Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Available online at www.sciencedirect.

com

Acta Materialia 60 (2012) 2719–2728


www.elsevier.com/locate/actamat

3-D phase-field simulation of grain growth: Topological analysis


versus mean-field approximations
Reza Darvishi Kamachali ⇑, Ingo Steinbach
Interdisciplinary Centre for Advanced Materials Simulation (ICAMS), Ruhr-University Bochum, 44801 Bochum, Germany

Received 21 December 2011; received in revised form 20 January 2012; accepted 23 January 2012
Available online 3 March 2012

Abstract

The characteristics of 3-D grain growth are investigated by a topological analysis of phase-field simulation results compared with
theoretical mean-field theories. We found that the size distribution of the grains starting from an arbitrary narrow distribution crosses
the self-similar Hillert distribution, and ends in a distribution with relatively longer tails of large grains in which the central peak shifted
towards smaller grain size. The distribution of topological classes, as characterized by the number of facets per grain, is found to be time-
invariant for the process as a whole. The obtained shape function is in good agreement with the analytical distribution function derived
based on the average N-hedron model [Rios PR, Glicksman ME. Act Mater 2008;56:1165]. The volumetric growth rate per topological
class also correlates well with the analytical approach obtained by Mullins [Mullins WW. Acta Mater 1956;3:900]. The relationship
between grain size and its shape, however, deviates from theoretical predictions.
Ó 2012 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Grain growth; Grain size distribution; Phase-field simulation; 3-D von Neumann; Grain shape distribution

1. Introduction have been applied to grain growth in different ways. Despite


the variety of the different approaches they mainly confirmed
Grain growth is a subject of long-standing interest the parabolic kinetics for both 2-D and 3-D grain growth
which despite intensive studies is still not fully understood. which had been introduced by Burke and Turnbull [8].
This is mainly due to the geometrical complexity of the An alternative description of the growth process in
individual elements which form polycrystalline materials: terms of “topological requirements of space-filling and
two-dimensional grain boundaries together with triple lines the geometrical needs of surface tension equilibrium” has
and higher-order junctions with different impacts on the been developed for 2-D grain growth by Smith, von Neu-
growth process. mann, and Mullins [9]. This analysis has proven to be very
In addition to extensive theoretical and experimental useful in predicting a stable size distribution for the grains,
developments which have attempted to capture the physical as characterized by the number of facets of neighbours per
origins of the phenomenon, computer simulation has been grain. A good overview of the existing theories on grain
increasingly applied to the subject in the last decades, as growth has been provided by Atkinson [10] at 1988. For
computational power has grown enormously. Various 3-D grain growth several analytical approaches have also
approaches, e.g. the Monte Carlo–Potts model [1], Surface been developed to estimate the rate of growth related to
Evolver [2], the front-tracking method [3], vertex dynamic the topological characteristics of the grains [11,12].
[4], cellular automata [5] and the phase-field model [6,7], Recently, MacPherson and Srolovitz [13] demonstrated a
more general solution for the 3-D von Neumann problem,
⇑ Corresponding author. Tel.: +49 2343222607. in which most of the analysis concentrates on regimes of
E-mail addresses: reza.darvishi@rub.de, r.darvishi.k@gmail.com (R. self-similarity where the size and shape distributions of
Darvishi Kamachali). grain structures can be scaled by a characteristic length.

1359-6454/$36.00 Ó 2012 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2012.01.037
2720 R. Darvishi Kamachali, I. Steinbach / Acta Materialia 60 (2012) 2719–2728

The existence of such a regime, however, requires strict Here, r is the interface tension, and r1 and r2 are the
conditions, which is discussed by Mullins [14]. local radii of curvature. The sign of the velocity is hidden
In our study, we perform large-scale 3-D multi phase- in the sign of the curvature in local coordinates, where a
field simulations of ideal grain growth, i.e. with constant negative sign corresponds to shrinking grains and a posi-
interface energy and mobility. In the phase-field method, tive sign for the growing ones. Burke and Turnbull simplify
grain boundary energy is defined by the gradient energy the case by assuming (i) the interface tension is constant
term between different domains which is built into the free and independent of the radii; (ii) the averaged radius of a
energy functional of the system. Therefore, the existing gra- grain, R, is proportional to the average of the radii of the
dient between domains directly reveals the grain boundary curvature, r 1, where r = C1R; (iii) the rate of change of
position. The Gibbs–Thomson effect, which is naturally the average radius is proportional to the interface velocity,
incorporated in the governing equations of the phase-field, v; dR/dt = C2v; and (iv) the size distribution is self-similar,
guarantees the representation of the curvature-driven i.e. it scales with a typical or average size over time R  R.
motion of the grain boundaries. However, it is not enough Employing these assumptions in Eqs. (1) and (2), the rate
for grain growth simulations, since there are considerable of change of the average radius is:
numbers of higher-order junctions which cooperate in the
dR C 2 Mr
growth process. For our propose, the multi phase-field con- ¼ ; ð3Þ
cept [15] is employed which enables us to reconstruct triple dt C1R
lines, quadruple points and higher orders of junctions in a which leads to parabolic kinetic for the growth:
physically sound picture. The validity of this aspect for the
R2  R20 ¼ K 1 t: ð4Þ
multi phase-field model has been examined in previous
2C 2 Mr
studies [16,17]. The most significant limitation for multi where K 1 ¼ is the growth constant. Eq. (4) is asserted
C1
phase-field simulations is the large amount of storage to be true for both 2-D and 3-D grain growth. It should be
which is needed during the computation. We overcome this noted that the above assumptions approximate the grain
problem in the current series of simulations by using a structure built by planar or distorted facets with the picture
dynamic storing algorithm that only reserves the available of spheroids embedded into a uniform medium. We discuss
field variables rather than all field variables. the intimate connection between these assumptions in
In this work, both statistical and topological perspec- Section 3.2.
tives of ideal grain growth are discussed. In this section,
we briefly review the theories on normal grain growth to 1.1.2. Hillert’s approach
which we compare our results. In Section 2, the multi Inspired by Greenwood’s equation for the coalescence
phase-field concept is briefly described and the parameters of particles, Hillert introduced a critical grain size above
of our simulations are given. The results and analysis are which grains grow and below which the grains shrink
discussed in Section 3. We display the role of topological [18]. He started from Eqs. (1) and (2) and introduced the
characteristics of the grain in the process of the growth. growth rate so as to be negative for small grains and posi-
We interrelate the size of the grains with their shape and tive for large ones:
present the grain shape distribution. The results are dis-  
dR 1 1
cussed against the background of analytical mean-field the- ¼ aMr  ; ð5Þ
dt Rcr R
ories and topological models.
or
1.1. Classical models for grain growth  
dR R
R ¼ aMr 1 : ð6Þ
dt Rcr
1.1.1. Burke and Turnbull analysis
The classical model for normal grain growth proposed Here, a is a constant, and the critical grain size, Rcr, is
by Burke and Turnbull [8] deduces a parabolic relationship related to the instantaneous average grain size R, which
for grain evolution. In this model it is assumed that the varies with time. From Eq. (5) small grains are expected
grain boundary migrates toward the centre of its curvature, to shrink at a high rate, and for large grains the growth rate
which in turn reduces the interfacial area as well as its asso- approaches a positive constant value which is relatively
ciated energy. The interface velocity of a single grain is small. Fig. 1 shows the growth rate in arbitrary units for
assumed to be linearly proportional to the relative pres- different grain sizes where Rcr = 5 and aMr = 1.
sure, P, between the inside and outside of the grain: Utilizing the Lifshitz–Slyozov method [19], Hillert ended
up with the following grain growth rate equation:
v ¼ MP ð1Þ
dRcr aMr
where M is constant. The pressure is proportional to the ¼ ð7Þ
dt cRcr
mean curvature:
 
1 1
P ¼r þ : ð2Þ  
r1 r2 1
The grain is projected to an equivalent sphere: 2r ¼ 1
þ r12 .
r1
R. Darvishi Kamachali, I. Steinbach / Acta Materialia 60 (2012) 2719–2728 2721

Fig. 1. (a) The growth rate of different grain sizes presented by Eq. (5) and (b) the schematic diagram of volumetric growth rate derived based on Eq. (9).

and Another apparent property of Eq. (5) is that for all


grains of the same size the growth rate will be the same,
R2cr  R2cr0 ¼ K 3 t; ð8Þ
which may be not the case in reality; one reason is that this
2aMr
where c is a constant and K 3 ¼ .
He has shown that for
c equation does not take into account the topological prop-
steady-state grain growth, c = 4 and Rcr ¼ 98 R. It is worth erties of the grains. This issue will be also addressed in
keeping in mind that the critical size in the case of coales- Section 3.2.
cence in the Lifshitz–Slyozov method has a physical mean-
ing related to the kinetics of nucleation and growth, while, 1.2. Topological illustration of grain growth
as Hillert himself mentioned, in the case of grain growth,
such a critical size is a postulate. Mullins [11] derived an expression for the average
The constant a is similar to the constant C2 in Eq. (3), normalized volumetric rate of change of a topological
resulting from the third assumption of the Burke–Turnbull class:
analysis which in fact corrects the interface velocity, pro-
1 dV
jecting a real-shaped grain into a sphere. Hillert estimated V 3 ¼ MrG1 ðnÞG2 ðnÞ ð10Þ
the parameter a = 1 for 3-D growth. dt
The rate of volumetric change of individual grains is not where G1 and G2 are given by
easy to derive directly from analytical expressions such as  13 " pffiffiffiffiffiffiffiffiffiffiffi!#
Eq. (3) or (5), because it needs a specific description for 3 p 1:86 n  1
G1 ðnÞ ¼  2 arctan ð11Þ
the shape of the grain. It is an important quantity, how- 4p 3 n2
ever, which appears frequently in the topological analysis
and
of 3-D grain growth. It also enables us to estimate the most
 1   
prominent relative sizes from a volumetric perspective. For 3 3 2 n2
a single grain of radius R: G2 ðnÞ ¼ 5:35n 3 pffiffiffiffiffiffiffiffiffiffiffi  0:375G1 ðnÞ : ð12Þ
4p 2 n1
dV dR The topological class number, n, is equal to the number
¼ SR2 ð9Þ
dt dt of faces for each grain in 3-D. Hilgenfildt et al. [12] devel-
where S is a shape factor, equal to 4p for spherical parti- oped another relationship for the normalized volume
cles. Substituting Eq. (5) into Eq. (9), we will find the cor- growth rate for 3-D bubbles very similar to the previous
responding volumetric growth rate in Hillert’s formulation. expression:
As the schematic graph shows in Fig. 1b, the volumetric
growth rate has a local minimum at 12 Rcr . 1 dV
V 3 ¼ MrGH ðnÞ ð13Þ
It has been frequently pointed out that in mean-field dt
approaches, one assumes that the variation of the size of with
a grain is independent of that of its neighbours, which is
not the case during grain growth [20]. To avoid this, Lout  1  h i  23
6 3 3 1 Dn p p
[21] ignored the mean-field assumptions and the curvature- GH ðnÞ ¼ tan 3  Dn ðn  2Þ tan ;
driven motion of the interfaces, and started with a pure 22=3 4p 2 3 xn
random walk theory which allows boundaries to move ran- ð14Þ
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
  !
domly. Pande [22] tried to combine the essentials of the p
mean-field approach and pure random-walk theory to Dn ¼ 2 arctan 4 sin2 1 ð15Þ
xn
deduce a new expression for grain growth.
2722 R. Darvishi Kamachali, I. Steinbach / Acta Materialia 60 (2012) 2719–2728

and interfacial energy is the dominant contribution to the total


  free energy:
2 Z
xn ¼ 6 1  : ð16Þ
n F ¼ f intf : ð20Þ
X
MacPherson and Srolovitz [13] derived the general solu-
tion for the growth rate for three or more dimensions The interfacial energy density of the system is given by
which in 3-D relates the growth rate with the mean width Ref. [15] as:
of the grains, K(D), and the sum over the length of all triple X N
4rab g2
lines, e: f intf ¼  2 r/a  r/b þ /a /b ; ð21Þ
  g p
1X
a¼1;a–b
dV
¼ 2pMr KðDÞ  eðDÞ : ð17Þ
dt 6 where g is the interfacial width and rab = r is the interface
energy between phases a and b taken to be equal for all
The latter is not exactly a topological relationship (like interfaces and a constant. N is the order of the interfaces
previous presentations) and measuring the model parame- or junctions, i.e. the number of grains which are in local
ters is difficult. In their paper, however, they demonstrate contact (2 for interfaces, 3 for a triple line, 4 for a quadru-
a more simplified relation: ple junction, etc.). Each domain of the grains is specified
dD pffiffiffi with a single phase variable /a which is 1 inside the domain
D ¼ C 01 Mr 6  C 02 n ð18Þ
dt and 0 in all other domains. Through the interface the field
in which C 01 and C 02 are constants, and D is the linear variable varies smoothly from 1 to 0 such that the integral
dimension of the domain which can be estimated by 4R over field variables is 1 in each point of the space.
for a sphere. This equation directly relates the radial The evaluation of the phase variables, therefore, reveals
growth rate to the topological aspect of the grain. If we the microstructural evolution of the system. This is given
compare this formulation with Eq. (6), it will be clear that within a generalized form of the Ginsburg–Landau
the grain radius in the mean-field approach is expected to equation:
scale with n1/2. This is a simple but important conclusion. XN  
_/a ¼ M dF dF
However, the range of validity for this relationship is not 
b¼1
N d/a d/b
determined. In this study, we investigate the issue of a
one-to-one relationship between the size and the shape of Mr XN

individual grains. ¼ ½I a  I b ; ð22Þ


N b¼1–a
Based on Glicksman’s model [23] for polycrystalline
structures, the so-called average N-hedra method (ANH), where
Rios and Glicksman [24] derived another relationship for p2
the growth rate which depends only on n and very similarly I c ¼ r2 /c þ /: ð23Þ
g2 c
predicts n  R2. They also obtained an approximation for
the growth rate as: Here M is the interfacial mobility between two phases.
For an ideal system, as we consider here, it is assumed that
dV 2=3 9 pffiffiffi pffiffiffiffiffiffi
 Mrð n  ncr Þ ð19Þ the interface energy and mobility are equal for all pairs of
dt 4 the grains. The interested reader can find more details of
in which ncr is the critical average N-hedra which has equiv- the phase-field model in Ref. [27].
alent zero mean curvature (zero growth rate). Based on Grain growth was simulated using our open-source
their theory, ncr = 13.3973. In their studies, Rios and package for phase-field modelling and simulations, Open-
Glicksman succeeded in directly linking between their Phase [28]. Since we are aiming to perform large scale sim-
topological model and mean-field assumptions. Based on ulations, the storage of the grain information has to be
the average N-hedron model, they have obtained Hillert’s handled by an efficient algorithm which stores only non-
affine kinetic equation and grain size distribution. More- zero field variables. There is no cut-off set for the multiple
over, they have also derived a conclusive analytical deriva- junctions. In particular, during the first stages of the simu-
tion for the corresponding grain shape distribution [25,26]. lations after random initialization there can exist junctions
In Section 3.3 the shape distribution function of grains is up to an order of 26, which is the upper limit for the cubic
presented and compared with available models. We also numerical grid.
ascertain the neutral value ncr from our simulation results. The simulations start from a narrow distribution of
grains which are small in size compared to the box size,
2. Multi phase-field modelling and simulation but sufficiently large with respect to the grid spacing Dx
and interfacial width g = 6Dx. Nevertheless, the early
The driving force for grain growth is the reduction of the stages of simulation are omitted in the analysis in order
total free energy of the system by means of decreasing total to ensure that numerical artefacts are avoided. The figures
interfacial area. For pure grain growth, we assume that the presented in the following discussion are from a 3-D simu-
R. Darvishi Kamachali, I. Steinbach / Acta Materialia 60 (2012) 2719–2728 2723

Fig. 2. The simulation box of 5123 grid points after 1000, 5000 and 10,000 time steps. The colour coding presents phase-field indices of the grains. The
visual “layering” is only due to the post-processing and the grains have a fully random texture. There are no grains with the same index. (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

lation of 30,000 initial grains in a cubic box of 512 grid cant in the simulation box. Moreover, the same behaviour
points in each direction. Fig. 2 shows three snapshots of has been observed from simulation boxes of different sizes
the simulation box at different time steps. We have chosen and initial distributions. These results show that, although
the physical dimensions for interfacial energy and mobility the grain size distribution crosses Hillert’s prediction, at
of 1 J m2 and 1014 m4 J1 s1, respectively, within the later stages of the growth the distribution function deviates
accepted range for metallic systems. The time step and grid from the mean-field prediction. A tendency to abnormal
spacing are dt = 101 s and Dx = 107 m. grain growth with a bimodal distribution, however, was
not observed. We assume that the final distribution in
3. Results and discussion our simulations fulfils the requirements for self-similarity.

3.1. Grain size distribution 3.2. Kinetics of the growth

Fig. 3a shows the evolution of the grain size distribution The time evolution of the squared mean grain size is
during the initial steps of the growth. The system evolves shown in Fig. 4. The curve is plotted over first 20,000 time
very rapidly from its initially narrow distribution and steps but note that almost the whole simulation (50,000 ts)
approaches a broader distribution. The x axis in the figure maintains this form. The initial stages of the growth are
presents the relative grain size with respect to the critical omitted.
grain size defined in Hillert’s analysis, 98 R.2 A parabolic equation is obtained by fitting to the curve
After 1500 time steps, the grain size distribution as:
obtained from the simulation matches nicely with Hillert’s
R2 ½m2  ¼ 5:1  1015 t þ 3:7  1013 : ð24Þ
prediction and tends to stay almost unchanged for a rela-
tively long time. In this period of growth, the number of The growth constant is K = AMr = 5.1  1015 m2 s1
existing grains decreases from about 16,000 to 4000. in which A  12. The constant A is defined as 2C C1
2
in the
Fig. 3b compares the grain size distributions for several Burke and Turnbull analysis. From the stability condition,
time steps with the Hillert function. The values are taken a similar constant, is calculated from the mean-field
from individual time steps without averaging. These results approximation as:
confirm the work by Kim et al. [29], who performed similar  2
8 2a
phase-field simulations. ¼ 0:395 ð25Þ
At a later stage, however, the peak of the grain size dis- 9 c
tribution gradually shifts to the left (Fig. 3c). Simulta- which is smaller than the obtained value A. It is worth not-
neously the frequency of big grains at the right-hand side ing that the constants C2 and a in the Burke–Turnbull and
increases and the distribution becomes more symmetric Hillert formulations correct the interface velocity due to
around 1 (Fig. 3d). The rate of change of the grain size dis- the shape effect.
tributions is very slow and the symmetric grain size distri- Since we are able to evaluate Eq. (6) and calculate the
bution also tends to keep its form for a relatively long time. constant a from our simulation, it is easy to test the
Statistically, the number of remaining grains is still signifi- mean-field assumptions in the case of 3-D grain growth.
Fig. 5 shows the normalized growth rate vs. the relative
2
For consistency and fair comparison between figures within this study, grain size for different time steps. Although the data points
we use the relative grain size with respect to the critical value, Rcr, instead are highly scattered, they can still be fitted along a line. The
of average grain size. Note that the Hillert function is not a homogeneous two important characteristics of this fitting are (i) the slope
function of degree 1, and therefore it is not straightforward to reformulate of the line which gives a and (ii) the relative grain radius,
his function for the relative size respect to the average grain size R. See the R
Appendix in Ref. [18]. Rcr
, for which the growth rate vanishes.
2724 R. Darvishi Kamachali, I. Steinbach / Acta Materialia 60 (2012) 2719–2728

Fig. 3. Simulated grain size distributions. (a) The grains evolve toward a smoother function and finally get close to the Hillert function. (b) For a relatively
long period of time, the grain size distribution coincides with the Hillert function. We name this period of growth the “Hillert regime”. (c) The distribution
gradually deviates from the Hillert regime and shifts to the left. (d) While the number of available grains is still considerable, a new, relatively stable
distribution appears. This deviation is found to be independent of the initial configurations.

which means that the relative velocity of the interfaces


changes with time. In fact, this violates the required condi-
tion for the self-similar regime mentioned in the previous
section. During the early stage of the simulation, where
the Hillert regime is found, the average value for a is about
1.25, which is above the assumed value, a = 1. Further-
more, it should be noted that for grains of the same size,
the dispersion of the growth rate values is significant
(Fig. 5). Therefore the individual grains could have a values
which deviate even more from 1. From the ANH model,
Rios and Glicksman [26] have calculated the aANH  0.81.
This is obtained for a specific case where the vertex-to-ver-
tex distance (quadra-junction spacing) for the entire grain’s
network was assumed to be independent of number of
neighbours.  
Fig. 4. Squared average grain size changes linearly over time, showing the
parabolic kinetics of grain growth.
The critical relative size RRcr at dR
dt
! 0 is also observed
to be slightly above 1 (Fig. 6b). The deviation is more pro-
nounced at the later stage of the growth. In fact, the
Fig. 6 shows these two parameters obtained from the asymptotic solution for the Lifshitz–Slyozov stability con-
simulations. The a value obtained from the linear approx- dition [13], based on which the maximum and critical grain
imation decreases at the later stages of the simulation, sizes are obtained, seems not to be exactly the case for the
R. Darvishi Kamachali, I. Steinbach / Acta Materialia 60 (2012) 2719–2728 2725

Fig. 5. The normalized growth rate for different time steps fitted to a linear curve. The colour coding present the topological classes of the grains. For
relatively small grains there is larger deviation from the mean-field approximation. (For interpretation of the references to colour in this figure legend, the
reader is referred to the web version of this article.)

Fig. 6. The evolution of a (a) and critical relative size (b). There is a departure from the mean-field assumptions made by Hillert.

grain growth. This perhaps explains the deviation of the Remarkably, despite all the discrepancies with the
distribution function from the mean-field prediction at mean-field theory, the parabolic kinetics of the grain
later stages, which is a natural result of the geometrical growth with a single growth constant (K) remains valid
complexity of the grains. This will be discussed in more for the entire growth process. If we restart from Eqs. (1)
detail in Section 3.3, where we focus on the topological and (2) and look again at assumptions (ii) and (iii) made
characteristics of the grains during the evolution. by Burke and Turnbull (Section 1.1.1), the property of
2726 R. Darvishi Kamachali, I. Steinbach / Acta Materialia 60 (2012) 2719–2728

parabolic equations requires that constants C1 and C2 have fastest shrinkage rate was predicted to be 0.5Rcr, which is
a fixed relation in order that the growth constant K not hard to distinguish in our simulations. The shape of the dis-
change with time. We have already mentioned that C2 var- tribution changes slightly along the simulation.
ies similarly to a. This is due to the type of correction which
Burke and Turnbull provide for the kinetic Eqs. (3) and (5). 3.3. 3-D von Neumann and topological analysis
We also know that C1 is a geometrical constant which
relates the average radius (calculated from the total volume As presented in Section 1.2, Mullins [11] and Hildefeldt
of the grain) with the average curvature of the grain. Thus, et al. [12] have given analytical expressions for 3-D grain
C1 is expected to vary inversely with C2. Of course, the rate growth. Fig. 8 shows the normalized volumetric rate of
of the changes for C1 and C2 is slow and interconnected. change from our simulation and both these analytical func-
In Fig. 7, the volumetric growth rate is plotted against tions. At the earlier stage of growth, the obtained values
the normalized radius. We have chosen the same time steps coincide with Mullins’s prediction while later on they devi-
as in Fig. 5. The shape of the distributions compares well ate slightly; the slope of the curve decreases slightly with
with the function obtained from the mean-field assumption time. The critical number of faces (ncr), which is the average
(Fig. 1b). This is also comparable with simulation results number of faces per grain at which the volumetric rate
obtained from the vertex model of Barrales Mora et al. changes its sign, is 13.5 for earlier stages. Later on, it
[30], who discussed the variety of growth rates for the increases towards 14, but it is impossible to extract a pre-
grains of the same size. They also plotted the growth rates cise value from the simulations. This value is predicted to
of individual grains vs. their size, and obtained results that be 13.769 by Lazar et al. [31] and 13.397 from the analyses
look similar to ours. From the mean-field theory, the of Rios-Glicksman [24] and Hilgenfeldt et al. [12]. Based on

Fig. 7. Volumetric rate of changes for different time steps. Compare with Fig. 1b. The shape of the distribution changes slightly with time. The colour
coding relates to the topological classes. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this
article.)
R. Darvishi Kamachali, I. Steinbach / Acta Materialia 60 (2012) 2719–2728 2727

tion is quite sharp, and quickly changes to a stable


distribution. Unlike the grain size distribution, the grain
shape distribution stays in an almost time-invariant regime
for the whole period of the simulation. Rios and Glicks-
man have derived an analytic form of the shape distribu-
tion function based on the ANH theory (Fig. 4 in Ref.
[26]), which is in good agreement with our current results.
Both the current study and the Rios–Glicksman function
correlate well with Surface Evolver data from Wakai
et al. [2]. Using the front-tracking method, Lazar et al.
[31] conducted large-scale 3-D grain growth simulations
and evaluated the shape distribution, which, like our simu-
lation result, is invariant in time. The distribution evalu-
ated from their simulations data is compared to our
results in Fig. 9.
Fig. 8. A comparison between normalized volumetric growth rate and the To explain the difference between the shape and size dis-
analytical expression. There is a good correlation between simulation
results and the Mullins prediction. Our simulation gives a value of between
tributions, we studied the relationship between the grain
13.5 and 14 for the neutral number, ncr. size and shape in more detail. We plot the topological class
of each grain vs. its relative grain size in Fig. 10a. Although
the points are scattered, more detailed analysis shows that
the scattering seems not to be uniform but that for a spe-
their Surface Evolver simulations, Wakai et al. [2] reported cific topological class, some grain sizes are more probable
that grains surrounded by more than 15 grains tend to than others. This is also reported by Feltham from experi-
grow, which is higher than the values obtained in the cur- mental observations [32]. Fig. 10b shows an example for
rent study. They also showed that at the later stage of the distribution of grains size for different topological
the growth, the neutral value slightly increases. classes.
A topological perspective on 3-D growth can be Feltham suggested a linear relationship between the size
obtained and compared in the same way as the size analy- and the topological class of the grains [32], while some
sis. The colour coding in Figs. 5 and 7 presents the topolog- other analyses found that the grain size scales with the
ical classes of the individual grains. What is salient in the square root of n [13]. Rios and Glicksman [24] explicitly
evaluation is that the growth rates of the grains with differ- derived the face-number-dependent growth rate in which
ent sizes but the same topological classes stay close pffiffiffi
the grain area in 3-D scales with n (Eq. (19)). We
together, even closer than that of the grains of the same size obtained the averaged topological class for each relative
but different topological class. Hence it seems to be useful grain size which is plotted in Fig. 10c. It is found that there
to explain the rate of growth in the topological class space is a one-to-one relationship between grain size and the
rather than the relative grain size space. Topological inves- average number of its faces, which does not change dra-
tigations usually use this kind of shape-dependent growth matically during the simulation. For grains larger than
rate [24]. 1
Rffi the curve was observed to be linear, which means
2pffifficr
The distribution of the grain shape (number of topolog- n / R. We argued in Section 1.2 that such a relationship
ical classes) is shown in Fig. 9. The initial shape distribu- is necessary for all grain sizes to fulfil the mean-field

Fig. 9. Grain shape distribution at the initial (a) and stable (b) stages of the grain growth simulation.
2728 R. Darvishi Kamachali, I. Steinbach / Acta Materialia 60 (2012) 2719–2728

Fig. 10. Topological analyses. (a) The topological class number (n) of each grain vs. its relative grain size. The black cubes are averaged values. (b) Grain
pffiffiffi
size distributions for three topological classes. (c) Square root of the averaged topological classes ( n) plotted against the relative grain size.

assumptions. For smaller grains, however, the form of rela- Acknowledgements


tionship is significantly different. This, in fact, can explain
the differences between the size and the shape distribution The authors acknowledge financial support through
functions. The origin of this behaviour is hidden in the geo- ThyssenKrupp Steel Europe AG, Bayer MaterialScience
metrical complexity of the system, which still is not fully AG, Salzgitter Mannesmann Forschung GmbH, Robert
understood. Bosch GmbH, Benteler Steel/Tube GmbH, Bayer Technol-
ogy Services GmbH and the state of North-Rhine Westpha-
lia as well as the European Commission in the framework of
4. Conclusion
the European Regional Development Fund (ERDF).
Using a multi phase-field model, we performed large-
References
scale 3-D grain growth simulations. The characteristics of
normal grain growth were investigated with respect to [1] Blikstein P, Tschiptschin AP. Mater Res 1999;2:133.
mean-field assumptions and topological properties. We [2] Wakai F, Enomoto N, Ogawa H. Acta Mater 2000;48:1297.
found that due to the geometrical complexity of a polycrys- [3] Frost HJ, Thompson CV. Curr Opin Sol Stat Mater Sci 1996;1:361.
talline system, the grain behaviour may deviate from the [4] Weygand D, Brechet Y, Lepinoux J. Philos Mag B 1998;78:329.
mean-field assumptions. The size distribution of the grains [5] Liu Y, Baudin T, Penelle R. Scrip Mater 1996;34:1679.
[6] Fan D, Chen LQ. Acta Mater 1997;45:611.
starting from an arbitrary narrow distribution crosses the [7] Krill III CE, Chen LQ. Acta Mater 2002;50:3057.
Hillert distribution to end in a distribution with longer tails [8] Burke JE, Turnbull D. Prog Metal Phys 1952;3:220.
of large grains and the central peak shifted toward smaller [9] Mullins WW. Acta Mater 1956;3:900.
grain size. The parabolic kinetics of growth, however (with [10] Atkinson HA. Acta Mater 1988;36:469.
a single growth constant), are unaffected by these devia- [11] Mullins WW. Acta Metall 1989;37:2979.
[12] Hilgenfeldt S, Kraynik AM, Koehler SA, Stone HA. Phys Rev Lett
tions. Following the arguments of the Burke–Turnbull 2001;86:2685.
analysis, we found it is necessary to have a correlation [13] MacPherson RD, Srolovitz DJ. Acta Mater 2007;446:1053.
between the effective mobility of interfaces, the local curva- [14] Mullins WW. J App Phys 1986;59:1341.
ture, and the grain size, in order to keep the parabolic rela- [15] Steinbach I, Pezzolla F. Physica D 1999;134:385.
tionship in its current form. [16] Guo W, Steinbach I. Int J Mater Res 2010;101:480.
[17] Guo W, Spatschek R, Steinbach I. Physica D 2011;240:382.
Unlike the grain size distribution, the grain shape distri- [18] Hillert M. Acta Metall 1965;13:227.
bution was almost self-similar and constant over the whole [19] Lifshitz IM, Slyozov VV. J Phys Chem Solids 1961;19:35.
process of growth. Our results are in good agreement with [20] Lucke K, Heckelmann I, Abbruzzese G. Acta Metall 1992;40:533.
analytical distribution function obtained by Rios and [21] Louat NP, Duesbery MS. Phil Mag A 1994;69:841.
Glicksman [26], Surface Evolver results obtained by Wakai [22] Pande CS. Acta Metall 1987;35:2671.
[23] Glicksman ME. Philos Mag 2005;85:3.
et al. [2] and the front-tracking simulation study by Lazar [24] Rios PR, Glicksman ME. Act Mater 2006;54:1041.
et al. [31]. We also determined a one-to-one relationship [25] Rios PR, Glicksman ME. Act Mater 2007;55:1565.
between the average relative size of the grains and their [26] Rios PR, Glicksman ME. Act Mater 2008;56:1165.
number of faces; for relatively large grains (larger than [27] Steinbach I. Model Simul Mater Sci Eng 2009;17:073001.
about 0.5Rcr) the number of faces increases with R2, as is [28] http://www.openphase.de.
[29] Kim SG, Kim DI, Kim WT, Park YB. Phys Rev E 2006;74:061605.
predicted from the topological analysis. For smaller grains, [30] Barraless Mora LA, Gottstein G, Shvindlerman LS. Acta Mater
below 0.5Rcr, the relationship changes, which is not 2008;56:5915.
reported in previous studies. This behaviour may describe [31] Lazar EA, Mason JK, MacPherson RD, Srolovitz DJ. Acta Mater
the relationship between the size and shape distributions 2011;59:6837.
of the grains. [32] Feltham P. Acta Metall 1957;5:97.

You might also like