THERMODYNAMICS OF GRAIN BOUNDARIES IN Ni-RICH NiAl E. Rabkin, L. Klinger, V.Semenov* Department of Materials Engineering, TECHNION-Israel Institute of Technology, 32000 Haifa, Israel *Institute of Solid State Physics, Russian Academy of Sciences, 142432 Chernogolovka, Russia Abstract Thermal grooving at grain boundaries in Ni-rich NiAl was studied by atomic force microscopy technique. The determined average ratio of grain boundary to surface energy for large-angle grain boundaries at 1400 C is 0.45, which is in a good agreement with the results of computer simulations. It has been found that in most cases thermal grooving at the grain boundaries is accompanied by relative shift of the adjacent grains. This shift is associated with the grain boundary sliding caused by the relaxation of internal substructure of the specimen. A model of grain boundary grooving with the simultaneous sliding is developed. The calculated grain boundary groove profiles are in a good agreement with the experimentally measured ones. 1. Introduction Good mechanical properties, low density, high melting temperature and high oxidation resistance of the ordered NiAl intermetallic compound with B2 structure attract attention to this material for more than three decades. The main obstacle for structural applications of NiAl is its severe intrinsic grain boundary (GB) brittleness at low temperatures. Therefore, the knowledge of GB atomic structure and energy is important for understanding the mechanism of intergranular failure in NiAl. While the atomistic simulation studies addressed both the structural and energetic aspects of GBs in NiAl [1-3], only the atomic structure of some GBs in NiAl was studied experimentally using high resolution electron microscopy [3, 4], whereas the data on GB energies are virtually nonexistent. The GB energy, b, is an important parameter directly related to the cohesive 198 strength, G, of the polycrystal, which in the case of ideally brittle intergranular fracture is given as G s1 s 2 b (1) where s1 and s2 - the energies of two surfaces formed as a result of GB fracture. Indirectly, the GB energy is also connected with GB diffusivity and intergranular corrosion resistance. The main goal of the present paper was experimental determination of GB energies in polycrystalline NiAl. Simultaneously with the energy measurements, three of five possible macroscopic geometrical degrees of freedom (DOFs) for the same set of GBs, namely, those associated with misorientation parameters of the adjacent grains, were identified. Finding the geometrical DOFs of low-energy GBs serves as a starting point in the concept of GB engineering [5]. Recently, a high potential of atomic force microscopy (AFM) for determining relative GB energies in metals [6] and ceramics [7] has been demonstrated. AFM combines the possibility to scan relatively large surface areas with the atomic resolution in the vertical direction, thus allowing determination of dihedral angle at the root of GB groove, , with a very high accuracy unattainable by other methods [6]. Under the assumption of surface isotropy, the relative GB energy is directly connected with by the relationship b 2 cos s 2 (2) Relative GB energy values determined by eq. (2) contain however a potential error caused by the neglection of the torque terms which stem from the inclinational dependence of s. Nevertheless, the values obtained in the study of GB grooving in ceramics [7] were consistent with the results of previous studies. Since the anisotropy of surface energy in metals and alloys is, in general, lower than in ceramics, the torque terms will be neglected in the present study, too. An additional important source of error in determining GB energy is GB sliding that occurs in the process of annealing and leads to a difference in the level of adjacent grains near the GB. The formation of GB steps caused by such a sliding have been recently revealed in Fe(Si) alloy [8]. 199 The driving force for this process is the dependence of b on the misorientation angle, whereas the selective absorption of lattice dislocations by GB has been suggested as a possible mechanism of GB sliding [8]. Earlier, a similar phenomenon of near-GB lattice rotations in Ni has been revealed by tracking local lattice orientation across GB with the aid of electron backscattering diffraction (EBSD) technique [9]. In the present study we have also found that during high-temperature annealing of NiAl polycrystal the GB grooving process occurs simultaneously with the GB sliding, with the amount of sliding varying from zero to approximately 1 m. If GB grooving is accompanied by GB sliding, the relationship between the groove width and its depth derived by Mullins [10] and used in [7] for determining is no longer valid. We suggested an algorithm for determining the GB energy value which is not affected by the process of GB sliding and also modified original Mullins’ theory of GB grooving by taking into account GB sliding process. 2. Experimental A rod of NiAl intermetallic compound with nominal composition 48 at.% Ni + 52 at.% Al was produced from Ni of 99.95 at.% purity and Al of 99.99 at. % purity by vacuum casting. The as-cast alloy was then remelted and purified by vacuum electron-beam floating zone melting. After two passes of the melted zone through the rod at the velocity 6 mm/min we obtained a polycrystal with averaged grain size 2 mm. The discs of 3 mm in thickness cut from the rod by spark erosion were annealed in vacuum of 10-5 Pa at 1400oC for 1 h to provide relaxation of internal stresses and stabilize microstructure. Metallographic examination of one of the discs in longitudinal section revealed columnar microstructure with GBs running approximately along the rod axis. Another disc, ground and polished to the mirror quality on SiC paper followed by diamond paste down to 0.3 m particle size was annealed for the second time at the same conditions. Its surface was then studied by the light microscopy (LM), scanning electron microscopy (SEM), with the attachment for electron backscattering diffraction, and AFM. Chemical composition of the specimen after the second annealing was determined by the energy dispersive X-ray analysis using JSM840 (JEOL, Japan) scanning electron microscope equipped with LINK ISIS (Oxford Instruments, England) energy 200 dispersive spectrometer (EDS). EDS measurements were carried out at 20 kV accelerating voltage and 1 nA probe current in Ni K and Al K radiation with 30° take-off angle using pure Ni and Al as standards. The standard deviation of the measured intensity for a single measurement with the acquisition time 100 s did not exceed 2 %. The result, averaged over 10 measurements and normalized to 100 %, gives Al content as 43.60.5 at.%. The partial loss of Al in comparison with the nominal composition (52 at.%) is due to its evaporation during the remelting process. The AFM measurements were done using the Autoprobe CP AFM of Park Scientific operated in the contact mode and the Si Ultralevers with nominal tip radius of 10 nm. Both the topography signal and the feedback loop error signal were collected. For the acquisition of AFM data, the AFM cantilever was placed near the chosen GB using the on-axis optical microscope of the AFM and then its tip was brought into contact with the specimen surface. Using the AFM operating system, it was possible, by observing the distance between the cantilever tip and GB, to adjust the slow scan direction to be parallel to the GB groove with the accuracy of 2. After the adjustment, AFM image containing 256 x 256 pixels was obtained by scanning in the direction perpendicular to the GB groove, with the GB being approximately in the middle of the scanned region. The effect of the finite size of AFM tip on the measured topography of GB grooves has been analysed in details in the previous publications [6, 7] but for the relatively large grooves (4-6 m wide) observed in the present work this effect can be neglected. Saylor and Rohrer determined the dihedral angle at the groove root by measuring the width (taken as the distance between the two maxima) and depth of the GB groove [7], while Schöllhammer et al. [6] used the direct fit of the observed topography by the functions suggested by Mullins [10]. However, both methods cannot be applied in the present work due to the observed grain sliding along GB. For that reason, the known relationship between the groove width and depth ceases to be valid and the solution of surface diffusion problem is to be modified [10, 11]. Moreover, in the AFM used in the study, scanning is performed by moving specimen mounted on the holder against the cantilever tip. The movement, implemented by piezoelectric contraction/expansion of the walls of the tube on which the specimen holder is mounted, displaces the specimen surface along the axis of the piezoelectric tube. As a result, the output AFM signal always contains some instrumental parabola-type 201 distortion which is added to the true specimen topography. This instrumental distortion, which is a common feature of many AFMs, can be removed by substracting the averaged parabola from the primary image. Some “flattening” of the primary AFM image is also required if the specimen surface is not exactly parallel to the scanning plane. Such a flattening may cause an additional error in determining the dihedral angle of GB groove from the groove width and depth. In the present research a procedure for extracting the dihedral angle of GB groove different from those employed in [6, 7] was developed. Each of the 256 line scans of the flattened image perpendicular to the GB groove was used in calculating dihedral angle. First, the point of absolute minimum was determined for every scan. Then, a number of points for interpolation was chosen along the scan line on both sides of the GB groove, with the length of the interpolated regions being about 2/3 of the distance between the minimum and each maximum of the profile. After parabolic interpolation, performed separately for each side of the groove profile, the intersection point of the two parabolas was determined by solving the corresponding quadratic equation. The example of this procedure is given in Fig. 1. 200 100 y, nm 0 1 -100 -200 -300 -400 Intersection: (8.9527; -333.43) 5 6 7 8 9 10 11 12 x, m Fig. 1. To the algorithm for determining the dihedral angles 1, 2 at the root of GB groove. The parabolic interpolation is applied separately to the left and right branches of the AFM profile. 202 The coordinates of the intersection point of two parabolas in Fig. 1 (8.9527 m; -333.43 nm) are slightly different from those of the measured minimum point (8.9444 m; -330.60 nm). The difference is due to the finite information density of the AFM image. Indeed, since the distance between neighbouring pixels is 78 nm in the above example, the location of the obtained minimum value does not necessarily corresponds to the root of GB groove. After determining the intersection point, the dihedral angles 1 and 2 (Fig. 1) were calculated from the slopes of interpolating parabolas at the intersection point. In the above example 1=82.015° and 2=74.41°. Based on these values, the relative GB energy was calculated according to equation b cos 1 cos 2 s (3) which is valid for the isotropic surface energy and immobile GB perpendicular to the specimen surface. Metallographic examination on the cross-sections showed that, as a rule, no GB migration was associated with the grooving process (no GB energy determinations were carried out for GBs which migrated on long distances due to some other reasons) and, because of the columnar microstructure, GBs were perpendicular to the specimen surface. In the above example, b/s=0.408. The values of b/s averaged for all 256 line scans typically show standard deviation as low as 0.01. The main advantage of the parabolic interpolation used in the present work is its consistency with the flattening procedure of primary AFM images, in which the linear or parabolic functions are employed. The measurements of the crystallographic orientations of the individual grains were performed by EBSD method carried out with the LINK OPAL (Oxford Instruments, England) equipment using the high resolution field emission gun SEM LEO 982 (Zeiss - Leica). Analytical conditions for EBSD measurements were as follows: accelerating voltage - 20 kV, beam current - about 3 nA, working distance - 21 mm, angle between the electron probe and the sample normal - 70°. Under these conditions the probe diameter was about 40 nm. The image of the analyzed area on the inclined surface was obtained employing the conventional secondary electron detector mounted on the side wall of the microscope chamber, or using the detector of forward scattered 203 electrons which is a part of the LINK OPAL assembly. After specific grains of interest have been selected, the microscope was switched to the spot mode for acquisition of EBSD patterns that were recorded on a CCD screen. The identification of crystal planes and directions appearing in the patterns (observed as bright bands and their intersections) was performed using LINK OPAL software. For the calibration of angular distances between bands on CCD screen we used EBSD pattern obtained from (001) Si single crystal mounted on a specimen holder close to the specimen. The error in the measurement of angular distances did not exceed 1º. The orientation matrix was determined for each grain of interest and Miller indices of the normal to the grain surface were calculated. The misorientation between two specific grains was expressed in terms of an axis common to the two crystals and the angular rotation around this axis (the angle/axis pair). 3. Results A representative AFM image of GB region (corresponding to GB No. 1 in Table 1) is shown in Fig. 2. Good reproducibility of both the groove shape and dihedral angle along the GB groove root should be mentioned. At the same time, far from the groove root the level of the left grain in Fig. 2 is visibly lower than the level of the right grain. Some relative shift of the adjacent grains was observed for the majority of studied GBs. Fig. 2. AFM image of GB No. 1. 204 Table 1 Orientations of adjacent grains, corresponding misorientation at their grain boundaries (GBs), and amount of sliding at and relative energies of GBs GB No. Surface I Surface II GB m b/s 1 (3 –1 0) (4 –1 0) 15.5<-33 –4 4> 0.13 0.390.01 2* (3 –1 0) (4 –1 0) 15.5<-33 –4 4> 0.02 0.400.01 3 (3 –1 0) (010) 14.4<2 –1 3> 1.2 4 (013) (010) 20.4<23 5 –24> 0.63 5 (810) (010) 43.0<17 2 –2> 0.5 6 (810) (701) 26.0<1 35 2> 0.14 7 (701) (801) 3.0<12 –8 9> 8 (1 -7 -1) (5 -1 0) 22.0<-17 0 1> 0.23 0.490.01 9 (5 -1 0) (061) 21.5<0 35 1> 0.0 0.410.01 10 (910) (5 -1 0) 19.2<34 8 -1> 0.15 0.490.02 11 (160) (010) 12.8<11 21 26> 0.41 12 (160) (010) 12.1<4 -5 -9> 0.32 13 (014) (010) 16.5<011> 0.53 (010) (010) 14.5<-1 -33 7> 0.12 14 0.590.02 0.110.01 0.260.01 0.330.01 *13 difference in in-plane inclination with respect to GB No.1. Table 1 summarizes the amount of sliding and calculated relative energy for all studied GBs. In the second and third columns of Table 1 Miller indexes of the surfaces of grains forming the GB are shown in such a way that the grain of higher level appears in the second column. Fig. 3a shows LM image of GB, a portion of which migrated slightly during annealing. AFM image of the area of GB migration (Fig. 3b) indicates that the region swept by migrating GB exhibits the intermediate level between those of the left and right grains. The root of the GB groove at the original GB position is blunted thus indicating that the GB left its original position in the middle of annealing process and there was enough time for smoothing the sharp groove root by the surface diffusion mechanism. The blunting of the groove root at the 205 original GB position is visualized by the line scan of the AFM image in the direction perpendicular to the GB (Fig. 3c). GB 100 m (a) 0.6 (b) 0.4 OGB y, m 0.2 0.0 -0.2 -0.4 0 (c) FGB (c) 10 20 30 40 50 x, m Figure 3. LM (a) and AFM (b) images of migrating grain boundary (GB) and its linear AFM profile (c) containing original (OGB) and final (FGB) GB positions. Note the blunted groove root at OGB position. 206 One of the possible reasons for the relative shift of adjacent grains might be the evaporation of Al from the specimen surface with the rate dependent on the surface orientation. If crystallographic orientations of adjacent grains are different, the difference in evaporation rate would result in different contraction rate of the grains leading to the formation of a step at their GB. To verify this possibility, Al distribution across the GBs exhibiting the largest amount of sliding was analyzed by EDS. LM image of the studied area is shown in Fig. 4a. White triangles that cross the GB are pointed to the lower grains, with the corresponding numbers indicating the height of the step between the level of two adjacent grains. Fig.4b contains EBSD patterns obtained from the four grains observed in Fig. 4a and used for indexing crystallographic orientation of their surfaces. Results of EDS analysis of Al distribution performed along lines aa’ and bb’ in Fig. 4a are presented in Fig. 4c. Both grains show the same Al content and no discontinuity of Al distribution across the GB is observed. (410) (010) a a’ b b’ 0.63 m (013) 0.13 1.1 m m (310) 500 m Fig. 4a. LM image of four grains, with their Miller indices indicated, which form two GB junctions; triangular marks across the boundaries point to the lower grain, with the amount of corresponding GB sliding shown near the triangles For the most right GB in Fig. 4a the AFM measurements of GB grooving were performed at two locations differing by inplane inclinations visible on the sample surface (GBs No. 1 and 2 in Table 1). As can be seen from Table 1, the amount of GB sliding in these two positions is very different. It can be therefore 207 4. Discussion 4.1. The mechanism of GB sliding The following reasons for the formation of steps at GBs during annealing can be named: - Residual shear stresses in the specimen produced during cooling after the previous annealing. This possibility is excluded due to the fact that certain amount of sliding occurs also at the final position of migrated GB (Fig. 3). Indeed, the relaxation of residual stresses would cause the GB sliding at the initial position, whereas, taking into account high annealing temperature (1400 C), all internal stresses should be fully relaxed before the GB arrives to its final position. - The presence of steps at GBs prior to annealing as a result of specimen preparation (polishing). In this case, as well as in the preceding one, no sliding is to be expected at the final position of migrating GBs. In addition, no GB steps were revealed by AFM prior to annealing, though LM allowed to distinguish separate grains because of their different reflectivity. - Selective evaporation of Al from the specimen surface during annealing. This possibility is not confirmed by the Al concentration profiles measured across GBs, which do not exhibit any discontinuity. Moreover, interdiffusion coefficient in Ni-43 at.% Al system extrapolated to 1400 C from the data of Shankar and Seigle [12] is 4.310-12 m2/s. This corresponds to 125 m wide Dt -law) after 1 h interdiffusion zone (estimated according to annealing. Therefore, if the shrinkage is caused by selective Al evaporation, the change in the grain surface level should occur gradually over a distanse exceeding 100 m because any concentration discontinuity in this range would be smoothed due to the bulk interdiffusion beneath the surface. Experimentally, however, the difference in the levels of two adjacent grains is clearly visible on distances lesser than 10 m across GBs (see Fig. 2 and 3c). In addition, the grain with (010) surface exhibits the lowest level in Fig. 4a. The (010) plane is the second densely packed plane, after (011), in the bcc lattice. Since the lower evaporation rate is expected for the densely packed planes this observation contradicts the hypothesis of 208 selective evaporation. It can be therefore concluded that selective Al evaporation cannot be the cause of the observed GB sliding. As follows from the above arguments, the change in the level of adjacent grains ought to be associated with some physical relaxation process that occurs during annealing either within the GBs or in their close vicinity. In the previous publication [8] the hypothesis of near-GB lattice rotations being responsible for the step formation at GBs in an FeSi alloy was put forward. The process is illustrated schematically in Fig. 5. GB y x Fig. 5. Illustration of the mechanism of near-GB lattice rotation. The driving force for the rotation is the decrease of GB energy. For simplicity, a low angle GB built of the array of edge dislocations is considered. Its energy is an increasing function of the misorientation angle and, therefore, there exists a thermodynamic driving force for the grain rotation which decreases the misorientation angle. Rotation of individual grains in the fine grained polycrystals is possible [13], while only the lattice regions close to GBs can change their orientation in the coarse grained structures. Such a local rotation of the lattice is made possible due to the mechanism of selective absorption of lattice dislocations at the GBs. Let us assume that, for some reason, the right grain in Fig. 5 is more active than the left one in supplying dislocations to the GB. There is an attraction force between the GB and the lattice 209 dislocations having the sign of their Burgers vector opposite to that of GB dislocations. These lattice dislocations became trapped in the GB and subsequently annihilate with the GB dislocations. As a result, the linear density of GB dislocations decreases and, hence, the misorientation angle between the adjacent grains decreases, too. Moreover, as a result of this process the right grain is left with an excess of dislocations having the same sign of Burgers vector as the GB dislocations. This excess leads to the near-GB lattice rotation that is consistent with the decrease of misorientation angle. The net macroscopic result of the dislocation rearrangement of this type is the formation of a step at the intersection of GB with the specimen surface (Fig. 5). The suggested mechanism is consistent with our experimental observations. Among the investigated GBs, GB No. 9 has the lowest angular deviation (1.3 around <1 –9.4 5> axis) from the special coincident site lattice (CSL) 13 boundary (26.62<010>), where stands for the reciprocal density of coincident sites in two misoriented lattices, and also exhibits the lowest change of level among all studied large angle GBs. This is consistent with the known fact that both the rate of GB sliding under the action of the applied shear stress and the rate of grain rotation due to the misorientation dependent GB energy tend to a minimum for the singular CSL GBs [14, 15]. We do not consider here the origin of the bulk dislocations responsible for the near-GB lattice rotation. This may be either growth dislocation inherited from the re-melting process or dislocations introduced into sub-surface layers in the process of mechanical polishing, or both. As has been recently demonstrated [16], a significant dislocations density is produced up to the depth of about 100 m below the surface during the final polishing stage of NiAl single crystal using the 0.3 m Al2O3. An interesting observation was made by J.W. Cahn and coworkers during their molecular dynamics simulations of the migration of a GB separating a small grain embedded in a larger one [17]. The shrinkage of the small grain was accompanied by the increase of the GB misorientation angle due to the increase in the density of GB 210 dislocations. This example clearly demonstrates that while the microstructure evolution is driven by the decrease of the total interfacial energy its actual direction can be determined by the decrease of either GB energy or GB area. In our case, the potential for the latter is limited because of the columnar microstructure, therefore, the former option is realized. Moreover, the idealized computer simulations do not take into account the defect structure of the bulk (dislocations). 4.2 GB energy Among GBs listed in Table 1 there are three low-angle GBs (No. 7, 12 and 14) with the misorintation angles of 3°, 12.1° and 14.5° having the relative energies of 0.11, 0.26 and 0.33, respectively. The averaged value of relative GB energy for the large angle GBs (those with the misorientation angle above 15°) is 0.45. It can be, therefore, concluded that the energy of low angle GBs increases with the misorintation angle and is considerably lower than that of large angle GBs. This result is in line with the current understanding of GB energetics [14]. It should be also noted that the determined value of relative GB energy in NiIAl (0.45) is closer to that in pure metals (0.30.4) [14] rather than in ceramics (1.0-1.5) [7]. As known from the computer simulation studies [1-2, 18], the GB energy is sensitive to the microscopic position of GB plane with respect to the two kinds of atomic sites. In the ordered B2 structure of NiAl, even of stoichiometric composition, Ni-rich, Al-rich or stoichiometric compositions are possible in the GB plane. For example, the energy of Al-rich GB is considerably higher than that of Ni-rich GB in Ni-rich alloy [18]. Based on these results, one would expect a high scattering of the measured GB energies, since the microscopic positions of the GB planes should be a random result of the specimen pre-history. However, the experimental results of the present work do not support the hypothesis of structural multiplicity of GBs in NiAl. Indeed, the value of GB energy remained constant with a good accuracy along the whole GB segment of AFM scans, the size of which varied from 12 to 30 m. 211 Moreover, two segments of the same GB separated by the distance of approximately 100 m and having different in-plane inclinations exhibited almost identical energies (GBs No. 1 and 2 in Table 1). It is, therefore, very probable that in the studied structure the majority of GBs belonged to the same class, presumably, of the Ni-rich GBs. We took an advantage of the fact that the preparation procedure of our sample resulted in a noticeable <001> texture and plotted the value of relative GB energy as a function of misorientation angle for GBs with their misorientation axis deviating from the <001> axis no more than by 15° (see Fig. 6). (GB No. 9 is not included in Fig. 6 because, according to the low value of GB sliding associated with the GB, it belongs to a special class of CSL boundaries characterized by a sharp local cusp in GB energy vs. misorientation dependence.) 0.8 b/ s 0.6 0.4 0.2 0.0 0 20 40 60 80 Misorientation around <001>, deg Fig. 6. The dependence of measured relative GB energies on misorientation angle for the GBs with the misorientation axis close to <001>. The solid line represents the fit by equation (4). D. Wolf demonstrated that formal substitution of the misorientation angle in the Read-Shockley formula for the energy of low angle GBs by the sin(2) gives an empirical relationship which fits very well the calculated energies of <001> tilt GBs in the whole range of 212 misorientations [19]. As can be seen in Fig. 6, there is a good fit between our experimental data and an equation of this type b 0 sin 2 A ln sin 2 (4) for 0=0.08515 and A=8.27435. According to equation (4) the maximum relative energy in the family of <001> tilt GBs is about 0.7. There are two sets of reliable computer simulation data (both for 0 K) for comparison with our experimental results. Yan et al. [1] calculated the energies of some symmetrical tilt <001> GBs in stoichiometric NiAl using the semi-empirical embedded atom model (EAM) interatomic potential that reproduces correctly different type of defects accommodating the deviations of composition from the stoichiometry. The calculated relative energies for the large angle GBs vary from 0.6 to 0.92. This is considerably higher than our experimental values. Mutasa and Farkas calculated the energies of 5 tilt <001> GBs in NiAl alloys of different composition [18]. The relative energy of non-singular Nirich (013)<001> GB in the Ni-rich alloy was approximately 0.5, which is comparable with the our measurements. It can be therefore concluded that the modified EAM potential for NiAl (that reproduces correctly not only the type of defects but also the diffusion parameters of pure Ni and Al), as developed by Mishin and Farkas [20] and used in Ref. [18], gives the best description of GB energetics in Ni-rich NiAl. 5. The model Let us assume that the GB stays planar during the annealing process. We will try now to find a solution for the surface topography evolution supposing that the surface diffusion mechanism is operative and under the same set of simplifications as those in the original Mullins’ paper [10]. Our basic assumption is that the change of the vertical coordinate of the surface, y (see Fig. 7), is a result of two independent processes, namely, the shape evolution by surface diffusion and the GB sliding, i.e. independent movement of two adjacent grains in the opposite directions by the distance h(t) 213 y 4 y h( t ) B t t x 4 (5) where y+(x,t) and y-(x,t) describe the solutions for the right and left grains, respectively, and h(0)=0 (6a) Far from the GB the curvature of the surface is zero and no surface diffusion occurs. Therefore, y , t ht (6b) In equation (5) B is the coefficient of Mullins: B s Ds kT (7) where , Ds and are the thickness of surface diffusion layer, surface diffusion coefficient and atomic volume, respectively, and kT has its usual meaning. S S 2h b y x GB Fig. 7. To the model of GB grooving with simultaneous GB sliding. 214 It should be noted that equations (5)-(7) are derived in the framework of original Mullins’ assumptions which are valid for the single-component system [10]. It is very likely that the surface excess energy in NiAl is very sensitive to the deviations of the local surface composition from the equilibrium value, which is the case for the GB energy [1,2, 18]. In this situation the equations (5)-(7) still can be used, but the surface diffusion coefficient should be properly renormalized [21]. The equations (5)-(7) should be completed by the following boundary conditions at the root of GB groove x=0 y 0 y 0 y x x 0 2 y x 2 3 y x 3 y x x 0 x 0 x 0 2 y x 2 3 y x 3 b s (8a) (8b) (8c) x 0 (8d) x 0 Equation (8a) represents the obvious continuity condition; equation (8b) is the small-slope approximation of the equilibrium condition (3); equations (8c) and (8d) represent the conditions of continuity of the chemical potential and the surface diffusion flux, respectively. The system of equations (5) - (8) describes the GB grooving process with simultaneous GB sliding, the amount of GB sliding, 2h(t), being the function of time. The solution of equations (5)(6) is sensitive to the function h(t). Only in one case, namely, ht t1 / 4 , these equations allow the self-similar solution exhibiting the time-independent shape of GB groove. For any other non-zero h(t) the solution depends essentially on time and no “universal” solutions like that suggested by Mullins [10] exist. As follows from the 215 experimental observations, the GB sliding occurs continuosly during the annealing, however, the exact time dependence of the amount of GB sliding is unknown. To get an idea about its time dependence, we performed simple calculations (not shown here) based on the following assumptions: (i) the elastic stress field of the GB is described by the known expression for the low angle GBs; (ii) the velocity of lattice dislocations is proportional to the elastic interaction force between an individual dislocation and the GB; (iii) all lattice dislocations that have approached the GB by the distance equal to the interdislocation spacing in the GB are absorbed by this GB. The calculations resulted in the time dependence of the type h(t ) 1 exp t (9a) with =const. We will, nevertheless, assume a different type of time dependence ht Bt 1/ 4 (9b) keeping in mind that this time dependence is chosen only for the mathematical convenience and only the qualitative comparison of the calculated and experimentally measured groove shapes is possible. Here =const is the only new parameter in the Mullins-type model of the phenomenon. It should be noted that dependence (9b) captures the main features of more realistic dependence (9a), namely, the rapid GB sliding at the beginning of annealing process and the decreasing sliding rate at the later stages of annealing. Let us introduce the new dimensionless coordinates u and g according to x u(Bt )1 / 4 y b Bt 1 / 4 g u Bt 1 / 4 2 s 216 (10a) (10b) In the dimensionless coordinates the system of equations (5) – (8) is transformed to 1 g ug ' 4 g , t 0 g'''' g 0 g 0 g ' 0 g ' 0 2 g ' ' 0 g ' ' 0 g ' ' ' 0 g ' ' ' 0 (11a) (11b) s b (11c) (11d) (11e) (11f) where the prime, double prime, etc. denote the derivatives d , du d2 etc., respectively. Note that for 0 (no GB sliding occurs) du 2 g(u)=g(-u) and the problem (11) is reduced to the Mullins’ problem [10]. The general solution of equations (11) can be represented in the form g u a1 g1 u a0 g 0 u (12) where a1 and a0 are the u-independent constants and g1, g0 are two linearly independent solutions of equation (11a) for which g 0 . These functions satisfy the following boundary conditions in the point u=0 g1 0 1 2 5 / 4 g0 0 1 g1 ' 0 1 217 0.78012 ; (13a) 5 / 4 0.64092 2 1 g1 ' ' 0 0.61372 2 3 / 4 g 0 ' 0 (13b) g0 ' ' 0 0 (13c) g1 ' ' ' 0 0 5 / 4 g 0 ' ' ' 0 0.361602 21 / 2 (13d) where is the Euler’s gamma-function. Following Mullins [10], the functions g(u) can be represented in the form of infinite power series. The constants a1 and a0 are determined from the boundary conditions (11c-f) a1 a1 1 and a0 a0 2 s b (14) Finally, with the definition (10b) we have y u, t b Bt 1 / 4 g1 u 2 s 1 g0 u 2 s b (15) The expression in the square brackets of the RHS of equation (15) calculated for the different values of parameter is shown in Fig. 8a. In Fig. 8b, two experimentally measured line profiles of the GB groove are shown for comparison. The calculated and measured profiles are in a good qualitative agreement with each other. This agreement proves that GB sliding process occurred continuously during the specimen annealing. It should be noted that the process of GB sliding decreases the width of GB groove, defined as a distance between the two maxima on the groove profile. For example, the width of Mullins groove (=0) is 4.6, while for /m=1 the value is decreased to 3.6 (see Fig. 8a). 218 1.5 1.0 y/m(Bt)1/4 Mullins' function /m=0.5 0.5 0.0 /m=0.3 -0.5 -1.0 /m=1 (a) -10 -5 0 x/(Bt)1/4 5 10 0 5 10 15 20 200 y, nm 0 (010) (160) (410) (310) -200 200 0 -200 (b) 0 5 10 15 20 x, m Figure 8. (a) Linear profiles of GB grooves calculated according to equation (15); m denotes b/2s. (b) Experimentally measured profiles of GBs exhibiting different amount of GB sliding. 5. Conclusions From the results of the present study the following conclusions can be drawn. 1. The topography of thermal GB grooves in the Ni-43.6 at.% Al intermetallic after annealing at 1400 C for 1 h was studied by AFM technique. It was found that in many cases the thermal GB grooving is accompanied by the sliding at GBs, the amount of sliding being below 1 m. The dependence of the relative GB 219 2. 3. 4. 5. energy and of the amount of GB sliding on misorientational degrees of freedom of the GBs was determined. The ratio of GB-to-surface energy for the low angle GBs is low (0.11 for the GB with the misorientation angle of 3) but increases with the misorientation angle, reaching an average value of 0.45 for the large angle GBs with the misorientation of approximately 15.5. The values of relative GB energy are in a good agreement with the results of computer simulation of Mutasa and Farkas [18]. It is demonstrated that the amount of GB sliding is sensitive to the in-plane inclination of the GB. It is shown that the observed GB sliding can be reasonably explained in terms of near-GB lattice rotation. The driving force for the lattice rotation is the misorientation dependence of GB energy, with the rotation itself occuring by the mechanism of selective absorption of lattice dislocations by the GB. The model of GB grooving with simultaneous GB sliding is developed in the small-slope approximation of Mullins [10]. The calculated topographies of the GB grooves are in a good qualitative agreement with the observed shapes. Acknowledgements This research was supported by the Ministry of Science and Culture of the German Federal State Niedersachsen, by the INTAS grant No. 97-0118 and by the Centre for Absorption in Science, Ministry of Immigrants Absorption, State of Israel. E.R. also wishes to thank the support of the Fund for the Promotion of Research at the Technion. Helpful discussions with Dr. Y. Fishman are heartily appreciated. 1. 2. 3. 4. 5. References Yan, M., Vitek, V., and Chen, S.P., Acta mater., 1996, 44, 4351. Mishin, Y. and Farkas, D., Phil. Mag. A, 1998, 78, 29. Fonda, R.W., Yan, M., and Luzzi, D.E., 1995, Phil. Mag. Lett., 71, 221. Nadarzinski, K. and Ernst, F., Phil. Mag. A, 1996, 74, 641. Reitz, W.E., JOM, 1998, No. 2, 39. 220 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. Schöllhammer, J., Chang, L.-S., Rabkin, E., Baretzky, B., Gust, W. and Mittemeijer, E.J., Z. Metallkd., 1999, 90, 687. Saylor, D.M. and Rohrer, G.S., J. Amer. Cer. Soc., 1999, 82, 1529. Rabkin, E., Semenov, V.N. and Bischoff, E., Z. Metallkd., 2000, 91, 165. Thomson, C.B. and Randle, V., Acta mater., 1997, 45, 4909. Mullins, W.W., J. Appl. Phys., 1957, 28, 333. Robertson, W.M., J. Appl. Phys., 1971, 42, 463. Shankar S. and Seigle, L.L., Metall. Trans. A, 1978, 9A, 1467. Harris, K.E., Singh, V.V., and King, A.H., Acta mater., 1998, 46, 2623. Sutton, A.P. and Balluffi, R.W., Interfaces in Crystalline Materials, Clarendon Press, Oxford, 1995. Martin, G., Phys. stat. sol. (b), 1992, 172, 121. Simkin, B.A. and Crimp, M.A., Phil. Mag. Lett., 2000, 80, 395. Cahn, J.W., Personal communication (1999). Mutasa, B. and Farkas, D., Metall. Mater. Trans. A, 1998, 29A, 2655. Wolf, D., Scripta metall., 1989, 23, 1713. Mishin, Y. and Farkas, D., Phil. Mag. A, 1996, 75, 169. Rabkin, E., Estrin, Y. and Gust, W., Mater. Sci. Engng. A, 1998, 249, 190. 221