You are on page 1of 8

Available online at www.sciencedirect.

com

Acta Materialia 57 (2009) 3947–3954


www.elsevier.com/locate/actamat

Phase field study of precipitate growth: Effect of misfit strain


and interface curvature
R. Mukherjee a, T.A. Abinandanan a,*, M.P. Gururajan b
a
Department of Materials Engineering, Indian Institute of Science, Bangalore 560 012, India
b
Department of Applied Mechanics, Indian Institute of Technology-Delhi, Hauz Khas, New Delhi 110 016, India

Received 1 March 2009; received in revised form 25 April 2009; accepted 29 April 2009
Available online 27 May 2009

Abstract

We have used phase field simulations to study the effect of misfit and interfacial curvature on diffusion-controlled growth of an iso-
lated precipitate in a supersaturated matrix. Treating our simulations as computer experiments, we compare our simulation results with
those based on the Zener–Frank and Laraia–Johnson–Voorhees theories for the growth of non-misfitting and misfitting precipitates,
respectively. The agreement between simulations and the Zener–Frank theory is very good in one-dimensional systems. In two-dimen-
sional systems with interfacial curvature (with and without misfit), we find good agreement between theory and simulations, but only at
large supersaturations, where we find that the Gibbs–Thomson effect is less completely realized. At small supersaturations, the conver-
gence of instantaneous growth coefficient in simulations towards its theoretical value could not be tracked to completion, because the
diffusional field reached the system boundary. Also at small supersaturations, the elevation in precipitate composition matches well with
the theoretically predicted Gibbs–Thomson effect in both misfitting and non-misfitting systems.
Ó 2009 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Phase field modeling; Phase transformation kinetics; Precipitation

1. Introduction phase diagram, and the precipitate grows under a supersat-


uration of ðc1 –cme Þ.
In their classic study of diffusional growth of an isolated The Zener–Frank (ZF) theory has been extended to
precipitate into a supersaturated matrix, Zener [1] and study diffusional growth problems in a variety of settings;
Frank [2] established the parabolic growth law: the square see Ref. [3] for a recent review. Laraia, Johnson and Voo-
of the precipitate size (radius) increases linearly with time. rhees (LJV) extended the theory to diffusional growth of
They used what is now referred to as a ‘‘sharp interface” elastically misfitting precipitates [4]; the work of LJV built
model. Fig. 1 shows a schematic of a composition profile on the work of Eshelby [5,6], Larche and Cahn [7–10],
in the precipitate p and matrix m phases during growth. Johnson and Alexander [11], and Leo and Sekerka [12].
Local equilibrium at the (sharp) interface yields the matrix The main result of the LJV study is that the parabolic
interfacial composition, cmI , which is used as a boundary growth law continues to be valid for precipitates with mis-
condition for solving the diffusion equation; the far-field fit, but the growth coefficient depends on (i) misfit, (ii) the
composition, c1 is the other boundary condition. Neglect- elastic moduli of the two phases, (iii) interfacial stress and
ing interface curvature (capillarity) effects, cmI is the same as (iv) compositional stress.
cme , the equilibrium matrix composition obtained from the In the sharp interface model used by LJV, the effect of
dilatational misfit is to raise the matrix interfacial compo-
sition from cme to cmE , and therefore to decrease the supersat-
*
Corresponding author. Tel.: +91 8022932676; fax: +91 8023600472.
uration (by a constant factor) to c1 –cmE . Thus, neglecting
E-mail address: abinand@materials.iisc.ernet.in (T.A. Abinandanan). capillarity effects, the LJV theory predicts that the growth

1359-6454/$36.00 Ó 2009 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2009.04.056
3948 R. Mukherjee et al. / Acta Materialia 57 (2009) 3947–3954

(a) model resemble that used in the LJV theory as closely as


possible. Thus, the first goal of our study is a critical
comparison of the sharp interface results of LJV with those
from our simulations based on a (diffuse interface) phase
field model.
In ZF and LJV theories, which neglect the effect of inter-
face curvature, the diffusion fields at different times are self-
similar. This self-similarity is broken when capillarity effect
is included, since cmI (in Fig. 1a and b) is size dependent.
Capillarity, then, not only decreases the growth rate a,
but also renders it size-dependent. The second goal of this
study is to evaluate the extent of this reduction in a.
Our results also allow us to address two other issues:

1. Our simulation results on systems with low supersatura-


tion can be used to verify the predictions of Johnson and
Alexander [11] and of Leo and Sekerka [12] on the com-
position of a misfitting precipitate.
2. Many phase field models (including the one used in this
(b) study) make use of a phase field variable g whose pri-
mary role is to help distinguish one phase from the
other. In some settings, such as precipitation of an
ordered phase, g may be associated with a long-range
order parameter. In other settings (e.g. in the present
work, or in solidification), g lacks a direct physical sig-
nificance, and its role in the model is justified through
mathematical convenience. In this study, we show first
that our computational model with a phase field vari-
able g reproduces well-known analytical results from
ZF theory in one-dimensional (1D) systems in which
capillary effects are absent.

Our study is organized as follows: Section 2 presents an


outline of our phase field model. In Section 3, we first
establish the essential correctness of the use of a phase field
variable g through a critical comparison of its results
against those of 1D ZF theory, in which capillarity and
Fig. 1. Schematic of the composition profile in the matrix and precipitate elastic effects are absent. We then use the same model to
phases during growth in (a) non-misfitting and (b) misfitting systems. The study the growth of non-misfitting and misfitting circular
solid curve in the matrix shows the profile considered in the theories of (a)
precipitates, and critically compare our results with those
ZF and (b) LJV. Capillarity modifies the profile, as indicated by dashed
curve. from ZF and LJV theories. These results are analyzed in
order to elucidate the elastic and capillarity effects. These
results are critically discussed in Section 4, followed by a
coefficient a in misfitting systems is the same as that in non- set of conclusions in Section 5.
misfitting systems with ðc1 –cmE Þ as the supersaturation.
This conclusion is valid for systems in which diffusivity is 2. Phase field model
independent of the state of stress; our study is restricted
to these systems (as already mentioned, LJV have studied 2.1. Formulation
other stress-induced effects).
Direct experimental verification of the LJV results is We consider a binary alloy at constant temperature; an
extremely difficult, not only due to the ‘‘isolated particle” isolated precipitate of phase p grows into a matrix phase m.
assumption, but also due to lack of reliable data on stress We normalize compositions in such a way that the scaled
effects on diffusion. Moreover, the assumption of isotropic equilibrium compositions of m and p phases are cme ¼ 0
interfacial energy can also be difficult to realize experimen- and cpe ¼ 1, respectively. A two-phase microstructure in this
tally in crystalline alloys. Thus, simulations can act as ‘‘com- alloy is described in terms of composition cðr; tÞ (conserved
puter experiments” for validating the LJV results, since their variable) and order parameter gðr; tÞ (non-conserved vari-
parameters can be so tuned to make the computational able) in a periodic domain. The order parameter field g is
R. Mukherjee et al. / Acta Materialia 57 (2009) 3947–3954 3949

defined in such a way that g ¼ 0 in the m-phase and g ¼ 1 Table 1


in the p-phase; see Eq. (6) below. Parameters used in the simulations.
The microstructural evolution in our system is governed Parameter type Parameters Non-dimensional values
by the Cahn–Hilliard equation [13] and the Allen–Cahn Elastic parameters G m
800
equation [14]: m 0.3
d 0.5, 1.0 and 2.0
@c eT 0.01
¼ r  Mrl; ð1Þ
@t /ðgÞ g3 ð10  15g þ 6g2 Þ  12
@g dðF =N V Þ bðgÞ g3 ð10  15g þ 6g2 Þ
¼ L ; ð2Þ C ijkl
eff
1 ijkl ijkl
2 ðC m þ C p Þ
@t dg
Simulation parameters Dx 0.4
where M is the atomic mobility, l is the chemical potential Dy 0.4
and L is the relaxation coefficient for the order parameter. Dt 0.2
Chemical potential l is defined as the variational derivative System size 20000 in 1D
1024  1024 in 2D
of the total free energy per atom, F, with respect to the lo-
Allowed error 6 108
cal composition c, in displacements
dðF =N V Þ
l¼ : ð3Þ
dc
rel and el are elastic stress and strain tensors, respectively,
The total free energy of the system F is assumed to be and e0 is the the position-dependent eigenstrain (misfit
given by the sum of a chemical contribution F ch and an strain). The total strain eij is:
elastic contribution F el :  
1 @ui @uj
F ¼ F ch þ F el : ð4Þ eij ¼ þ ; ð12Þ
2 @rj @ri
The chemical contribution F ch is given by the following
where ui is the the displacement field.
functional:
Z Assuming both the m and p phases to be linear elastic,
2 2
F ch ¼ N V ½f ðc; gÞ þ jc ðrcÞ þ jg ðrgÞ dX; ð5Þ we have:
X
rel el
kl ¼ C ijkl eij ; ð13Þ
where f ðc; gÞ is the bulk free energy per atom, jc and jg are
the gradient energy coefficients for gradients in composi- where C ijkl is the elastic modulus tensor.
tion and order parameter, respectively, and N V is the num- The stress field rel ij obeys the equation of mechanical

ber of atoms per unit volume. equilibrium:


The bulk free energy density f ðc; gÞ is given by, rel
ij;j ¼ 0: ð14Þ
2
f ðc; gÞ ¼ f m ðcÞð1  W ðgÞÞ þ f p ðcÞW ðgÞ þ P g2 ð1  gÞ ; ð6Þ The eigenstrain e0ij and the elastic moduli C ijkl are
where P, a constant, sets the height of the free energy bar- assumed to depend on the order parameter as follows:
rier between the m-phase and p-phase. e0ij ðgÞ ¼ bðgÞeT dij ; ð15Þ
In Eq. (6), f m ðcÞ and f p ðcÞ stand for free energies of m
(matrix) and p (precipitate) phases, respectively. In our C ijkl ðgÞ ¼ C eff
ijkl þ /ðgÞDC ijkl ; ð16Þ
work, they are assumed to take the following simple forms: where eT is a constant that determines the strength of the
eigenstrain, dij is the Kronecker delta, bðgÞ and /ðgÞ are
f m ðcÞ ¼ Ac2 ; ð7Þ
scalar (interpolation) functions, where C eff ijkl is an ‘‘effective”
p 2
f ðcÞ ¼ Bð1  cÞ ; ð8Þ modulus, C pijkl and C mijkl are the elastic moduli tensor of the
p and m phases, respectively, and DC ijkl ¼ C pijkl  C mijkl . All
with positive constants A and B. In Eq. (6), W ðgÞ is an the model parameters used in our simulations are listed
interpolation function [15], given by: in their non-dimensional form in Table 1. Our non-dimen-
8 sionalization procedure is the same as that used in Ref.
< 0;
> for g < 0;
[16].
W ðgÞ ¼ g3 ð10  15g þ 6g2 Þ; for 0 6 g 6 1; ð9Þ For a configuration at time t, the equation of mechani-
>
:
1; for g > 1: cal equilibrium, Eq. (14), is solved using an iterative Fou-
rier spectral technique with periodic boundary conditions
The elastic contribution to the free energy is:
Z (described in detail in Gururajan and Abinandanan [16];
1 see also Refs. [17–21]) to yield the elastic stress and strain
F el ¼ rel ; el dX ð10Þ
2 X ij ij fields; these, in turn, are used to integrate the Cahn–Hil-
where liard (Eq. (1)) and Allen–Cahn (Eq. (2)) equations over a
time step Dt to yield the configuration at t þ Dt. For this
0
eel
ij ¼ eij  eij ; ð11Þ time integration step, we use a semi-implicit Fourier
3950 R. Mukherjee et al. / Acta Materialia 57 (2009) 3947–3954

spectral technique, due to Chen and Shen [22]; the Fourier (a)
transforms are performed using the freeware called FFTW,
developed by Frigo and Johnson [23]. Microstructural evo-
lution is simulated through a repeated application of this
procedure on the new configuration at the end of each time
step.

3. Results

In our simulations, we place a small particle of initial


radius Ro and composition cpe at the centre of our simula-
tion cell with periodic boundary conditions; the initial
composition outside the particle is set to c1 everywhere;
with the normalizing scheme we have used, c1 is numeri-
cally the same as the supersaturation (or equilibrium vol-
ume fraction): n ¼ ðc1  cme Þ=ðcpe  cme Þ. As the simulation
proceeds, we track the particle radius, operationally (b)
defined as the distance from the centre where g ¼ 0:5,
and hence the growth coefficient a. The simulations end
when either the diffusion field or the elastic stress field from
neighbouring simulation cells begin to overlap; the values
of a obtained just before the simulations end are used in
the plots. They are also listed in Table 1 along with Rf ,
the final precipitate radius, and nE ¼ ðc1 –cmE Þ=ðcpe –cme Þ, the
effective (i.e. elasticity-corrected) supersaturation.

3.1. Zener–frank theory

We first compare our 1D simulation results with those


from the ZF theory. In Fig. 2, we have plotted the growth
coefficient a against matrix supersaturation n. The data
points from our 1D simulations are in excellent agreement Fig. 3. A comparison of scaled composition profiles in (a) a 1D and (b) a
with the analytical results of ZF theory for 1D systems (the 2D system with n ¼ 0:4. The dashed curve in the matrix is from the ZF
lowest curve) for all supersaturations. This agreement is sharp interface theory, while the solid and dotted curves are from phase
field simulations at two different particle sizes. The distance on the x-axis
not just in the growth coefficient, but also in the composi-
is scaled by the instantaneous particle size. In the legend, S and P stand for
tion profile in the matrix phase in Fig. 3a. sharp interface theory and phase field simulations, respectively.

1D Thus, this excellent agreement in the baseline case— the


1.6
2D one without interface curvature and without misfit
3D
1D strains—is our main evidence for the essential correctness
2D
of the use of a phase field variable g in our model.
1.2 Fig. 2 also displays a comparison of the 2D growth coef-
ficients from the ZF theory and our simulations. Values of
a from our simulation are in excellent agreement with those
α

0.8 from ZF theory for a supersaturation of n ¼ 0:4 (within


0.4% at a final particle size of Rf ¼ 60), but become less
so at lower supersaturations. At n ¼ 0:1, growth coeffi-
0.4
cients from ZF theory and our simulations differ by about
13% at Rf ¼ 21. The main conclusion from these results is
0 that particles of smaller size, which experience a stronger
0 0.2 0.4 0.6 0.8 effect due to interface curvature, grow slower than pre-
ξ dicted by ZF theory.
As shown in Fig. 1a, the curvature effect may be
Fig. 2. Dependence of growth coefficient a on matrix supersaturation n.
The three curves are from the ZF (sharp interface) theory for 1D, 2D and
explained using the elevation of matrix interfacial composi-
3D growth. The data points are from our 1D and 2D phase field tion from cme from zero to cmI , which results in a reduced
simulations. supersaturation. Since the reduction in supersaturation
R. Mukherjee et al. / Acta Materialia 57 (2009) 3947–3954 3951

1.8 No Misfit, S
c∞ = 0.1 , S
c∞ = 0.4 , S Misfit, S
1.6 1.6 Misfit, P
c∞ = 0.1 , P
c∞ = 0.4 , P
1.4
1.2
1.2

α
α

0.8
0.8

0.6
0.4
0.4

0.2
0
0 0 0.2 0.4 0.6 0.8
0 10 20 30 40 50 60 70
R ξ

Fig. 4. Effect of capillarity on instantaneous growth coefficient a, plotted Fig. 5. Dependence of growth coefficient a on matrix supersaturation n in
as a function of instantaneous particle radius R in 2D, non-misfitting a misfitting system with d ¼ 0:5. The dashed curve is for LJV sharp
systems. The horizontal lines are from the ZF sharp interface theory interface theory in 2D, and the data points are from 2D simulations. The
(without considering capillarity effects), and the data points are from our solid curve, from the 2D ZF theory for non-misfitting precipitates, is also
2D phase field simulations. shown for comparison.

for a given particle size is a larger fraction of the original and simulations is about 18% at Rf ¼ 21. Thus, the com-
supersaturation in a less concentrated matrix, the curvature bined effect of misfit and capillarity in Fig. 5 mimics that
effect on growth coefficient is much higher at n ¼ 0:1 than of capillarity alone in Fig. 2.
at n ¼ 0:4. Thus, in Fig. 4, in which the instantaneous val- In Fig. 6, we find a good agreement between the simu-
ues of growth coefficient are plotted against instantaneous lated (scaled) composition profile for two different times
particle size, the data points from our simulations are much with that for the LJV model. This agreement is similar to
closer to the ZF result in the latter, even at a particle size of that for the non-misfitting case in Fig. 3b.
R ¼ 20. We return to a discussion of the curvature effect in
Section 4. 3.3. Effect of interface curvature
Similar to the 1D case in Fig. 3a, the composition profile
in the matrix at two different sizes obtained from our sim- A more detailed assessment of the role of capillarity is
ulations show good agreement, in Fig. 3b, with that from possible by considering cpI , the precipitate interfacial com-
the ZF theory for a matrix composition of n ¼ 0:4. position.
 In the  sharp interface model, the difference

DcpI ¼ cpI  cpe , is proportional to DcmI ¼ cmI  cme :
3.2. Effect of a misfit strain: LJV theory

In these simulations, we have used a dilatational misfit


of 1%. The precipitate and matrix phases are elastically iso- S
1 time R 35, P
tropic, with a Poisson’s ratio of 0.3. The shear moduli of R 55, P
the two phases, however, can be different. We have consid- 0.8
ered three different cases with ðGp =Gm Þ ¼ d ¼ 0:5, 1.0 and
2.0, though, for the sake of clarity, we present our plots
0.6
only for d ¼ 0:5.
c

As we mentioned in Section 1, if curvature effects are


0.4
neglected, the growth coefficient vs. supersaturation curve
(dashed curve in Fig. 5) from the LJV theory for the misfit-
0.2
ting case is obtained by shifting the ZF curve (solid curve,
for the non-misfitting case) to the right by cmE .
0
In Fig. 5, the growth coefficients a from our simulations
are plotted against supersaturation, along with the LJV 1 2 3
result for a system with elastically soft precipitates Scaled Distance
ðd ¼ 0:5Þ. The agreement is quite good at high supersatura-
Fig. 6. A comparison of scaled composition profiles in a 2D misfitting
tions (with a difference of just 0.1% at a particle size of
system with n ¼ 0:4. The dashed curve in the matrix is from the ZF sharp
Rf ¼ 57 in an alloy with n ¼ 0:4Þ. With decreasing solute interface theory, while the solid and dotted curves are from phase field
concentration, however, the agreement becomes worse; at simulations at two different particle sizes. The distance on the x-axis is
n ¼ 0:1, for example, the difference between the LJV theory scaled by the instantaneous particle size.
3952 R. Mukherjee et al. / Acta Materialia 57 (2009) 3947–3954

(a) S In simulations using diffuse interface models, interfacial


1.08
c∞ = 0.1, P compositions cpI and cmI cannot be measured, as the compo-
c∞ = 0.4, P
sition changes continuously across an interface region of
finite width. However, if we assume that the chemical
potential gradients are negligible inside the precipitate,
1.06
the composition anywhere in the precipitate should be
the same as cpI .
p

In Fig. 7, we have plotted cpI , obtained from simulations


cI

1.04
against the inverse of particle size R in (a) non-misfitting
and (b) misfitting systems. In Fig. 7a, the straight line rep-
resents the Gibbs–Thomson result (Eq. (18) without the
1.02
stress and strain terms), and the data points from our sim-
ulations are compositions at the centre of the precipitate;
for clarity, we have presented data only for two supersatu-
1 rations: n ¼ 0:1 and n ¼ 0:4. This figure shows that, in a
0 0.02 0.04 0.06 0.08 0.1
growth setting, particles growing in more concentrated
1/R
alloys display a smaller Gibbs–Thomson effect.
(b) In Fig. 7b, we compare the theoretical value of cpI (the
dashed line) with those from simulations for n ¼ 0:1 and
1.08
n ¼ 0:4. This figure also shows that the Gibbs–Thomson
effect is less completely realized for more supersaturated
alloys, leading to a better match between a from simula-
1.06
tions and that from theory (see Figs. 2 and 5).
Thus, Fig. 7 reveals yet another reason for the closer
cIp

agreement between ZF theory and our 2D simulations for


1.04 more concentrated alloys: Gibbs–Thomson effect is less
completely realized in systems with a higher supersaturation.
At n ¼ 0:1, even though our results were obtained in a
1.02 growth setting, the small supersaturation ensures that the
S precipitate experiences a large part of the Gibbs–Thomson
P, c∞ = 0.1
P, c∞ = 0.4 effect. Thus, in the misfitting case, DcpI ¼ 0:066 from our sim-
1 ulations (at Rf ¼ 21) matches well with the theoretical result
0 0.02 0.04 0.06 0.08 0.1
of DcpI ¼ 0:067 calculated from Eq. (18). This agreement in
1/R Fig. 7b may be taken as an indirect verification of the predic-
Fig. 7. Dependence of precipitate composition on particle radius in (a) tions of Johnson and Alexander [11] and Leo and Sekerka
non-misfitting and (b) misfitting systems. The straight line is for cpI , [12] on the phase compositions across a sharp interface.
precipitate composition at a curved interface, given by the Gibbs–
Thomson effect (see Eqs. (17) and (18)). The data points represent the
composition at the centre of the precipitate in simulations with n ¼ 0:1
4. Discussion
(open circles) and n ¼ 0:4 (filled circles).
As we mentioned in Section 1, the primary aim of this
Wm article is to use our ‘‘computer experiments” to validate
DcpI ¼ p DcmI ð17Þ
W the LJV theory of elastic stress effects during precipitate
where W ¼ ½@ 2 f =@c2 ce in the designated phase, and DcmI is growth. The following features of our phase field model
given by the generalized Gibbs–Thomson effect [24]: make it resemble closely the alloy model used in the LJV
rmij ðemij  epij Þ þ ½rpij ðepij  eT dij Þ  rmij emij =2
theory: isotropic interfacial energy, isotropic atomic mobil-
vc
DcmI ¼ þ p : ð18Þ ity, constant atomic diffusivity in the matrix (and in the
ðcpe  cme ÞWm ðce  cme ÞWm
precipitate too, though this is not an essential part of the
In the above equation, rij and eij are the stresses and LJV theory), isotropic elastic moduli, and constant misfit
strains, respectively, in the designated phase, eT is the dila- in the precipitate phase. While the LJV theory includes
tational eigenstrain and v is the mean interface curvature. cases where composition gradients may engender stress,
In our 2D simulations, v ¼ 1=R; from Eqs. (7) and (8) (with our model parameters are such that they do not.
the parameter values of A ¼ B ¼ 1), we get Wm ¼ Wp ¼ 2. However, interfacial curvature (and hence the Gibbs–
For the values of misfit, shear modulus and Poisson’s ratio Thomson effect) is one feature which is always present in
used in this study (see Table 1), the (constant) elasticity phase field simulations (except in 1D), but absent in the
contribution to DC mI —the first term on the right-hand side LJV and ZF theories. Its primary consequence is to
of Eq. (18)) — is 0.044, 0.056 or 0.066 for systems with dampen the growth rate. However, this dampening effect
d ¼ ðGp =Gm Þ ¼ 0:5, 1.0 or 2.0, respectively. keeps decreasing with increasing particle size, and the
R. Mukherjee et al. / Acta Materialia 57 (2009) 3947–3954 3953

Table 2 instabilities are promoted at larger sizes. In order to test


Comparison of growth coefficients from theory ðat Þ and simulations ðas Þ. the role of fluctuations in our system, we introduced a ran-
System n nE Rf at as 100  ðas  at Þ=at dom noise of magnitude 0:05% in both c and g at each time
No misfit 0.1 0.1 21 0.369 0.321 13.0 step in our 2D simulations with a circular particle. At the end
0.2 0.2 35 0.628 0.604 3.8 of these simulations the particle remains circular, indicating
0.3 0.3 48 0.897 0.885 1.3 that the fluctuations have not had any effect on particle shape
0.4 0.4 60 1.199 1.194 0.4
in the size regime explored in this study.
Misfit, d ¼ 0:5 0.1 0.056 21 0.246 0.202 17.9
0.2 0.156 32 0.514 0.502 02.3
5. Conclusions
0.3 0.256 44 0.775 0.776 +0.1
0.4 0.356 57 1.061 1.060 0.1
Misfit, d ¼ 1:0 0.1 0.044 19 0.208 0.145 30.3
0.2 0.144 31 0.482 0.469 02.7
(1) Using a our phase field simulations as computer
0.3 0.244 43 0.742 0.744 +0.3 experiments, we have validated the LJV theory of
0.4 0.344 56 1.023 1.031 +0.8 growth of misfitting particles in systems in which
Misfit,d ¼ 2:0 0.1 0.034 16 0.178 0.068 61.8
composition differences do not engender stress.
0.2 0.134 30 0.458 0.444 3.1 (2) Capillarity effects decrease instantaneous growth
0.3 0.234 42 0.717 0.722 +0.7 rate; this dampening effect is more pronounced at
0.4 0.334 55 0.996 1.003 +0.7 smaller sizes and at smaller supersaturations.
(3) At constant particle size, capillarity has a smaller
effect in alloys with larger supersaturations, because
instantaneous growth coefficient keeps rising towards the the Gibbs–Thomson effect is less completely realized.
theoretical values (e.g. Fig. 4). It is clear from Figs. 2 and (4) At low supersaturations, the Gibbs-Thomson effect is
5 that capillarity has a larger dampening effect on a at more completely realized, and our results on precipi-
lower supersaturations. In misfitting systems, the available tate compositions are in agreement with those pre-
(effective) supersaturation nE ¼ ðc1  cmE Þ=ðcpe  cme Þ is dicted from the theory of thermodynamics of
smaller when the precipitate is stiffer (see Table 2). Thus stressed solids.
we find that a values from simulations are smaller than the-
oretical values by 18%, 30% and 62%, when d ¼ 0:5; 1:0
Acknowledgements
and 2.0, respectively, in systems with n ¼ 0:1.
Our 2D simulations (with and without misfit) show that
The authors thank the Volkswagen Foundation for
the Gibbs–Thomson effect is only partially realized during
financial support for this work. They also thank Prof. V.
growth, especially at large supersaturations. Since the
Jayaram and Dr. H. Ramanarayan for discussions.
Gibbs–Thomson effect dampens the growth rate, its partial
realization helps bring the growth rate closer to the theoret-
References
ically predicted rate in these alloys—see Figs. 2 and 5 for
2D growth. [1] Zener C. Theory of growth of spherical precipitates from solid
Clearly, with higher (lower) atomic mobility in the p solution. J Appl Phys 1949;20(10):950–3.
phase, the Gibbs–Thomson effect can be realized more (less) [2] Frank FC. Radially symmetric phase growth controlled by diffusion.
completely. Thus, if the atomic mobility in the p phase is Proc Roy Soc London. Ser A, Math Phys Sci 1950;201(1067):586–99.
[3] Sekerka RF, Wang S-L. Moving phase boundary problems. In:
higher (lower), the approach of the growth coefficient (in Aaronson HI, editor. Lectures on the theory of phase transforma-
Fig. 4) towards the ZF or LJV result will be delayed tions. Warrendale (PA): TMS of AIME; 1999. p. 231–84.
(hastened) to larger precipitate sizes. Our simulations, how- [4] Laraia VJ, Johnson WC, Voorhees PW. Growth of a coherent
ever, cannot address this issue, since they use a constant precipitate from a supersaturated solution. J Mater Res 1988;3(2):
mobility in both m and p phases. 257–66.
[5] Eshelby JD. The determination of the elastic field of an ellipsoidal
We now turn to other similar studies that compared inclusion, and related problems. Proc Roy Soc London. Ser A, Math
results from sharp and diffuse interface models in a kinetic Phys Sci (1934–1990) 1957;241(1226):376–96.
setting. Specifically, we mention Chen and co-workers, who [6] Eshelby JD. Elastic inclusions and inhomogeneities. In: Sneddon N,
have compared their phase field results on diffusion fields Hill R, editors. Progress in solid mechanics, vol. II; 1961. p. 87–140.
around a precipitate with those obtained by solving a diffu- [7] Larche F, Cahn JW. A linear theory of thermochemical equilibrium
of solids under stress. Acta Metal 1973;21(8):1051–63.
sion equation. However, the focus of their work was on dif- [8] Larche F, Cahn J. The effect of self-stress on diffusion in solids. Acta
fusion fields in ternary systems [25] or on lengthening and Metall 1982;30(10):1835–46.
thickening of plate-like precipitates [26]. [9] Cahn JW, Larche FC. A simple model for coherent equilibrium. Acta
Thermally induced fluctuations in composition (and Metall 1984;32(11):1915–23.
order parameter) may induce instabilities in the shape of a [10] Larche FC, Cahn JW. The interactions of composition and stress in
crystalline solids. Acta Metall 1985;33(3):331–57.
growing precipitate; in addition to the classical Mullins– [11] Johnson WC, Alexander JID. Interfacial conditions for thermome-
Sekerka instability [27,28], stress-induced shape bifurcation chanical equilibrium in two-phase crystals. J Appl Phys 1986;59(8):
[29,30] may also be caused by fluctuations. These shape 2735–46.
3954 R. Mukherjee et al. / Acta Materialia 57 (2009) 3947–3954

[12] Leo P, Sekerka RF. The effect of surface stress on crystal-melt and [22] Chen LQ, Shen J. Applications of semi-implicit fourier-spectral
crystal–crystal equilibrium. Acta Metall 1989;37(12):3119–38. method to phase field equations. Comput Phys Commun 1998;108(2):
[13] Cahn JW. On spinodal decomposition. Acta Metall 1961;9:795–801. 147–58.
[14] Allen SM, Cahn JW. A microscopic theory for antiphase boundary [23] Frigo M, Johnson SG. FFTW: an adaptive software architecture for
motion and its application to antiphase domain coarsening. Acta the FFT (<http://www.fftw.org>). In: Proceedings of the 1998 IEEE
Metall 1979;27(6):1085–95. international conference on acoustics, speech and signal processing,
[15] Wang SL, Sekerka RF, Wheeler AA, Murray BT, Coriell SR, Braun vol. 3; 1998. p. 1381–4.
RJ, et al. Thermodynamically-consistent phase-field models for [24] Johnson W. Precipitate shape evolution under applied stress –
solidification. Physica D 1993;69(1–2):189–200. thermodynamics and kinetics. Metall Trans A 1987;18(2):233–48.
[16] Gururajan MP, Abinandanan TA. Phase field study of precipitate [25] Chen LQ, Ma N, Wu K, Wang Y. Quantitative phase field modeling
rafting under a uniaxial stress. Acta Mater 2007;55:5015–26. of diffusion-controlled precipitate growth and dissolution in Ti–Al–V.
[17] Anthoine A. Derivation of the in-plane elastic characteristics of Scripta Mater 2004;50(6):471–6.
masonry through homogenization theory. Int J Solids Struct 1995; [26] Hu SY, Murray J, Weiland H, Liu ZK, Chen LQ. Thermodynamic
32(2):137–63. description and growth kinetics of stoichiometric precipitates in the
[18] Moulinec H, Suquet P. A numerical method for computing the phase-field approach. Comput Coupling Phase Diag Thermochem
overall response of nonlinear composites with complex microstruc- 2007;31(2):303–12.
ture. Comput Methods Appl Mech Eng 1998;157(1):69–94. [27] Mullins WW, Sekerka RF. Morphological stability of a particle
[19] Michel H, Moulinec JC, Suquet P. Effective properties of composite growing by diffusion of heat flow. J Appl Phys 1963;34(2):323–9.
materials with periodic microstructure: a computational approach. [28] Doherty RD. Role of interfaces in kinetics of internal shape changes.
Comput Methods Appl Mech Eng 1999;172(1):109–43. Metal Sci 1982;16:1–13.
[20] Hu SY, Chen LQ. A phase-field model for evolving microstructures [29] Johnson WC, Cahn JW. Elastically induced shape bifurcations of
with strong elastic inhomogeneity. Acta Mater 2001;49(11):1879–90. inclusions. Acta Metall 1984;32(11):1925–33.
[21] Zhu J, Chen L-Q, Shen J. Morphological evolution during phase [30] Johnson WC. Influence of elastic stress on phase transformations. In:
separation and coarsening with strong inhomogeneous elasticity. Aaronson HI, editor. Lectures on the theory of phase transforma-
Model Simulat Mater Sci Eng 2001;9(6):499–511. tions. Warrendale (PA): TMS of AIME; 1999. p. 35–133.

You might also like