You are on page 1of 10

Acta mater.

48 (2000) 30713080
www.elsevier.com/locate/actamat

ABNORMAL GRAIN GROWTH AND GRAIN BOUNDARY


FACETING IN A MODEL Ni-BASE SUPERALLOY
S. B. LEE 1{, D. Y. YOON 1 and M. F. HENRY 2
1

Department of Materials Science and Engineering, Korea Advanced Institute of Science and
Technology, Taejon 305-701, South Korea and 2General Electric Corporate Research and Development
Center, One Research Circle, Niskayuna, NY 12309, USA
(Received 3 January 2000; accepted 23 April 2000)
AbstractNormal or abnormal grain growth in a model Ni-base superalloy is observed to depend on the
grain boundary structure when heat-treated in a solid solution temperature range above the solvus temperature (11508C) of the g ' phase. When heat-treated at 12008C abnormal grain growth occurs and most
of the grain boundaries are observed to be faceted by optical microscopy, transmission electron microscopy, and scanning electron microscopy at the intergranular fracture surface. Some of the grain boundary facet planes are expected to be singular corresponding to the cusps in the polar plot of the boundary
energy against the inclination angle, and it is proposed that if these boundary segments move by a boundary step mechanism, the abnormal grain growth can occur. When heat-treated at 13008C normal grain
growth occurs, the grain boundaries are defaceted, and hence atomically rough. Normal growth is expected
if the migration rate of the rough grain boundaries increases linearly with the driving force arising from
the grain size dierence. The correlation between the grain boundary structural transition and the growth
behavior thus appears to be general in pure metals and solid solution alloys. 7 2000 Acta Metallurgica
Inc. Published by Elsevier Science Ltd. All rights reserved.
Keywords: Nickel alloy; Abnormal grain growth; Grain boundary faceting

1. INTRODUCTION

At temperatures close to 0 K all crystals in equilibrium with a surrounding vapor or liquid are predicted [17] and have indeed been observed [811]
to be polyhedral with atomically at singular surface planes. At high temperatures each singular surface plane can undergo roughening transition as
predicted initially by Burton et al. [12]. At the
roughening temperature TR the planar surface is
predicted to become curved [13] and at high temperatures some grains even become spherical [14
19]. The curved surface shape thus manifests the
rough atomic structures which have been conrmed
by diraction methods [20, 21]. The growth behavior of a single crystal from melt or solution was
found to depend critically on the surface structure
[2224]. It was also proposed that both the coarsening of many grains dispersed in a liquid matrix [25,
26] and the growth of catalytic metal particles on a
substrate [27, 28] also depended critically on their
shape and hence on their surface structure.
In the early 1970s Hart [29, 30] proposed a

{ To whom all correspondence should be addressed.

rst order structural transition of a grain boundary treating it as a two-dimensional phase. At


about the same time Gleiter [31] also proposed a
grain boundary phase transition based on the observation of apparently discontinuous change of
the energies of two grain boundaries in Pb at
about 0.73Tm and 0.76Tm, where Tm is the melting
point. (These were probably defaceting transitions
as pointed out by Cahn [32].) Simpson et al. [33]
also suggested that the observed slope changes of
the log of the grain boundary migration rates
against the reciprocal temperature were due to
grain boundary transitions. Later (in 1986)
Rottman [34] predicted the roughening transition of
low angle grain boundaries by making an analogy
with a stepped surface. Shvindlerman and Straumal
[35] showed a collection of numerous results in various metals and non-metals which indicated the
roughening transitions of grain boundaries with coincidence site lattice (CSL) orientations or near
CSL orientations at temperatures between about
0.6Tm and 0.9Tm. Recently, Westmacott and
Dahmen [36] observed that a small aluminum grain
embedded in another large one with a large misorientation angle had a polyhedral shape at low temperatures. At a high temperature the at symmetric

1359-6454/00/$20.00 7 2000 Acta Metallurgica Inc. Published by Elsevier Science Ltd. All rights reserved.
PII: S 1 3 5 9 - 6 4 5 4 ( 0 0 ) 0 0 1 1 9 - 1

3072

LEE et al.: ABNORMAL GRAIN GROWTH

grain boundary segments became rounded indicating the roughening transition. It was shown to be
reversible during temperature cycling. Although the
rounding of the planar grain boundaries in the
equilibrium shape is a direct indication of the
roughening transition as for the crystal surface, no
direct in situ observation of a rough grain boundary
at atomic scale has yet been made.
In polycrystals the grain boundary planes do
not normally correspond to those which appear
in the equilibrium shape. If the grain boundary
energy varies strongly with the grain boundary
normal (or the inclination angle) or particularly
if there are cusps in the polar plot of the grain
boundary energy s against the normal direction,
the grain boundaries can be faceted [1, 2, 32]
with zigzag shapes. The faceting of both general
and special grain boundaries have indeed been
observed in a number of metals [3740] and oxides [4143]. Such impurities and additives as O
in Ni [37, 38], Bi in Cu [4447], Te in Fe [48,
49], and CaO or SiO2 in Al2O3 [50] have been
observed to induce the grain boundary faceting. At
high temperatures some of the grain boundary facet
planes are expected to be rough corresponding to
the edges of the curved boundary segments in the
equilibrium shape [1, 2, 13, 32], and with temperature increase the defaceting transition proceeds with
an increase of the rough facet plane area. This is a
rst order phase transition unlike the roughening
transition of a singular grain boundary which can
be a higher order transition if the analogy to the
surface roughening transition holds [5153]. If the
defaceting proceeds with temperature increase until
the average boundary orientation becomes equal to
an orientation of a curved boundary segment in the
equilibrium shape, the defaceting transition is complete with a rough structure for the entire boundary
and a macroscopically curved shape. Thus in defaceting transition the singular boundary segments

Fig. 1. The initial optical microstructure of the specimen


before the heat treatments.

which coexist with the rough boundary segments


transform into the rough boundaries. As pointed
out by Cahn [32], this is equivalent to the dissolution of an intermetallic compound into a solid solution. The defaceting transition arises from the
roughening of the edges and corners of an equilibrium shape, while the roughening transition
usually refers to the change of a singular boundary
corresponding to a cusp in the s-plot and hence its
blunting.
As pointed out earlier, the grain boundary transition observed by Gleiter [31] in Pb was probably a
defaceting transition and many of the observations
collected by Shvindlerman and Straumal [35] may
also have been defaceting rather than roughening
transitions. The rst direct and deliberate observations of the defaceting transition were made by
Hsieh and Ballu [39] for asymmetric tilt grain
boundaries with CSL orientations in Al and Au.
The defacetingfaceting transitions were observed
to be reversible, and the defaceting was complete in
Al at 0.54Tm and in Au at 0.96Tm [39].
The grain boundary properties are expected and
indeed have been observed to depend on the boundary structural transition as cited by Shvindlerman
and Straumal [35]. We have recently observed [37]
that the grain coarsening behavior in pure polycrystalline Ni varied with the boundary structural
change. At high temperatures close to the melting
point in a carburizing atmosphere, the grain boundaries were defaceted with curved shapes and hence
had an atomically rough structure. Then normal
grain growth was observed. However, at low temperatures the grain boundaries were faceted and
abnormal grain growth (AGG) occurred. In a low
vacuum some or all of the grain boundaries were
faceted at all temperatures tested and AGG
occurred. Such a correlation between the grain
boundary faceting and AGG was also observed in
pure Ag when heat-treated at dierent temperatures
in either oxygen or vacuum [54]. It has been
suggested [25, 37] that AGG occurs with faceted
grain boundaries because they move by a boundary
step mechanism which has been proposed by
Gleiter [55, 56].
The purpose of this work is to test the correlation
between the grain boundary faceting and AGG in a
model Ni-base superalloy with most of the metallic
elements (Co, Al, Ti, Cr, and Mo) that are found in
the commercial alloys but without any C. Like in
commercial Ni-base superalloys coherent g '-precipitates form in this model alloy when heat-treated at
temperatures below the solvus temperature of
11508C [57]. The advantage of using this alloy is
that because the grain boundaries are strongly
pinned by the g '-precipitates the initial structures
with ne grains can be obtained by heat-treating
below the solvus temperature. The heat treatments
for grain growth were performed at 1200 and

LEE et al.: ABNORMAL GRAIN GROWTH

13008C above the g ' solvus temperature where the


alloy existed as a single phase solid solution.
2. EXPERIMENTAL PROCEDURE

The alloy of Ni24Co4Al4Ti5Cr5Mo (by


wt%) was made by spray forming. The ingots were
hot isostatically pressed at 10508C for 6 h under a
pressure of 20 ksi and specimens of 7 mm in diameter and 12 mm in length were cut out from the
same region of the ingot. The specimens were heattreated by rapidly pushing to the center zone of a
tube furnace preheated to either 1200 or 13008C.
After holding for various periods ranging from
5 min to 12 h at these temperatures in a owing Ar
atmosphere, the specimens were quenched in water.
In order to examine the grain boundary morphology at the fracture surface, notched impact
specimens of 1  1  5:5 cm3 were made and heattreated under the same conditions. These were fractured with a Charpy impact tester after immersing
in a liquid nitrogen bath. The fracture surface was
examined under a scanning electron microscope
(SEM).

3073

3. RESULTS AND DISCUSSION

Before the heat treatments (but after the hot isostatic pressing) the grains were fairly uniform in
size ranging from about 100 to 200 mm as shown in
Fig. 1. Because the hot isostatic treatment was performed at 10508C, which is below the solvus temperature (11508C) of the g ' phase, the grain growth
was probably limited by the g '-precipitates. After
heat-treating for 5 min at 12008C (above the g ' solvus temperature) large grains appeared surrounded
by small grains which had approximately the same
size as those in the initial state. Thus a typical
AGG structure was observed as shown in Fig. 2(a).
After 10 min at the same temperature more large
grains appeared as shown in Fig. 2(b), and after 2 h
most of the small grains had disappeared and the
large grains had impinged upon each other as
shown in Fig. 2(c). The equivalent sphere diameters
of the grains were measured by an image analyzer
and their size distributions for the specimens in
Fig. 2 are shown in Fig. 3. For the specimens heattreated for 5 and 10 min about 150200 grains were
measured, but for the specimens heat-treated for

Fig. 2. The optical microstructures after heat treatments at 12008C for (a) 5 min, (b) 10 min, and (c)
2 h.

3074

LEE et al.: ABNORMAL GRAIN GROWTH

2 h, their grain sizes were so large that only about


50 grains could be measured. After heat-treating for
5 min at 13008C, there was very little growth of the
grains as shown in Fig. 4(a), and after 10 min, there
appeared to be a slight growth as shown in
Fig. 4(b). There was a substantial growth after 2 h
as shown in Fig. 4(c). These micrographs and the
measured grain size distributions of Fig. 5 show
that the grain growth at 13008C was normal.
Comparisons of the micrographs (Figs 2 and 4) and
the grain size distributions (Figs 3 and 5) show that
the largest grains after heat-treating at 12008C were
larger than the largest grains after heat-treating at
13008C for the same periods.
The specimens heat-treated at 12008C showed

some jagged grain boundaries which could be seen


even under an optical microscope as shown in
Fig. 6(a), but the specimens heat-treated at 13008C
showed only curved grain boundaries as shown in
Fig. 6(b). The grain boundary morphology was
more clearly revealed at the fracture surface of the
specimens which were broken by impact as shown
in Figs 7 and 8. The fracture appeared to be predominantly intergranular and the specimens heat-treated at 12008C showed striations with steps at most
of the grain boundaries as shown in Fig. 7(a). The
faceted morphology of the grain boundaries was
more clearly visible at higher magnications as
shown in Fig. 7(b). Previously, faceted grain boundaries were also observed at the intergranular frac-

Fig. 3. The measured distributions of the equivalent sphere diameters of the grains in the specimens
heat-treated at 12008C for (a) 5 min, (b) 10 min, and (c) 2 h.

LEE et al.: ABNORMAL GRAIN GROWTH

ture surfaces of Ni [38] and Fe with Te [48, 49]. In


contrast, the specimens of this alloy heat-treated at
13008C showed smoothly curved grain boundaries
without any faceting as exhibited in Fig. 8. These
observations show that the defaceting transition
occurred for most of the grain boundaries at temperatures between 1200 and 13008C. Some of the
grain boundary facet planes observed after heattreating at 12008C are expected to be singular with
the local minimum boundary energies corresponding to the cusps in the s-plot. The defaceted grain
boundaries at 13008C are expected to have an
atomically rough structure.
These grain boundary shapes were conrmed by
observations under transmission electron microscopy (TEM). Out of the six grain boundaries in
the specimen heat-treated for 10 min at 12008C
[Fig. 2(b)] examined under TEM, four grain boundaries clearly showed faceted grain boundaries as
exhibited in Fig. 9(a). Five grain boundaries examined in the specimen heat-treated at 13008C for
10 min, as exhibited in Fig. 9(b), did not show any
faceting but smoothly curved ne bumps which are
also visible at the fracture surfaces shown in

3075

Fig. 8(b). These bumps are the g '-precipitates which


nucleated at the grain boundaries during quenching
after the heat treatment. The g '-precipitates of sizes
slightly smaller than 0.1 mm, which are visible in the
grains in both Figs 9(a) and (b), are spherical
because their lattice parameter is almost equal to
that of the matrix g phase [57]. These bumps were
absent at the grain boundaries heat-treated at
12008C as can be seen in Figs 9(a) and 7(b). It thus
appears that the preferential nucleation of the g 'precipitates at the grain boundaries occurred only
in the specimen heat-treated at 13008C. The rough
grain boundaries at 13008C will tend to undergo
faceting transitions during rapid cooling (by
quenching) at temperatures close to 12008C (which
is higher than the g ' solvus temperature), but the
transition is apparently slow enough to retain the
rough structure until the g '-precipitates begin to
form below the g ' solvus temperature. Although it
is likely that the precipitation of a second phase at
grain boundaries will depend on the grain boundary
structure, denitive conclusions cannot be drawn
until further studies are made.
The observations in this alloy thus show that

Fig. 4. The optical microstructures after heat treatments at 13008C for (a) 5 min, (b) 10 min, and (c)
2 h.

3076

LEE et al.: ABNORMAL GRAIN GROWTH

Fig. 5. The measured distributions of the equivalent sphere diameters of the grains in the specimens
heat-treated at 13008C for (a) 5 min, (b) 10 min, and (c) 2 h.

when the grain boundaries are faceted AGG occurs


and when defaceted normal growth occurs as
shown previously in Ni [37] and Ag [54]. Gleiter
[31] also noted that AGG occurred at temperatures
below the transition temperatures of the two grain
boundaries in Pb. A similar correlation between the
interface atomic structure and the coarsening behavior exists also for the grains dispersed in liquid
matrix [25, 26]. Only when the grains are polyhedral
with at singular surfaces AGG occurs [2628, 58,
59]. When the grains are spherical with a rough surface normal growth occurs and is controlled by diffusion in the liquid matrix [1719]. It is well known
that the growth behavior of crystals from melt or
solution depends on the crystal surface structure

[12, 2224]. If it is atomically rough, the growth is


continuous with the rate linearly increasing with the
supersaturation. For a system of many spherical
grains in liquid matrix, normal diusion controlled
growth was predicted [60, 61] and veried experimentally [1719].
If the surface is atomically at, the crystals which
are free of defects grow by two-dimensional nucleation as proposed by Burton et al. [12]. Then the
growth rate is predicted to be very low at low
supersaturation and increase abruptly when a
threshold supersaturation is exceeded as for the
nucleation process in three dimensions. Such crystal
growth by two-dimensional nucleation was veried
experimentally, for example, for 4He [62] and Ga

LEE et al.: ABNORMAL GRAIN GROWTH

[63] crystals. For a system of many polyhedral


grains dispersed in a liquid matrix, the driving force
for their coarsening arising from the size dierence
is usually assumed to be proportional to 1=r 1=r,
where r is the size of a particular grain and r is
roughly the average size which has to be determined
by the ux balance condition among all the grains.
As proposed by Park et al. [26], those grains which
are slightly larger than r will grow very slowly
because they are under low driving forces and only
those which are large enough to exceed the critical
driving force for rapid two-dimensional nucleation
will grow at substantial rates, thus resulting in
AGG. Wynblatt [28] observed abnormal growth of
faceted Pt particles deposited on alumina substrate
in oxygen atmosphere and proposed the same
growth mechanism. Wynblatt and Gjostein [27]
obtained numerical solutions for the two-dimensional nucleation mechanism which exhibited AGG
in agreement with the observations.
If there are dislocations the steps at the surface
can grow in spiral form and the growth rate is predicted to vary parabolically with the driving force

Fig. 6. The typical optical micrographs of grain boundaries in the specimens heat-treated for 10 min at (a)
12008C and (b) 13008C. The arrow in (a) indicates a
faceted grain boundary.

3077

for a single screw dislocation emerging vertically at


the surface [12]. At low driving forces the growth
rate will be much higher than that expected by twodimensional nucleation but still substantially lower
than that for a rough surface. At high driving
forces the two-dimensional nucleation can be the
dominant mechanism for growth and the growth
rate can approach that for a thermodynamically
rough surface by kinetic roughening [64, 65].
Because the growth rate will be still nonlinear
against the driving force, AGG can probably occur
even with dislocations which must be present in the
grains under most of the experimental conditions.
Gleiter [55, 56] proposed that even in single
phase polycrystalline solids the grain growth
occurred by the same step growth mechanism based
on his observations of apparently spiral growth of
grain boundary steps in an Al alloy [55]. Such a
mechanism is likely to operate with singular grain
boundaries which are implicit in his model. Thus if
the faceted grain boundaries move by the step
mechanism either at dislocations or by two-dimensional nucleation, AGG can occur. If the grain
boundaries are rough, their migration rate will be
determined by the atom jump rate across the grain
boundary, which will increase linearly with the driving force. Then normal growth is expected as indeed

Fig. 7. The intergranular fracture surface of the specimen


heat-treated at 12008C for 2 h at (a) a low and (b) a high
magnication.

3078

LEE et al.: ABNORMAL GRAIN GROWTH

Recently, strong evidence for the step growth of


faceted grain boundaries was observed by Lee et al.
[69] in BaTiO3 with excess TiO2. They found that
when heat-treated in air, the grain boundaries were
faceted and only those grains with double twins
grew to large sizes elongated in the direction of the
twins. They attributed this behavior to the reentrant edges produced at the junctions of the
double twins with the faceted grain boundaries.
When heat-treated in H2 the grain boundaries
became defaceted (and therefore rough), and normal growth occurred in all directions without any
preference to the directions of the double twins.
These observations are only consistent with the conclusion that the double twins inuence the step
growth at the faceted grain boundaries and become
ineective at the rough grain boundaries.
4. CONCLUSIONS

Fig. 8. The intergranular fracture surface of the specimen


heat-treated at 13008C for 2 h at (a) a low and (b) a high
magnication.

observed in Ni [37], Ag [54], and the Ni-based


superalloy in this work. The analysis of Thompson
et al. [66] and the simulation of Srolovitz et al. [67]
also predicted normal growth of grains with isotropic grain boundary energy. On the other hand, the
simulation results of Park [68] showed AGG with
faceted grain boundaries.

It now appears to be quite certain that grain


boundaries can become rough at high temperatures.
The singular grain boundaries with the normal
directions corresponding to the cusps in the s-plots
can undergo the roughening transition and the
faceted grain boundaries can also become rough by
the defaceting transition. The correlation between
the faceted grain boundaries and AGG, and
between the rough grain boundaries and normal
growth appears to be general in pure metals and
single phase alloys. But the step growth mechanism
proposed for AGG with faceted grain boundaries
obviously needs to be claried possibly by more in
situ observations of the growth behavior under
TEM and quantitative determination of the dependence of the migration rate on the driving force. In
particular, if the faceted grain boundaries consist of
both singular and rough segments, their migration
behavior is yet to be understood, although it is

Fig. 9. The TEM micrographs of the grain boundaries in the specimens heat-treated for 10 min at (a)
12008C and (b) 13008C.

LEE et al.: ABNORMAL GRAIN GROWTH

possible that the migration rate is controlled by the


singular segments. If the step growth hypothesis is
valid, the AGG behavior is expected to depend critically on the heat treatment temperature because of
the variation of the step free energy. Also dislocations at grain boundaries produced by small deformations are expected to inuence the growth
behavior. Our observations appear to be qualitatively consistent with these expectations.
AcknowledgementsThis work was supported by the
Korea Science and Engineering Foundation through the
Center for Interface Science and Engineering of Materials,
by the General Electric Corporate Research and
Development Center, USA, and by the Creative Research
Initiative Center for Microstructure Science of Materials.
The authors are grateful to an anonymous reviewer for
drawing their attention to the publications of Wynblatt
and Gjotsten [27, 28].

REFERENCES
1. Herring, C., Phys. Rev., 1951, 82, 87.
2. Herring, C., in Structure and Properties of Solid
Surfaces, ed. R. Gomer and C. S. Smith. University
of Chicago Press, New York, 1953, p. 5.
3. Fisher, D. S. and Weeks, J. D., Phys. Rev. Lett.,
1983, 50, 1077.
4. Drechsler, M. and Nicholas, J. F., J. Phys. Chem.
Solids, 1967, 28, 2609.
5. Fradkin, E., Phys. Rev., 1983, B28, 5338.
6. Rottman, C., Phys. Rev., 1984, B29, 328.
7. Rottman, C., Phys. Rep., 1984, 103, 59.
8. Balibar, S., Edwards, D. O. and Laroche, C., Phys.
Rev. Lett., 1979, 42, 782.
9. Landau, J., Lipson, S. G., Maattanen, L. M., Balfour,
L. S. and Edwards, D. O., Phys. Rev. Lett., 1980, 45,
31.
10. Avron, J. E., Balfour, L. S., Kuper, C. G., Landau,
J., Lipson, S. G. and Schulman, L. S., Phys. Rev.
Lett., 1980, 45, 814.
11. Keshishev, K. O., Parshin, A. Ya. and Babkin, A. B.,
Soviet Phys. JETP, 1981, 53, 362.
12. Burton, W. K., Cabrera, N. and Frank, F. C., Phil.
Trans. R. Soc. Lond., 1951, A243, 299.
13. Jayaprakash, C., Saam, W. F. and Teitel, S., Phys.
Rev. Lett., 1983, 50, 2017.
14. Warren, R., J. Mater. Sci., 1972, 7, 1434.
15. Warren, R. and Waldon, M. B., Powder Metall.,
1972, 15, 166.
16. Sarian, S. and Weart, W. H., Trans. metall. Soc.
A.I.M.E., 1965, 233, 1990.
17. Kang, T. K. and Yoon, D. Y., Metall. Trans., 1978,
9A, 433.
18. Kang, S. S. and Yoon, D. N., Metall. Trans., 1982,
13A, 1405.
19. Lee, W. H., Baik, Y. J. and Yoon, D. Y., Acta
metall., 1993, 41, 1263.
20. Engel, T., in Chemistry and Physics of Solid Surfaces,
Vol. 7, ed. R. Vanselow and R. Howe. SpringerVerlag, Berlin, 1988, p. 407.
21. Conrad, E. H., Prog. Surf. Sci., 1992, 39, 65.
22. Jackson, K. A., in Growth and Perfection of Crystals,
ed. R. H. Doremus, B. W. Roberts and D. Turnbull.
Wiley, New York, 1958, p. 319.
23. Leamy, H. J., Gilmer, G. H. and Jackson, K. A., in
Surface Physics of Materials, Vol. 1, ed. J. M.
Blakely. Academic Press, New York, 1975, p. 121.
24. Gilmer, G. H. and Jackson, K. A., in Crystal Growth

25.

26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.

37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.

3079

and Materials. North-Holland, Amsterdam, 1977, p.


79.
Yoon, D. Y., Park, C. W. and Koo, J. B., in Proc.
Int. Workshop on Ceramic Interfaces: Properties and
Applications IV, ed. S. J. L. Kang and H. I. Yoo. The
Institute of Materials, London, 1999, in press.
Park, Y. J., Hwang, N. M. and Yoon, D. Y., Metall.
Mater. Trans., 1996, 27A, 2809.
Wynblatt, P. and Gjostein, N. A., Acta metall., 1976,
24, 1165.
Wynblatt, P., Acta metall., 1976, 24, 1175.
Hart, E. W., in Ultrane-Grain Metals, ed. J. J. Burke
and V. Weiss. Syracuse University Press, Syracuse,
NY, 1970, p. 255.
Hart, E. W., in The Nature and Behavior of Grain
Boundaries, ed. H. Hu. Plenum Press, New York,
1972, p. 155.
Gleiter, H., Z. Metallk., 1970, 61, 282.
Cahn, J., J. Physique Colloq., 1982, 43(C6), 199.
Simpson, C. J., Aust, K. T. and Winegard, W. C.,
Metall. Trans., 1971, 2, 987.
Rottman, C., Phys. Rev. Lett., 1986, 57, 735.
Shvindlerman, L. S. and Straumal, B. B., Acta metall.,
1985, 33, 1735.
Westmacott, K. H. and Dahmen, U., in Interface:
Structure and Properties, ed. S. Ranganathan, C. S.
Pande, B. B. Rath and D. A. Smith. Trans. Tech.
Publications, Switzerland, 1993, p. 133.
Lee, S. B., Hwang, N. M., Yoon, D. Y. and Henry,
M. F., Metall. Mater. Trans. A, 2000, 31A, 985.
Henry, G., Plateau, J., Wache, X., Gerber, M., Behar,
I. and Crussard, C., Mem. scient. Revue Metall., 1959,
56, 417.
Hsieh, T. E. and Ballu, R. W., Acta metall., 1989,
37, 2133.
Goodhew, P. J., Tan, T. Y. and Ballu, R. W., Acta
metall., 1978, 26, 557.
Vaudin, D., Ruhle, M. and Sass, S. L., Acta metall.,
1983, 31, 1109.
Barber, J. and Tighe, N. J., Phil. Mag., 1966, 14, 531.
Carter, C. B., Kolstedt, D. J. and Sass, S. L., J. Am.
Ceram. Soc., 1980, 63, 623.
Ference, T. G. and Ballu, R. W., Scripta metall.,
1988, 22, 1929.
Donald, A. M. and Brown, L. M., Acta metall., 1979,
27, 59.
Donald, A. M., Phil. Mag., 1976, 34, 1185.
Menyhard, M., Blum, B. and McMahon Jr, C. J.,
Acta metall., 1989, 37, 549.
Rellick, J. R., McMahon Jr, C. J., Marcus, H. L. and
Palmberg, P. W., Metall. Trans., 1971, 2, 1492.
Pichard, C., Rieu, J. and Goux, C., Mem. scient.
Revue Metall., 1973, 70, 13.
Park, C. W. and Yoon, D. Y., J. Am. Ceram. Soc., in
press.
Kosterlitz, J. M. and Thouless, D. J., J. Phys., 1973,
C6, 1181.
Chui, S. T. and Weeks, J. D., Phys. Rev., 1976, B14,
4978.
van Beijeren, H., Phys. Rev. Lett., 1977, 38, 993.
Koo, J. B. and Yoon, D. Y., Metall. Mater. Trans. A,
in press.
Gleiter, H., Acta metall., 1969, 17, 565.
Gleiter, H., Acta metall., 1969, 17, 853.
Henry, M. F., Yoo, Y. S., Yoon, D. Y. and Choi, J.,
Metall. Trans., 1993, 24A, 1733.
Jun, J. Y., M.S. thesis, Seoul National University,
Korea, 1993.
Choi, J. H., M.S. thesis, Korea Advanced Institute of
Science and Technology, Korea, 1997.
Lifshitz, I. M. and Slyozov, V. V., J. Phys. Chem.
Solids, 1961, 19, 35.
Wagner, C., Z. Elektrochem., 1961, 65, 581.

3080

LEE et al.: ABNORMAL GRAIN GROWTH

62. Wolf, P. E., Gallet, F., Balibar, S., Rolley, E. and


Nozieres, P., J. Physique, 1985, 46, 1987.
63. Peteves, S. D. and Abbaschian, R., Metall. Trans.,
1991, 22A, 1259.
64. Weeks, J. D. and Gilmer, G. H., in Advances in
Chemical Physics, Vol. 40, ed. I. Prigogine and S. A.
Rice. John Wiley, New York, 1979, p. 157.
65. Liu, X.-Y., Bennema, P. and van der Eerden, J. P.,
Nature, 1992, 356, 778.

66. Thompson, C. V., Frost, H. J. and Spaepen, F., Acta


metall., 1987, 35, 887.
67. Srolovitz, D. J., Grest, G. S. and Anderson, M. P.,
Acta metall., 1985, 33, 2233.
68. Park, C. W., Ph.D. thesis, Korea Advanced Institute
of Science and Technology, Korea, 2000.
69. Lee, B. K., Chung, S. Y. and Kang, S. J. L., Acta
mater., 2000, 48, 1575.

You might also like