and E. R. Parker, Stress-Induced Movement of Crystal Boundaries, Acta Met, 1, (1953) 223. 11) D. W. Bainbridge, C. H. Li and E, H. Edwards, Recent Observations on the Motion of Small Angle Dislocation Boundaries, Acta Met, 2 (1954) 322. 12) E. W. Muller and K. Bahadur, Field lonization of Gases at Metal Surfaces, Phys. Review, 102 (1956) 618. 13) S. S. Brenner, Field Ion Microscopy Studies of Surfaces, in Metal Surfaces, ASM, Metals Park, Ohio, 1963. 14) F. Hargreaves and R. J. Hills, Work Softening and a Theory of Intercrystalline Cohesion, J. Inst. Metals, 41 (1929) 257. 15) N. F. Mott, Slip at Grain Boundaries and Grain Growth in Metals, Proc. . Phys. Soc., London, 60 (1948), 391. 16) T. S. Ke, A Study of Recrystallization and Grain Growth by Measurements of Internal Friction, Trans. AIME, 188 (1950) 581. 17 W. T. Read and W. Schockley, Dislocation Models of Crystal Boundaries, Phys. Rev. 78 (1950) 275. 18) J. H. van der Merwe, On the Stresses and Energies Associated with Intercrystol- line Boundaries, Proc. Phys. Soc., 63A (1950) 616. 19) R. Smoluchowski, Theory of Grain Boundary Diffusion, Phys. Rev. 87 (1953) 482. 20) J. C. M. Li, High-Angle Tilt Boundary. A Dislocation Core Model, J. App. Phys., 32 (1961) 525. 21) P. A. Beck, Theory of Annealing Textures, Trans.AIME, 191 (1951) 475. 22) W. G. Burgers and T. J. Tiedema, Notes on the Theory of Annealing Textures, Acta Met., 1 (1953) 234. 23) C. G. Dunn, On the Theory of Secondary Recrystallization Texture Formation in Face - Centered Cubic Metals, Acta. Met., 2 (1954) 386. 24) S. Kohara, M. N. Parthasarathi and P. A. Beck, Annealing Texture in a Rolled and Artificially Nucleated Aluminum Single Crystal, Trans. AIME, 212, (1958) 875. 25) B. Liebmann, K. Lucke and G. Masing, Untersuchen uber die Orien- tierungsabhangegkeit der Wachstungeschwindigkeit bei der primaren Rekristallisation von Aluminium Einkristallen, Z. Metallkunde, 47 (1956) 57. 26) H. Yoshida, B. Liebmann and K. Lucke, Orientation of Recrystallized Grains in Strained Aluminum Single Crystals, Acta. Met. I (1957) 51. 27) K. T. Aust and J. W. Rutter, Annealing Twins and Coincidence Site Boundaries in High-Purity Lead. Trans AIME 218 (1960) 1023. 28) K.T. Aust, Annealing Twins and Coincidence Site Boundaries in Zone- Rafined Aluminum, Trans. AIME, 221 (1961) 758. 29) K. T. Aust and J. W. Rutter, Effect of Grain Boundary Mobility and Energy on Preferred Orientation in Annealed High Purity Lead, Trans. AIME, 224,(1962) 111. 30) K. T. Aust, G. Ferran and G. Cizeron, Influence de la purete du cuivre sur les relations d'orientation entre cristaux de recristallization et cristaux stries de solidification, C.R. Acad. Sc., Paris, 257 (1963) 3595. 31) D. C. Larson and Bruce Chalmbers, Recrystallization of Single Crystals of Aluminum, Trans. AIME, 230 (1964) 908. 32) Y. C. Lin, Orientations of Grains Growing into Strained Single Crystals of a Copper -1.0 Atomic % Phosphorus Alloy, Trans. AIME, 230 (1964) 1513. 33) K. Kucke and G. Ibe, P.ecrystallization of Single Crystals, Technical Rep. to ARO(D) from Institute fur Allgemeine Metallphisik der . Technischen Hochschulen, Aachen, Sept. 1964. .... . . 34) C. D. Graham, Jr. and R. W. Cahn, Grain Growth Rates and Orientation Relationships in the Recrystallization of Aluminum Single Crystals, Trans. AIME, 200 (1956) 517. 35) K. T. Aust and J. W. Rutter, Grain Boundary Migration in High Purity Lead and Dilute Lead-Tin Alloys, Trans. AIME, 215 (1959) 119. 36) K. T. Aust and J. W. Rutter, Temperature Dependence of Grain Migra- tion in High Purity Lead Containing Small Additions of Tin, Trans. AIME, 215 (1959) 820. 37) J. W. Rutter and K. T. Aust, Kinetics of Grain Boundary Migration in High Purity Lead Containing very small Additions of Silver and Gold, Trans. AIME, 218 (1960) 682. 38) C. Frois and O. Dimitrov, Existence of Preferred Orientations in Zone- Refined Aluminum Containing Small Additions of Foreign Elements, Compt. Rend., 252 (1961) 1465. 39) B. B. Rath and Paul Gordon, Single Boundary Migration in Deformed Crystals of Aluminum, Technical Rep. to ARO(D) and ONR from Illinois Institute of Technology, October 1962. 40) C. G. Dunn and P. K. Koh, Primary Recrystallization Textures in Cold Rolled Si-Fe Crystals, Trans. AIME, 212 (1958). 41) J. W. Rutter and K. T. Aust, Migration of (100) Tilt Boundaries in High Purity Lead, Acta Met., 13 (1965) 181. 42) Paul Gordon and R. A. Vandermeer, The Mechanism of Boundary Migra- tion in Recrystallization, Trans. AIME, 224 (1962) 917. 43) Paul Gordon, A rationalization of the Data on Grain Boundary Migration in Zone-Refined Metals as Influenced by Dissolved Impurities, Trans. AIME, 227 (1963) 699. 44) G. F. Bolling and W. C. Winegard, Some Effects of impurities on Grain Growth in Zone-refined Lead, Acta. Met. 6 (1958) 288. 45) E.L. Holmes and W. C. Winegard, Effects of Lead, Bismuth, Silver and Antimony on Grain Growth in Zone-Refined Tin, J. Inst. Metals, 88 (1960) 468. 46) E. L. Holmes and W. C. Winegard, The Effect of Lead and Bismuth on Grain Growth in Zone-Refined Tin, Trans. AIME, 224 (1962) 945. 47) Paul Gordon and T. A. El-Bassyouni, The Effect of Purity on Grain Growth in Aluminum, Trans. AIME, 233 (1965) 391. " 48) J. K. MacKenzie, The Distribution of Rotation Axes in a Random Aggre- gate of Cubic Crystals, Acta. Met., 12, (1964) 223. 49) M. L. Kronberg and F. H. Wilson, Secondary Recrystallization in Copper, Trans. AIME, 185 (1949) 501. 50) D. C. Brandon, B. Ralph, S. Ranganathan and M. S. Wald, A Field-lon Microscope Study of Atomic Configuration at Grain Boundaries, Acta Met., 12 (1964) 813. 51) J. C. M. Li, Discussion to Grain Boundary Migration, by K. T. Aust and J. W. Rutter in Recovery and Recrystallization of Metals, Interscience Publishers, New York (1963) 160. 52) J. Hren, An Analysis of the Atomic Configuration of an Incoherent Twin Boundary with the Field Ion Microscope, Acta Met., 13 (1965) 479. 53) G. W. Rathenau, L'Etat Solide, Solvay Conference Report, Brussels, 1952. - - - . . . - - - - . - .- "VE in - * - : . . K LOR Gamin ' I . ir 1. . . I 54) J. W. Cahn, The Impurity-Drag Effect in Grain Boundary Motion, Acta. Met. 10 (1962) 789. 55) K. Lucke and K. Detert, A Quantitative Theory of Grain Boundary Motion and Recrystallization in Metals in the Presence of Impurities, Acta Met., 5 (1957) 628. 56) E. S. Machlin, Theory of Solute Atom Limited Grain Boundary Migration, Trans. AIME, 224 (1962) 1153. 57) J. E. Burke and D. Turnbull, Recrystallization and Grain Growth," Progress in Metal Physics, 3 (1952) 220. LIST OF REFERENCES (continued) 58. P. A. Beck, Recrystallization of lead, Trans. AIME, Vol. 137, (1940), p. 222. 59. J. 8. Smart, Jr. and A. A. Smith, Jr., Effect of Certain Fifth- Period Elements on Some Properties of High-Purity Copper, Trans . AIME, Vol. 152, (1943), p.103. 60. K. Lücke, G. Masing, and P. Nolting, Influence of Small Additions of Mn, Ni, and Zn on the Recrystallization of Super Purity Aluminum. Zeit. f. Metallk., Vol. 47, (1956), p. 64. 61. K. Detert and K. Lücke, Influence of Defined Small Amounts of Impurities on the Recrystallization of Aluminum. Brown University Report No. AFOSR-TN-103AD-82016 (1956). 62. G. F. Bolling and W. C. Winegard. Grain Growth in Zone-Refined Lead. Acta Met., Vol. 6, (1958), p. 283. 63. E. L. Holmes and W. C. Winegard, Grain Growth in Zone-Refined Tin. Acta Me:., Vol. 7 (1959), p. 411. 64. E. L. Holmes and W. C. Winegard, Effect of solute Atoms on Grain Boundary Migration in Pure Metals. Canadian J. Phys., Vol. 39, (1961), p. 1223. 65. C. Frois and 0. Dimitrov, Influence de Faibles Addition de Cuivre et de Magnesium sur la Recristallisation de l'Aluminium de Zone Fondue. Mem. Scient, Rev, Metallurg., vol. 59, (1962), p. 643. Shinin - .. Si . . .. . . . 7. 66. C. Fro18 and 0. Dimitrov, 7 Colloque de Metallurgie, Ecrouissage, Restauration, Recristallisation, Presses Universitaires de France, Paris, France, (1963), p. 181. 67. D. Turnbull, Theory of Grain Boundary Migration Rates, Trans. AIME, Vol. 191, (1951), p. 661. 68. P. A. Beck, P. R. Sperry and H. Hu, The Orientation Dependence of the Rate of Grain Boundary Migration. J. Appl. Phys., Vol. 21, (1950), p. 420. 69. E. L. Holmes and W. C. Winegard, Normal Grain Growth in Zone Refined High-Purity Metals, Canadian J. Phys., Vol. 37, (1959), p. 496. 70. K. Detert and G. Dressler, Rekristallisationsverhalten von Zonengeschmolzenum Nickel. Acta Met., Vol. 13, (1965), p. 845. 71. K. T. Aust and J. W. Rutter, ultra-High Purity Metals, American Society for Metals, Metals Park, Ohio, (1962), p. 115. 72. T. S. Lundy, and J. F. Murdock, Diffusion of A126 and Mn54 in Aluminum. J. Appl. Phys., Vol. 33, (1962), p. 1671. 73. R. E. Hoffman, F. W. Pikus, and R. A. Ward, Self-Diffusion in • Solid Nickel, Trang. AIME, vol. 206, (1956), p. 483. 74. B. Okkerse, Self Diffusion in Lead, Acta Met., Vol. 2, (1954), p. 551. 75. J. D. Meakin and E. Kokholm, Self Diffusion in Tin Single Crystals, Trans. Met. Soc. AIME, Vol. 218, (1960), p. 463. 76. W. Lange and D. Bergner, Messung der Kowngrenzenselbstdiffusion in Polykristallinem Zinn, Phys. Status Solidi, vol. 2, (1962), p. 1410. 77. W. R. Upthegrove and M. J. Sinnot, Grain Boundary Self-Diffusion of Nickel, Trans. ASM, Vol. 50, (1958), p. 1031. . 78. M. Wintenberger, Elimination des lacunes dans les Aluminiums Treis Purs, Acta Met., Vol. 1, (1959), p. 549. ttg - pen 79. W. DeSorbo and D. Turnbull, Kinetics of Vacancy Motion in High- Purity Aluminum, Phys. Rev., Vol. 115, (1959), p. 560. 80. D. Schumacher, W. Schule, and A. Seeger, Untersuchung Atomater Fehlstellen in Verformten.. and Abgeschrecktem, Zeit, Naturforsch, Vol. 17a, (1962), p. 228. 81.. A. C. Damask and G. Ji Dienes, Point Defects in Metals, Gordon and Breach Science Publishers, New York, 1963, p. 260. 82. J. D. Murphy, Interdiffusion in Dilute Al-Cu Solid Solutions, Acta Met., Vol. 9, (1961), p. 563. 83. P. Gordon and T. A. El-Bassyouni-Private Communication. 84. J. H. Westbrook and K. T. Aust, Solute Hardening at Interfaces in High Purity lead - I Grain and Twin Boundaries, Acta Met., Vol. 11, (1963), p. 1151. 1.: anamuotit.ro tination intrino woo 22 FIGURE TITLES (continued) Fig 15. Schematic variation of log migration rate with temperature for constant driving force and composition. Fig. 16. Schematic variation of log migration rate versus reciprocal of absolute temperature for several values of composition. 17. Experimental data for recrystallization in aluminum and in copper in aluminum alloys. From Gordon and Vandermeer (42). Fig. 18. Apparent activation energy for boundary migration as a function of atomic fraction of copper in zone-refined aluminum. From Gordon and Vandermeer (42). Fig. 19. Activation energy versus concentration for grain growth in zone-refined tin with lead added. From Holmes and Wine gard (64). Fig. 20. Schematic variation of log migration rate with composition for constant driving force and temperature. 21. Variation of the rate of growth of grains during recrystallization as a function of concentration for copper and magnesium in aluminum. From Frois and Dimitrov (65). Fig. 22. Experimental variation of log migration rate with log composition for copper in aluminum. From Gordon and Vandermeer (42). Fig. 23. Logarithmic plot of rate of grain boundary migration at 300°C versus tin concentration for "random" grain boundaries in zone- refined lead. From Aust and Rutter (35). Fig. 24. Schematic 11lustration of migration rate-temperature curves in breakaway region for (a) T3 «T, and (b) Tom TcFrom Gordon (43). - - - - - Fig. 25. The interaction energy profiles E(x) and diffusivity profiles D(x) used in calculating the impurity drag parameters a and Appendix A. rate Fig. 26. Variation of boundary migration with composition for various driving forces 11lustrating the three types of transition behavior expected with the Cahn theory. Fig. 27. (a) Schematic plot of the growth of recrystallized grains as a function of time for compositions in the two-velocity region of Cahn's theory. (0) Variation of boundary migration with composition at constant temperature and for a high driving force. Fig. 28. Experimental plot of grain boundary mobility versus reciprocal of temperature for zone-refined aluminum. Open circles - Aust and Rutter (71). Open triangles - Frois and Dimitrov (65). Filled triangles - Gordon and Vandermeer. (42). Filled circles – Gordon and El Bassyonni (83). rate Fig. 29. Experimental variation of reciprocal of migration rate with concentration at several temperatures for copper in zone- . . . . . . . . . . . . ' ', .1 2 1...' . 1 . 4 refined aluminum. Fig. 30. Teroperature dependence of the impurity drag parameters A, B, and als for boundary migration in aluminum-copper alloys. Fig. 31. Log migration velocity versus log concentration at 139°C. . Theoretical curves calculated from Eq. 7 for various B. Data points from Gordon and Vandermeer (42). Fig. 32. Variation of log migration velocity with reciprocal of absolute temperature for three alloys of copper in zone-refined aluminum. Theoretical curves calculated from Cahn theory. Data points from Gordon and Vandermeer (42). 29 Fig. 33. Grain growth rates at three temperatures for various alloys of lead in zone-refined tin. Data taken from Holmes and Winegard (46). Fig. 34. Experimental variation of reciprocal of grain growth rate squared with concentration at several temperatures for lead in tin. Data taken from Holmes and Winegard (46). *. Y : . . - . 1 . . . . . . Livrh منانم مخمد Figure 1. Movement of a 2° boundary in a zinc single crystal under uniaxial stress. The boundary is the vertical change in shading; the irregular horizontal line is a small step in the sample surface which serves as a reference line. Top: original position of boundary. Middle: boundary caused to move 0.1 mm to right. Bottom: boundary caused to move 0.4 mm to left by stress reversal. From reference 9. . . . . * * * * แต่เมตตา Lountar air-load Figunr 2 Displacement rate of edge dislocation bound. aries as a function of the boundary angle under the action of a constant shear stress of 9.19 psi at 350'C. Ref. 11. * . can . . . so ? . * :. 125 A Figure 3. Field-ion microscope picture of a 27° grain : boundary in tungsten. Arrons indicate position of boundary. From reference 13. . . ':.'.'. --- loo olok 010 100 j'111. A. Orientation relation Intworn artifirinlly marlontoed grains and cloformiral mntrix in anung wirre the matrix Wein oxtended 20 por cent. mori the murdention wahrnusel by cutting only, Thin nntrix is plotted in Atiiniloral projection. Orientation of thin 111) poles of now crymnis churnc. ferized by 40° LIPID rotniona. + Orientation of the (111) polrk of now crgintola chiaror. forizcil by 400 (111) rotation.. Jelenl ait untion of (111) polos nitor clockwino 40° 11111 rotntion. A Helmond mituntion of (111) polna nitor countnr.clockwimu 40'1111) rotation. 13 Belonil nituntion of tho (111) polea nftor coumtor.clock wino 4029 [111rointion. From reference 26. 1 . - - .. . . . OTI O • bool OTI (n) %one-rcfined Icad with tin additions less than 0.0001 WI pet: preferred grains with 0 = 25 to 60 deg about axes shown (0), and "next-to prcſerred" grains with = 16 to 55 deg about axes shown (). e ooi (6) Zone-refined lead with tin additions from 0.0005 to 0.004 wt pct: preferred grains with a primarily 30 to 50 deg about axos within 12 deg of <111), and 26 to 28 deg about axes within 6 deg of < 100>. Figure 5. Rotation axes in lead and lead-tin alloys. From reference 35. on 0 0 Olie Figuro 6. Composite stereographic plot of Aust and Rutter data for "random" axes of rotation which relate each now grain with matrix crystal by smallest amount. Note regions of low axes density enclosed by arcs near principle poles, Open circles - zone-refined lead, ref.35; horizontal-line circles zone-refined lead, ref. 29; vertical-line circles - silver alloys, and crossed-circles - gold alloys in zone-refined lead, ref. 37. Q . RATE OF GRAIN BOUNDARY MIGRATION mm PER MIN. AT 300°C 000 لللللللللللاهما 0.0001 .002 boundaries and "special" grain boundaries. Ref. 36. at 300°C (logarithmio) vo wt pot Sn for "random" grain Fig. 7. A comparison of rate of grain boundary migration WEIGHT PEACENT OF TH "RANDOM GRAN BOUNDARIES 20-21 (1000) (36-42(III) AT Sie suc h e n . . ' - .4 . -.- - .- . . war .. . . . . , ' mimi na Simoni **.ch .. . .-.W! ewtatim , *. '* ..Wita m ine r ii RANDOM HEASURED ACTIVATION ENERGY, 0, (KCAL PER GM ATOMI BOUNDARIES لیلا 0.0025 Q0001 0.0005 0.0010 0.0019 0.0020 TIN CONCENTRATION (WT. PCT) Figure 8. Apparent activation energies for boundary migration versus tin content in tin-in-lead alloys. Reference 36. (010) POLES (010) POLES GRAIN BN GRAIN 8 GRAIN A u _ VI o I . div BOUNDARY BOUNDARY Fig. 9. Symmetrical high angle coincidence tilt boundaries in simple square lattice. Open-circle atoms form co- incidence sublattice common to both grains. Left: Two grains related by 370 rotation around (001) (the direction perpendicular to page) forming high density coincidence site boundary with l-in-5 caincidence site density. Note minimal disturbance of lattices at boundary. Right: Two grains related by rotation around 2001) of a few degress less than 37º. Coincidence sublattice dislocation at boundary superimposes substructure on boundary. Note increased boundary width and porosity, particularily at dislocation core, and creation of coherency strain well into both grains. -. - - . . . . . . . . . - - . . . . 1 is - ..... . _ Normalized porosity Function, fo Normalized concentration Function, fc Concentration - - 11. ? Functions • . • Porosity Functions 11 . . : Mobilities iii 11 . LOV-density coincidence-site , or "random 1. :: : - . - 10 x 10 TOTii. Cliiliiiii KLUFFEL C33.CO. coco 1 .: Medium-density B-coinoidence-site • boundary: .. 1 - om od nadensity 1 . - coincidence- site boundary. . Normalized Mobi. 1 1 1 . . . 1 D O... 0 FC VÔ. Schematic" plots of boundary mobilities and mobility functions versus the orientation function 40. The normalized mobility is taken equal to the product of a normalized porosity function to f1(P)/17(p). and a normalized concentration function to: 12(1/08)/1,(1/ololsee text). A9 expresses the angles by which the relativo grain and the boundary orientations differ from those of ideal coincidence relationships. ACTIVATION ENERGY, O (K. CAL PER GM ATOM} ORIENTATION DIFFERENCE, I, IDEGS) <100> TILT Figure 11. Apparent activation energies for boundary migration, tilt boundaries in zone-refined lead. Reference 41. 1000 0 0 300°C RATE OF GPAIN BOUTDART MIGRATION (en bes kind 2007 A LOWER DRIVING (MERGY IN SPECIMEN 200°C ORIENTATION DIFFCRENCE 8 Idegsl (100TALT Fig. 12. Migration rates of tilt boundaries in zone-refined lead. Ref. 41. . RO to ippotton nr777787907 poder By 0 C : • • ن ومعه مرررررره نفر در «م ه 2 مره اعداد را ب . . • - - - . .. 1 1 :: mbwe mantras que ofereiwei aowl sexoj 1410 Louz 4012041727 zudleddet Sioso 30cm . - - . . 1 . Z.N.Pb Bicrystals . . I 1 . . 32154!9933 2009 . 2014 . 11 . . . - .. .! . 1 1 expeint) . I . I s expl ii 1 11 . - . . . . .... . ' . 1 1 2012 1 C i. 1 1 ... . 1.1.1.... 1 1 . : I 1 - - 1 1 . - - - I. 1 . . . : -•.-. . . . . . KEUFFEL & ESSER CO. ri: 15. X 23 C ka ? 10X10 TO THE CENT::-? Eto silvio W I CAC 11 - - -. - - : T -- Concentration: Functions 1 Normalized concentration . . _'Uotiunt 1 0 1 . . . . . . . ..... 1 • 1 1 . Or 1 • E - - 11 l . 1 . 0 .:.:. I 11 . . - - . ..! - :.:. . - . - . . . . . . . . . . . :. . . . . - . . . . Nornalized Mobility 1 . 1 :.:.:- w . i 0 10 :-::- 11 000 ::... - . . .. . . . . . 1 . . : 11 11 . . . - .. • . . • . . . :: ::::::: :: : :::: bil: 70Schematic plots of concentration functions mobilities versus tAoror-conditions of .: boundary-impurity interaction energy. J::::::::::::::..:::.:1:1::;.:::.:li :13:7:1:::::::::1:::::::: 1 + . - 1 1 UIT - 1 . - 10 1 . . 1 . • 11 - • : 1 1 * . I . . 11 . ! ORNL-DWG 65-10137 SCHEMATIC · IMPURITY INDEPENDENT In VELOCITY TRANSITION REGION IMPURITY CONTROLLED RECIPROCAL TEMPERATURE ORNL-LR-DWG 64003R SCHEMATIC Cq>C2>C3>C4>C5 > Co>C7 MEASUREABLE RANGE OF G G, LOGARITHM OF BOUNDARY MIGRATION RATE 66 4000/T (OK) . 104 Al + Cu 40% Doform. Alloy Fig. 3-(a) Schematic variation of log G V8 roviprocal absolute temperature for : several values of Cor and (b) .semilog oxporimental plots of G va reciprooal absolute temperature for all the com- positions investigated. 6 67 IT 6, CM. PER SEC. 3. 0.0256 4. 0.0124 6. 0.0069 6. 0.0034 7. 0.0017 6. 0.00043 9. 0.00021 JO. PURE --- - 20 2.2 24 26 32 28 3.0 IC0O/T.K. 3.4 3.6 3.8 kis. .. .- . . RES .. . - zvokim ... n on in com. . .. * ORNL-DWG 65-10258 Cu IN AL Al + Cu 40 % DEFORMATION E+QL APPARENT ACTIVATION ENERGY FOR BOUNDARY MIGRATION (kcal/mole) somet GE (ESTIMATED) > 10-7 . 10-6 10-5 . 10-4 ATOMIC FRACTION OF COPPER akanan Anthramane au acan arm. - - Measured Activation Energy kcal/gatom -. -*-- . . 0.05 0.10 . Concentration, Atomic % Lead . F1o. 2. Graph of activation enersy plotted against concentration for grain growth in zone.refined tin with lead added.. . . . . . --. . . plu , ORNL-DWG 65–40138 SCHEMATIC IMPURITY INDEPENDENT TRANSITION REGION en VELOCITY IMPURITY CONTROLLED V SLOPE=-4. Fig 2o . en CONCENTRATION Velemyo) 104 . 105 . . . · 10*8] f 30°C 100 10 0 Colom.106 Colom.100 Fig. I. - Variation de la vitesse de croissance des germes de recris- lallisation en fonction de la concentration alomique en cuivre ou en magnésium de l'aluminium. Figal TN ? ¢ o q o Actual Data Points A Extrapolated and . Interpolated, . -4Al+Cu 40% Deformed a average . 900 LOG G, CM. PER SEC. Temperature 7 8 9 . slope:-) NO یااللللليا -7 -6 -5 -4 -3 LOG ATOMIC FRACTION Cu. . Fig. 7-(a) Schematlo and (b) experimental variation of log G with log composition for constant driving force and tem- perature. WN AWAY erinevate ...." i ..... ...... eport w R -.6-........ : .:: mingi : . indiniai .. i lmesinin V zeb.it t ir bi EXPERIMENTAL THEORETICAL (LUOKE AND QETERT RATE OF GRAIN BOUNDARY MIGRATION ·mm PER MIN. AT 300°C 0015 .0001 0000! .000 .00 . .O WT. POT. 106 106 10 WT. CONCENTN. TIN CONCENTRATION Fig. 4-Logarithmic plot of rate of grain boundary migra- tion at 300°C vs un concentration for "random" grain boundaries in zone-refined leade Fig 23 (6) Tox Tc away rogion for (a) T, « Tc and tion rate-temperature curves in break Fig. 1-Schematic illustration of migra- Behavior Breakaway VT U Bohavior Broakaway Тух Т. Fig zu .. .it . . ... . .ei . c . a .... ..- - -' : ' L A sa descondonos . .'-'. .. -r. - -. .- ---- UT: . - . . .. . .. p . * . ORNL-DWG 65-10154 0(X) -CONST OSX58 Olx) Elx) -Eoſ Elx)= 50/8x-Eo DGB D(x) = DGB Os x 58m 01x) = DL *28/m ģ 01x)=DGB a Os x Sol.. Dlx) * 8 DL r8/2+3/4 40 8 *2814 Ol) log DGB 0(X) = DG8 exp0f60 q=-RTIE, en Place log D(x) log QL 8/4 & Dooh 01x) = DGB Osxs 814 Elx) 0(x)=OL io -EO 8/4 SX58 . Dlx} E(X)=-E. Osx38/ Elx)=4/3 E0 /&x=4/3E0 8145*58 OLT - 810 --.: . . . uri..... ..... .: : Figas v o , W . *. *;.. .-. .- .'- . - ... - - - - -. ORNL-DWG 65-10141 HIGH DRIVING FORCE - INTERMEDIATE DRIVING FORCE . .. LOW DRIVING FORCE 0.014 Fig 26 10-1 100 101 .. 10² 103 104 . . . . ORNL-DWG 65–10436 MACROSCOPIC VELOCITY G - Ò, AVERAGE DIAMETER OF LARGEST RECRYSTALLIZED GRAIN To MICROSCOPIC VELOCITY GA -MICROSCOPIC VELOCITY G2 SCHEMATIC 097 t, ANNEALING TIME ORNL-DWG 65-10140 100 HIGH VELOCITY REGION GB E TRANSITION + 011 Mit time. Asismis- LOW VELOCITY NREGION 0.01 - 104 10 .. og . in with the iis . . ORNL-DWG 65-10139 log † (com/otocie) 2100017 (OK) in Am s !: *venimente i ngredients in the media vinde inte minden rendin vantagem M i nimalistinitati .... daras n itaria IP- RECIPROCAL OF VELOCITY (cm / sec )** (x106) N w A 125°C 50 100 CONCENTRATION OF COPPER ( atomic fraction ) 150 170°C 200 189°C - 250 (x10-6) ... 155°C 139°C .... ORNL-DWG 65-10150 . ORNL-DWG 65-10135 log DRAG PARAMETERS a AND B Q/22 - log DRAG PARAMETER *182 1 2.4 2.8 3.2 3.6 4.0 4000/ (°K) Fig 30 . . . . WY ***, London Dia ORNL-DWG 65-10134 B=102 i B=103 'B=104 B=105 log VELOCITY. (cm/sec) . 10-9 10-8 10-7 . 10-6 40-5 10-4 10-3 CONCENTRATION (atomic fraction), 10-2 ORNL-DWG 65-10142 log VELOCITY (cm/sec) ZONE REFINED AL A4.3 ppm Cu - 68 ppm cu . 1.4 1.8 2.2 2.6 3.0 3.4 3.8 zabis 1000/700K) ... .. . . . . . .. to .. . 13 V r . .- Y ORNL-DWG 65-10143 -2.5 log GRAIN GROWTH RATE (cm/sec) 227 °C 188 °C 134 °C -4.5 10-5 bania 33 40-4 10-3 CONCENTRATION OF LEAD (atomic fraction) L hawa *Mieren -- ORNL-DWG 65-10144 188 °C (x109) RECIPROCAL OF GRAIN GROWTH RATE SQUARED (cm/sec)-2 0.4 0.3 34 °C 0.2 0.1 227 °C . 2 . 4 . 6 8. 10 (x1024) CONCENTRATION OF LEAD (atomic fraction) 5 4 . . . . A . A . Skr . . . END . DATE FILMED 12/ 7 /65