Вы находитесь на странице: 1из 11

Acta Materialia 55 (2007) 16291639

www.actamat-journals.com

Eects of anisotropic growth on the deviations from


JohnsonMehlAvrami kinetics
Feng Liu *, Gencang Yang
State Key Laboratory of Solidication Processing, Northwestern Polytechnical University, Xian, Shanxi 710072, China
Received 19 June 2006; received in revised form 19 October 2006; accepted 19 October 2006
Available online 26 December 2006
This article is for the purpose of congratulating the 80th birthday of Chinese academician of Science, Prof. Yaohe Zhou.

Abstract
The eects of nucleation, growth and impingement on deviations from classical JohnsonMehlAvrami (JMA) kinetics are described.
On the basis of Monte Carlo (MC) simulations of isothermal phase transformations, the deviations from the JMA kinetics due to the
eects of anisotropic growth have been investigated further. An analytical approach has been adopted to describe phase transformations
subjected to the eects of anisotropic growth. In combination with MC simulations, the eect of the memory time has been shown
when the transition from JMA behavior to blocking behavior occurs during the transformation. On this basis, a physically realistic
Avrami exponent has been deduced. A comparative study between the analytical description and a gammaalpha transformation occurring in an FeMn alloy under an applied load has been carried out.
 2006 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: JMA kinetics; Anisotropic growth; Avrami exponent

1. Introduction
The JohnsonMehlAvrami (JMA) equation [110] is
generally applied to describe phase transformations involving nucleation and growth. This equation provides an
expression for the volume fraction of material transformed
as a function of time, f(t). The JMA description is precise if
the assumptions imposed in their original derivations are
not violated. However, for many important physical situations, these conditions are not satised, and generalizations
of the JMA theory are required. Although some extensions
of the JMA theory have been made [1115], additional
work needs to be completed, especially in the case of grain
formation through anisotropic growth.
Monte Carlo (MC) simulations are well suited to test
JMA kinetics because they fully account for the statistical
nature of nucleation and subsequent anisotropic growth
*

Corresponding author. Tel.: +86 029 88460374.


E-mail address: liufeng@nwpu.edu.cn (F. Liu).

(mutual hindrance of growing grains). The eects of anisotropic growth on deviations from ideal JMA kinetics have
been numerically quantied previously [16]. It was
observed that the deviations become signicant as the
eects of anisotropic growth increase.
MC calculations incorporating deviations from JMA
kinetics due to anisotropic growth have recently been
extended [17], assuming a time(/temperature)-dependent
nucleation rate and a time-independent and temperaturedependent growth rate in combination with dierent
growth modes. A concept of memory time (i.e. the rst
time steps when transformation is purely JMA-like before
the blocking of growing grains) has been introduced in
the transformation kinetics to evaluate the eect of anisotropic growth. A novel analytical description (see Section
2.2) has been developed which incorporates classical JMA
kinetics, but also accounts for the blocking if the anisotropically growing grains have perpendicular anisotropy axes.
All the MC results could be reproduced and are thus well
explained by the novel analytical description [17].

1359-6454/$30.00  2006 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2006.10.022

1630

F. Liu, G. Yang / Acta Materialia 55 (2007) 16291639

However, the MC simulations described above and


the analytical descriptions involve some inherent aws,
with respect to both the Avrami exponent and its general
application (see Section 2.2.4). In the present work, analytical expressions for the Avrami exponent of transformations subjected to the eects of anisotropic growth
have been obtained by applying the analytical approach
presented in Ref. [8]. On this basis, an analytical description of MC simulations is given (see Sections 3 and 4.1).
Moreover, a comparative study between the analytical
description and the gammaalpha transformation occurring in an FeMn alloy under applied loads has been
carried out (see Section 4.2). This study not only shows
possible applications of the current analytical phase
transformation model, but also provides new insight
and understanding of phase transformations subjected
to the eects of anisotropic growth in real materials.
2. Theoretical background
2.1. Classical JMA kinetics
JMA kinetics hold only for extreme conditions, i.e.
either pure continuous nucleation or pure site saturation
as nucleation mechanisms [58]. This phenomenon has
been demonstrated using the analytical phase transformation model [8]. If the explicit eects of transient nucleation are excluded, the nucleation rate per unit of
volume (in three dimensions) or per unit of area (in
two dimensions) of untransformed material, I, as a function of time, t, and temperature, T, can be expressed as
[7,8]


Qn
It; T I 0 exp
1
RT
with R as the gas constant and Qn the activation energy for
nucleation. The term site saturation is used where the number of (supercritical) nuclei does not change during the
transformation, i.e. all nuclei of number N* per unit volume are already present at t = 0 [7,8]:
IT N  dt  0

with d(t  0) denoting the Dirac function.


Previous work [1820] proposed to introduce the
nucleation index to allow for the dependence of the
nucleation rate on the degree of transformation. The continuous nucleation rate equation (for isothermal transformation) in that case can be expressed as I(t) = an 0 ta1,
with n 0 as a constant (a = 0 for zero nucleation rate
(i.e. site saturation), a = 1 for constant nucleation rate
and a > 1 for increasing nucleation rate with the progress
of transformation). The eect of the nucleation index on
the transformation is not the main focus in the present
work; however, detailed descriptions are available in
Refs. [1821]. For the case of interface-controlled growth,
the textbook equation for the interface velocity G is given
by [5]:





DGa
DG
1  exp 
G G0 exp 
RT
RT

3a

with G0 as the pre-exponential factor for growth, DGa the


activation energy for growth that equals the interface energy barrier and DG the driving force, which is the energy
dierence between the new phase and the parent phase.
For large undercooling or superheating, i.e. DG (>0) is
suciently large compared with RT, Eq. (3a) becomes
[5,7,8]


Q
G G0 exp  G
3b
RT
with QG (=DGa) as the activation energy for growth and G0
the temperature-independent interface velocity. For the
case of diusion-controlled growth, QG equals the activation energy for diusion, QD, and G0 equals the preexponential factor for diusion D0.
The overall kinetics of transformations involving nucleation and growth, i.e. the volume fraction transformed, f,
as a function of time t, for isothermal conditions is generally described by:
n

f 1  expkt
 
Q
kT k 0 exp
RT

4
5

The expressions for n, Q and k0 are given in Table 1.


2.2. Deviations from JMA kinetics
The necessary associated size-dependent growth [9,10]
and transient nucleation [11,12] or a mixture of nucleation
modes [7,8] lead to deviations from JMA kinetics.
Anisotropic growth (possibly leading to the blocking of
growing grains by the neighboring ones) is another important origin for deviations from JMA kinetics [13]. Therefore, deviations from JMA kinetics may be caused by (i)
nucleation or (ii) growth and impingement.
Table 1
Expressions for the (time and temperature independent growth exponent,
n, the overall activation energy, Q, and the rate constant, k0, to be inserted
into Eqs. (4) and (5) for isothermal and isochronal annealing, respectively

Continuous nucleation
n
Q

Isothermal

Isochronal

d/m+1

d/m + 1

d
mQG QN

d
mQG QN

k n0

n
d=m
gN 0 m0
d=m1

Site saturation
n
Q

d/m
QG

k n0

gN  m0

n
d=m
gN 0 m0 C C
d=m

d=m1QG

d/m
QG
d=m

d=m

gN  m0

d=m

QG

For C c see Ref. [8]. These values are valid for the JMA kinetics based on
continuous nucleation (a = 1) or site saturation, with m as growth mode
parameter (m = 1 for interface-controlled growth; m = 2 for diusion
controlled growth), and d as the dimensionality of the growth (d = 1, 2, 3).

F. Liu, G. Yang / Acta Materialia 55 (2007) 16291639

2.2.2. Inuence of growth and impingement


As shown in Ref. [17], isotropic growth and parallel
growth of anisotropically growing grains with identical
convex shapes, in combination with continuous nucleation, does obey JMA kinetics, which is compatible with
impingement due to randomly dispersed nuclei. It is
assumed here that the nuclei are dispersed randomly
throughout the whole volume, but suppose that at time
t the actually transformed volume is Vt. If the time is
increased by dt, the extended and the actual transformed
volumes will increase by dVe and dVt, respectively. From
the change in the extended volume, dVe, only a part will
contribute to the change in the actually transformed volume, dVt, namely, a part as large as the untransformed
volume fraction [15]. Hence


V Vt
df
dV t
1  f
6
dV e ;
dxe
V
Fig. 2(b) and (c) in Ref. [17] shows that the s (square), r
(rectangular), rNRp (parallel anisotropic) and np (parallel
needle) growth modes, to a large extent, obey JMA
kinetics, i.e. the data in an ln(ln(1  f)) vs. ln(t) plot
fall on a straight line, and the Avrami exponent, derived
from the slope in Fig. 2(b) of Ref. [17], is constant for
these four modes almost throughout the entire interval
f = 0 to f = 1.
In the case of anisotropically growing particles, i.e.
anisotropic orthogonal growth rNRo and 1-D orthogonal
growth no [17], the average time interval for the randomly dispersed particles to grow before blocking by
other particles is smaller than the average time interval
for isotropic growth, leading to strong deviations from
JMA kinetics. One phenomenological approach regarding
this blocking eect has been proposed by extending
Eq. (6) to [17,21,22]:
df
e
1  f
dxe

where e P 1. Impingement due to Eq. (7) is more severe


(i.e. the dierence between fand xe is larger than the difference due to Eq. (6)) and increases with e (see Fig. 1).

0.8

real fraction, f

2.2.1. Inuence of nucleation


In principle, JMA kinetics are not applicable if a mixture of nucleation mechanisms are active. A situation of
mixed nucleation occurs when pre-existing nuclei are present at the onset of the transformation, and a constant
nucleation rate holds upon transformation [7,8]. Therefore, the Avrami exponent has an initial value of 1, 2
or 3, but as the transformation proceeds it changes to a
value of 2, 3 or 4 that holds for continuous nucleation
and linear growth (one-, two- or three-dimensional (1-D,
2-D or 3-D)), respectively.
Furthermore, nucleation has to occur randomly
throughout the innite space, and transient nucleation cannot be considered [15] for JMA kinetics to hold.

1631

0.6

0.4

=3
=2
=1.5
=1.2
=1

0.2

0
0

10

extended fraction, x

Fig. 1. The transformed fraction, f, as a function of the extended fraction,


xe, for the case of impingement by anisotropically growing grains,
corresponding to values of e P 1.

Impingement due to Eq. (7) implies that the eects of


anisotropic growth change only the relationship between
f and xe (see Fig. 1), but does not change xe itself, which
is not compatible with MC simulations and leads to poorer
ts [17].
2.2.3. Analytical description based on MC simulations [17]
In order to describe the eects of anisotropic growth on
the deviations from JMA kinetics, a relatively simple and
transparent analytical description has been developed,
extending from and incorporating the JMA theory. This
analytical description accurately reproduces the numerical
results of MC simulations [17]. The process starts with formulations that closely follow MC simulations with discrete
time steps and the corresponding increment of the fraction
transformed within every single period but ends with traditional continuum equations, as in the usual JMA kinetics
[17].
Generally, impingement leads to the blocking of the
growing grains and thus reduces the time interval for grains
to grow, i.e. to contribute to the fraction transformed. In
standard JMA kinetics, this time interval (i.e. the memory) always goes back to t = 0, so that all the preceding
time steps do contribute [17]. Upon transformation, their
contributions become less important due only to the
impingement factor (1  f); see Eq. (6).
Following this philosophy, an exact agreement between
the analytical description and MC simulations was
achieved for the np growth mode. Accordingly, a standard
JMA equation for this parallel 1-D growth with continuous
nucleation in 2-D space results [17]:
 ln1  f wIGt2

where w is the width of the growing 1-D grain and G is the


growth rate (unit length per unit time). If w = 1/2, Eq. (8) is

1632

F. Liu, G. Yang / Acta Materialia 55 (2007) 16291639

compatible with Eq. (4) for n = 2, and k = (IG/2)1/2. In the


case of blocking events originating from the no growth
mode, the memory does not revert to t = 0 but is limited
to a certain number of previous time steps [17]. Thus an
equation dierent from Eq. (8) results [17]:
p
2
 ln1  f 2Cw I t  t wIGt
9

simulations is carried out, and physically realistic solutions for the Avrami exponent are analytically obtained
(see Section 3).

where t*, the time when the transition from JMA behavior
to blocking behavior occurs, is given by:
s
2
3 C
t
10
IG2

For the np growth mode with continuous nucleation in a


2-D space, combining Eq. (8) (with w = 1/2) and (1)(4)
gives

with C as a constant with a value close to 2 [17]. For t < t*,


JMA kinetics, i.e. the last term in Eq. (9), holds. Analogously, an equation for the rNRo growth mode was obtained as [17]:


p
2
4C
 ln1  f 2Cw It  t 1 

NR
3NR
p
2
3=2
3
IG2 t
11
 I Gt  t
3NR
where t*, the time after which Eq. (11) becomes valid, is still
given by Eq. (10). A detailed description for derivations of
Eqs. (8), (9) and (11) is available in Ref. [17].
2.2.4. Inherent aws in the analytical description based on
MC simulations
The Avrami method of plotting ln(ln(1  f)) vs. ln(t)
is commonly adopted to deduce the Avrami exponent, n,
for transformations following JMA or non-JMA kinetics.
Namely, whether the data in an ln(ln(1  f)) vs. ln(t)
plot fall on a straight line or not determines whether a
transformation follows JMA or non-JMA kinetics [5,17].
Actually, the ln(ln(1  f)) vs. ln(t) plot is not applicable
for non-JMA transformations. The Avrami exponent for
non-JMA transformations on the basis of Eqs. (9) and
(11) cannot be deduced reasonably by using an Avrami
plot, even if they t MC simulations well [17].
In practice, an analytical description of transformations is favored, as numerical calculations can be very
cumbersome (one complication might be the distinction
between independent and dependent t parameters
[21,23]). Analytical expressions have an advantage over
numerical calculations because the inuence of, for example, the dierent nucleation, growth and impingement
models can be easily identied and investigated
[21,23,25]. On this basis, an accurate, exible analytical
phase transformation model has been developed that
incorporates a choice of nucleation, growth and impingement mechanisms [8,21] which has been successfully
applied [21,23,25].
The analytical description (i.e. Eqs. (8), (9) and (11))
aims solely at interpreting MC simulations [17]. However,
its application to general transformations is still an open
question. Applying the analytical approach [8], reinterpretation of the analytical description on the basis of MC

3. An analytical phase transformation model based on MC


simulations

ln ln1  x ln1=2I 0 G0 

QN QG
2 ln t
RT

12

Following JMA kinetics, a constant Avrami exponent,


n = 2, and a constant eective activation energy Q =
(QN + QG)/2, are obtained (see Table 1).
For the no growth mode with continuous nucleation in a
2-D space, Eq. (9) can be used to describe the non-JMA
transformation subjected to the eect of anisotropic
growth. From Eq. (9), the extended fraction can be considered to be composed of two parts, i.e. xe = xe1 + xe2, one
part with n1 = 0.5, the other with n2 = 2. The ratio between
the two extended fractions is given by
wI 1 G1 t2
r2 xe2
p
2Cw I 1 teff r1 xe1

13

where te = t  t*. Analogous to the analytical approach


adopted in Ref. [8], dierent values of I2 and G2 can be chosen in such a way that xe is due only to the part with
n1 = 0.5
p
p
2
14a
x0e1 2Cw I 1 teff wI 1 G1 t 2Cw I 2 teff
or that xe is only due to the part with n2 = 2,
p
2
2
x0e2 2Cw I 1 teff wI 1 G1 t wI 2 G2 t

14b

Integrating Eqs. (13), (14a) and (14b) gives


!
r1
r2
X
X
1
0
0
x i
xe2 i
xe
r1 r2 i1 e1
i1
and
"

1

p11r2
r 1
2 1r21
r1
xe 2Cw I 2 teff
wI 2 G2 t

#
15

From Eqs. (1), (3), (7) and (17), one can obtain
"
1

p 1r2
r2 1
ln1  x 2Cw I 02 1r1 wI 02 G02 1r1
0
 exp @

1
QG QN
1r2 =r1 1

1
r
2 1r2
1

t  t


RT
1

t 1rr21

1

1
1r2 =r1



1
Q
2 N

1
A

#
16

F. Liu, G. Yang / Acta Materialia 55 (2007) 16291639

 
U
QN
exp
2 3
RT
G2 t


p
2
4C p

U 2Cw I t  t 1 

I Gt  t 3=2
NR
3NR

where

I 02



 132
p
2
N QG
t
w I 01 t  t 1=2 2Cw N 01 G01 exp  1=2QRT
A5
p
I 02 4@
w t
20

17a
3



exp
1

 5
G02 4 

1=2
2Ct p

I 01 t  t 2CI 01 G01 exp  1=2QN QG t2
1=2QN QG
RT

RT

2
3
IG2 t
3NR

With reference to Ref. [8], the kinetic parameters for the


current transformation can be given as

17b

With reference to Ref. [8], the kinetic parameters for the


current transformation can be given as
1
2
3=2

n 
 1 2 
r2
1

r2 =r1
2 1 r1
1 r2

n3
18a

r1

 1r12 =r1

20a

1 r4 =r3

1 1 n2 Q2

n1 Q 1

3
2

(see Fig. 3(b))

(see Fig. 2(b))


1
r
1r2

1633

r2
r1

1 B
1 1 QG
r @
1r4
r
3
1 r2

Q QN
Q1 G
;
2

Q2 QN
"

1
p 11r2
r 1
1r2
n
r1
k 0 2Cw I 02 wI 02 G02 1

k n0

18c

For the rNRo growth mode with continuous nucleation in


a 2-D space, Eq. (11) is obtained (see Section 2.2.3). Analogously, the whole extended fraction on the right side of
Eq. (11) can be considered to be composed of two parts:
one part with n = 3, the other with a mixture of n = 0.5
and n = 1.5. Finally, an analytical expression analogous
to Eq. (16) is obtained as

C
Q2N A

18b

1
1 1 2QG QN
1

r4
r3

20b

"

1 

p p

r 1
2
1r2

1rr43
1
2Cw
I 02 I 02 G1 t  t 1 
NR

4C

3NR

1 
!
  p
p
I 02 exp QRTN I 02 G02 t  t 1rr21 1 1rr43
t  t

2
I 02 G202

3NR

1
r 1
1 r4
3

#
20c

2
6
1 
6
!
Q  p
:
p
1 

1
6
p p

r 1
r
2
4C I 02 exp RTN I 02 G02 t t 1rr21 1rr43
6
1r4
1r2


1
3
I 02 I 02 G2 t t 1

ln1x 6 2Cw
6
t t
NR
3NR
6
4
0



2
I 02 G202
3NR

Q C
C
7
B 1r B
 
1
1
3
B 1r4 @  1 QG 2N A  1 2QG QN C
r 1 7
1r4
r2
r4
C
7
B
3
1
1
1
1

3
3
2

t
1 r
1 r
r4
r 1 2 r2
C
7
1
3
exp B
C t t 1r21 1r1 1r3
7
B
C
7
B
RT
C
7
B
C
7
B
A
5
@

1
r 1
1 r4
3

19
where

where

 
t  t n1
QG
p

G02
exp
n1 1
3
RT

Ut  t  t

2G1 teff
r2
;
3wNR  2 r1
3

h
For n1 see Appendix.

2
I G2 t
r
3NR 1 1
i 4
p
p


3=2
2
4C
r3
I 1 G1 t  t
2Cw I 1 t  t 1  NR
3NR

1634

F. Liu, G. Yang / Acta Materialia 55 (2007) 16291639


1

a
0.8

transformed fraction, f

transformed fraction, f

0.8

t*3=1000s

0.6

0.4
*
t2=500s

0.2

0.6

*
3

0.4

t =180s

0.2

t2 =140s

t1=100s
0
0

*
1

t =100s

2000

4000

6000

100

transformation time, s

200

300

400

500

600

transformation time, s

t*1

t*3

t2

t*

t2

t*3

Avrami exponent, n

Avrami exponent, n

2.5
1.5

1.5

0.5
0

0.2

0.4

0.6

0.8

transformed fraction, f
Fig. 2. The transformed fraction, f, as a function of transformation time, t
(a) and the Avrami exponent, n, as a function of f (b) for pure continuous
nucleation and no growth mode: N0 = 4 107 (s1 m3), QG = 1.2 105
(kJ mol1), QN = 2 105 (kJ mol1), d/m = 1, m0 = 1 105 (m s1),
T = 900 K. In (a), the thick (overlapping) solid lines represent the fraction
transformed within (0! t1 ), (0! t2 ) and (0! t3 ), and the thin solid, thick
dashed and thick dashdotted lines represents the fraction transformed
within (t1 ! t), (t2 ! t) and (t3 ! t), respectively.

The detailed derivation of Eq. (11) is given in the


Appendix.
4. Discussion
4.1. Eects of anisotropic growth, extended fraction and
Avrami exponent
For all the nucleation modes considered, the extended
volume can be shown to be always given by the addition
of two parts [8]: one part that can be conceived as due to
pure site saturation and the other that can be conceived
as due to pure continuous nucleation. By extensive calculations, the following explicit analytical expressions for the
extended transformed fraction can be obtained [21,23]:

0.5
0

0.2

0.4

0.6

0.8

transformed fraction, f
Fig. 3. The transformed fraction, f as a function of transformation time, t
(a) and the Avrami exponent, n, as a function of f (b) for pure continuous
nucleation and rRNo growth mode: N0 = 1 105 (s1 m3), QG =
1.1 105 (kJ mol1), QN = 1 105 (kJ mol1), d/m = 2, m0 = 1.8 105
(m s1), T = 740 K. In (a), the thick (overlapping) solid lines represent the
fraction transformed within (0! t1 ), (0! t2 ) and (0! t3 ), and the thin
solid, thick dashed and thick dashdotted lines represents the fraction
transformed within (t1 ! t), (t2 ! t) and (t3 ! t), respectively.

(i) for isothermal transformation,





ntQt
nt nt
xe k 0 t t exp 
RT
(ii) for isochronal transformation,

!
 2 nT
nT QT
nT RT
exp 
xe k 0 T
RT
U

21

22

Explicit expressions for n, Q and k0, in terms of general


nucleation and growth mechanisms, for both isothermal
and isochronal annealings have been given in Refs.
[8,21,23]. The kinetic parameters (n, Q and K0) are determined by the extended fraction, as has been demonstrated

F. Liu, G. Yang / Acta Materialia 55 (2007) 16291639

for mixed nucleation using the analytical phase transformation model [8,21,23,25]. The cases of site saturation
and continuous nucleation follow directly from Eqs. (21)
and (22):
(i) for isothermal transformation [6,24]



nQ
xe k n0 tn exp 
bn
RT
Q
with b k 0 t exp RT
[6], and
(ii) for isochronal transformation,
Z t


n
Q
n
xe k 0
exp 
dt bn
RT
t
0

23

24

with


RT 2
Q
b
k 0 exp 
RT
UQ
from Ref. [24].
For both isothermal and isochronal transformations, n
is the constant Avrami exponent and Q is the constant
eective activation energy (see Table 1). Combining Eq.
(23) with Eq. (4) gives
ln ln1  f n lnk 0 

nQ
n ln t
RT

25

In a classical analysis (e.g. as in Eq. (8) where n, Q and k0


are
constant
upon
isothermal
transformation)
ln(ln(1  f)) is plotted as a function of ln(t), resulting in
straight lines with slopes of n = 2, 3 or 4 in the case of continuous nucleation and linear growth (1-, 2- or 3-D) (see
Table 1).
Since Eqs. (9) and (11) follow non-JMA kinetics, the
Avrami method cannot be used to deduce the Avrami
exponent (see Section 2.2.4). Numerical calculations based
on I0, G0,QG, QN and T can be performed to construct
phase transformations under the eects of anisotropic
growth. Note that, for numerical calculations, terms such
as 2Cw and w in Eq. (9) and 2Cw(1-2/NR), 4C/3NR and
2/3NR in Eq. (11) are incorporated with I0 and G0. In
the present work, combining Eq. (16) and (19) with Eqs.
(18a) and (20a) gives the relationship of ft and nf, as
shown in Figs. 2 and 3, respectively.
With reference to [17], the eects of anisotropic growth
are characterized by the value of t*, i.e. the time step when
the transition from JMA to blocking behavior occurs. Larger t* implies smaller eects of anisotropic growth and in
turn, higher transformation rates, in accordance with MC
simulations [17].
The Avrami exponent from Eq. (18a) is determined
mainly by r2/r1, the ratio between the two extended fractions originating from JMA and blocking behavior. If
t < t*, then the value of r2/r1 tends to be innite, and n is
equal to 2. As shown in Fig. 2(a), the thick (overlapping)
solid lines within 0 ! t1 , 0 ! t2 and 0 ! t3 represent
JMA-like transformations (i.e. n = 2; see Fig. 2(b)). There-

1635

after, the transformation is blocked due to the eects of


anisotropic growth. For t > t*, the extended fraction originating from blocking behavior increases with transformation, thus decreasing n. Particularly if t  t*, the value of
r2/r1 tends to be zero, and n is close to 0.5 (see Fig. 2(b)).
For suciently large t*, the value of r2/r1 (within t* ! t)
is limited, and thus n decreases almost linearly with f (see
the nf curve with t3 1000 s in Fig. 2(b)). In contrast,
for large eects of anisotropic growth (e.g. t* = 100 s), an
initial strong decrease of n occurs after which a nearplateau is reached where n hardly varies with f (see the
nf curve with t1 100 s in Fig. 2(b)).
In terms of Eq. (20a), the Avrami exponent is determined by both r4/r3, the ratio between the two extended
fractions contributed from JMA and blocking behavior,
and r2/r1, the ratio between the two kinds of blocking
behavior. If t < t*, then the value of r4/r3 tends to be innite, and n is equal to 3 (see Fig. 3(a) and (b)). For t > t*,
the increase in r2/r1 and the decrease in r4/r3 result in a specic evolution of n with f. As shown in Fig. 3(b), if t > t*, a
continuous decrease of n with f occurs, and after a minimum value is reached, a continuous increase of n occurs
until the end of the transformation. Note that the minimum value of n shifts to higher values of f with increasing
t*, i.e. with the decreasing eects of anisotropic growth. If
t  t*, then the value of r4/r3 tends to zero, and n is equal
to 3/2-1/(1+ r2/r1) (see the nf curve with t3 in Fig. 3(b)).
Only if the value of r2/r1 becomes innite with t* ! t is a
nal value of n = 1.5 achieved. Otherwise, a nal value of
n < 1.5 results.
In combination with Section 3, it is concluded that the
variations of the Avrami exponent are consistent with the
extended fraction. The eects of anisotropic growth leads
to an evolution of the extended fraction from JMA (Eq.
(8)) to non-JMA behaviors (Eqs. (9) and (11)), changing n
accordingly. These phenomena were also observed by MC
simulations [17,26,27]. Therefore, the Avrami exponents
deduced from the analytical description on the basis of
MC simulations (i.e. Eqs. (8), (9) and (11)) provide physically realistic explanations for deviations from JMA kinetics due to the eects of anisotropic growth.
4.2. Future application of the analytical description
4.2.1. Dilatometric experiments and analysis
The preparation of alloy samples (Fe2.13 at.% Mn)
and the calibration procedure for the raw experimental
data are described in Ref. [28]. The sample was heated from
room temperature up to 1223 K at a rate of 20 K min1
and kept at this temperature for 30 min. Then it was cooled
down (at a rate of 20 K min1) to room temperature. Uniaxial compressive loads were applied using a servohydraulic system at 1173 K and were cancelled at 873 K during the
continuous cooling of the alloy samples.
The length change recorded for samples under applied
loads during cooling is shown in Fig. 4. From the experimental data, the fractions of c and a phases (e.g. fa) and

1636

F. Liu, G. Yang / Acta Materialia 55 (2007) 16291639


1

12

a
10
0.8

8
0.6

l, *105 m

6
4

0.4

2
0

0.2

2
0

400

600

800

1000

1200

20

60

80

100

0.05

Fig. 4. Measured length changes of Fe2.13 at.% Mn during continuous


cooling (20 K min1) from 1223 K to room temperature. Uniaxial
compressive (e.g. 150 N) load was applied at 1173 K and cancelled at
873 K.

experimental
optimization

b
0.04

0.03

df/dt

the transformation rate, dfa/dt, during c ! a transformation can be calculated using a lever rule [29].

40

transformation time, s

Temperature, K

0.02

4.2.2. JMA and non-JMA transformation behaviors


On the basis of the obtained dilatometric experimental
results, the alloy samples can be categorized into two types
(i.e. A and B), depending on their initial grain size and initial grain size distribution [28]. In the present work, we
focus on the transformation behavior under the eects of
anisotropic growth and illustrate the possible future application of the analytical description (see Sections 2.2.3 and
3). The c ! a transformation mechanism under applied
loads, including the inuence of the initial grain size and
initial grain size distribution, will be published elsewhere
[28].
From the dilatometric measurements performed for type
A samples under an 150N applied load, the evolution of fa
with t and the evolution of dfa/dt with fa, are shown in
Fig. 5(a) and (b), respectively. Analogous results performed for type B samples under 50, 150 and 400 N applied
loads are shown in Fig. 6(a)(d).
As shown in Fig. 5(a) and (b), a usual S-shaped curve is
observed for fa,and dfa/dt exhibits only one maximum
peak. The optimization procedure (i.e. for Fig. 5(b)) is
explained in Ref. [28]. Given the same impingement mode,
the peak maximum of dfa/dt occurs at the same fvalue,
irrespective of the nucleation and growth modes considered, and the applied annealing temperature or applied
heating rate [21]. From Fig. 5(b), it is shown that the peak
maximum in dfa/dt vs. fa corresponds to fa = 0.540.58.
Within the experimental uncertainty, the gammaalpha
transformation for type A can be concluded to follow
JMA kinetics, i.e. random nucleation and isotropic growth
[21,23,25, 28,30].

0.01

0
0

0.2

0.4

0.6

0.8

Fig. 5. The ferrite fraction, fa as a function of time, t (a), and the


transformation rate, dfa/dt, as a function of fa(b) as determined from the
length change measured for Fe2.13 at.% Mn (A) samples under 150 N
applied load with a cooling rate of 20 K min1.

For type B samples subjected to 50, 150 and 400 N


applied loads, unusual S-shaped curves are observed for
the evolution of fa in c ! a transformations (see
Fig. 6(a)(c)). From Fig. 6(a), (b) and (d), a single peak
maximum of dfa/dt corresponds to fa  0.540.58,
whereas double or multi-peaks in dfa/dt vs. fa can be
inferred from Fig. 6(c), illustrating that c ! a transformations of type B do follow non-JMA kinetics.
From Eq. (7), it is implied that a high value of e gives
high anisotropy, corresponding to the peak maximum in
dfa/dt vs. fa at small fa reected in the relation between f
and xe in Fig. 1 and by a comparison between Figs. 5(b)
and 6(d). However, Eq. (7), as a phenomenological description, cannot provide a physically realistic explanation for
the deviation due to the eects of anisotropic growth.
Actually, the analytical phase transformation model
(Section 3) based on MC simulations deals only with isothermal transformations (see Figs. 2 and 3). The c ! a
transformations as shown in Figs. 5(a) and 6(a)(c) are
conducted isochronally. Therefore, the current analytical

F. Liu, G. Yang / Acta Materialia 55 (2007) 16291639


1

1637

c
0.8

0.6

0.6

t*

0.8

0.4

0.4

0.2

0.2
*

t
0
0

20

40

60

0
0

80

20

40

60

80

100

120

transformation time, s

transformation time, s
0.2

experimental
optimization

b
0.8
0.15

df /dt

0.6

0.1

0.4
0.05

0.2
*

t
0
0

20

40

60

80

0
0

0.2

0.4

0.6

0.8

transformation time, s

Fig. 6. The ferrite fraction, fa, as a function of time, t, as determined from the length change measured for Fe2.13 at.% Mn (B) samples under 50 (a),
150 (b) and 400 N (c) applied load with a cooling rate of 20 K min1. The transformation rate, dfa/dt, as a function of fa (d) as determined from the
length change measured for Fe2.13 at.% Mn (B) samples under 50 N applied load with a cooling rate of 20 K min1.

description cannot be adopted to describe the deviations


due to the eects of anisotropic growth. However, a
detailed comparison between Figs. 6(a)(c), 2(a) and 3(a)
demonstrates that all the fat curves are analogous, i.e.
they are composed of two segments indicative of two transformation behaviours (JMA and blocking behaviors). With
increased applied loads, the inection point between
the two segments moves to higher t and fa values (see the
evolution of t* in Fig. 6(a)(c)). In combination with Section 4.1, it is concluded that an increased load gives larger
memory time, in accordance with lower eects of anisotropic growth [17].
Thus the current analytical description based on MC
simulations does create a wider possibility to investigate
the eects of anisotropic growth. By this means, the current
analytical description can probably be applied to more general transformations, bringing new insight and understanding of phase transformations subjected to the eects of
anisotropic growth in real materials.

5. Conclusions
On the basis of MC simulations [17] of isothermal phase
transformations, the eects of anisotropic growth on deviations from the classical JMA kinetics have been
investigated.
(1) Adopting the analytical approach used to derive the
analytical phase transformation model [8], the eects
of anisotropic growth have been reinterpreted using
an analytical transformation model. In combination
with MC simulations, the eect of the memory time
when the transition from JMA to blocking behavior
occurs on the transformation has been shown, and
physically realistic Avrami exponents were deduced.
(2) In combination with MC simulations, it is concluded
that, the eects of anisotropic growth change the
extended fraction from JMA (Eq. (8)) to non-JMA
(Eqs. (9) and (11)), thus leading to a change of n.

1638

F. Liu, G. Yang / Acta Materialia 55 (2007) 16291639

(3) A comparative study, performed between the analytical description and the gammaalpha transformation occurring in a FeMn alloy under applied
loads, shows that, based on MC simulations, the
present analytical phase transformation model creates a wider possibility to investigate the eects of
anisotropic growth. By this means, the current analytical description can probably be applied to more
general transformations, bringing new insight and
understanding to phase transformations subjected
to the eects of anisotropic growth in real materials.
Acknowledgements
The authors are grateful for the nancial support of New
Century Excellent Person Supporting Project (NCET-05870), the Fundamental Research Project of National Defense of China (A2720060295), the Project Sponsored by
the Scientic Research Foundation for the Returned Overseas Chinese Scholars, State Education Ministry
(N6CJ0002), the Scientic and Technological Creative
Foundation of Youth in Northwestern Polytechnical University, and the Natural Science Foundation of China
(Grant No. 50501020, 50395103 and 50431030). F. Liu is
also grateful to Prof. E.J. Mittemeijer and Prof. F. Sommer
for their valuable and essential instruction and cooperation.

where



p p

p
2

0
2Cw
I 01 I 01 G1 t  t 1 
I 02
NR
"
p
I 020 G020

A4
!#
QG  p
p
4C I 01 exp RT I 01 G01 t  t
3NR
t  t
A5

Analogously, dierent values of I02 and G02 can be chosen


in such a way that xe is only due to the part with n2 = 3.
Combining Eqs. (A2), (A4) and (A5) gives
2
 1r
6
p p

1r2
2
6

1
I 01 I 01 G1 t  t 1 
:
6 2Cw
4
NR
! 1 1
  p
p
I 01 exp QRTG I 01 G01 t  t 1rr21

t  t
3
1
QN
1 1 QG
2
r2
3 1 7
B
C
1 r
2 1r2 7
1
B
C
 exp B
Cteff r1 7
5
@
A
RT
4C
3NR
0



2
2Q QN  3
I 01 G201 exp  G
t
3NR
RT


2
2Q QN  3
I 02 G202 exp  G

t ;
3NR
RT

Appendix
First, the mixed part (in Eq. (11)) is considered. For this
part, the ratio of the extended fraction with n = 1.5 to that
with n = 0.5 is given as
2Gteff
r2

3wNR  2t r1

A1

Analogous to the treatment performed for Eq. (9), Eq. (11)


can be rewritten as
"

 1r
p
1r2
2
1
0
xe 2Cw I 2 1 
NR
#

 1 1
3 1
r
2 1r2
4C p
2
1r2
3
r
1
1
I 20 G20
IG2 t
teff


3NR
3NR
A2
The whole extended fraction can be considered to be composed of two parts: xe = xe1 + xe2, one part with n = 3, the
other with n1 32  11r2 . The ratio between the two
r1

extended fractions is given as


"

2
I G2 t 3
3NR 1 1

r r
11r2  4C p 1rr2 1
p
1 2 1
2
r1
1
2Cw I 20 1  NR
teff 2r2 r1
I 20 G20
3NR

r4
r3

A3

or that xe is only due to the part with n1,


2
 1r
6
p p

1r2
2
6

1
I 01 I 01 G1 t  t 1 
6 2Cw
4
NR
! 1 1
  p
p
I 01 exp QRTG I 01 G01 t  t 1rr21

t  t
3
1
QN
1
 1 QG
2
r
3 1 7
B
C
1 r2
2 1r2 7
1
B
C
 exp B
Cteff r1 7
5
@
A
RT
4C
3NR
0



2
2Q QN  3

I 01 G201 exp  G
t
RT
2 3NR
 1r
6
p p

1r2
2
6
1
6 2Cw
I 02 I 02 G02 t  t 1 
4
NR
! 1 1
  p
p
I 02 exp QRTG I 02 G02 t  t 1rr21

t  t
1
3
QN
1
 1 QG
2
r
3 1 7
B
C
1 r2
2 1r2 7
1
B
C
 exp B
Cteff r1 7
@
A
5
RT
4C
3NR
0

F. Liu, G. Yang / Acta Materialia 55 (2007) 16291639

if,
2
 1r
6
p p

1r2
2
6
1
I 01 I 01 G1 t  t 1 
6 2Cw
4
NR
! 1 1
  p
p
I 01 exp QRTG I 01 G01 t  t 1rr21

t  t
3
1
QN
1 1 QG
2
r2
3 1 7
B
C
1 r
2 1r2 7
1
B
C
 exp B
Cteff r1 7
5
@
A
RT
4C
3NR
0



2
2Q QN  3
I 01 G201 exp  G
t U
3NR
RT

then Eq. 19 can be obtained. Finally, the kinetic parameters are deduced, as given by Eqs. (20a), (20b) and (20c).
References
[1] Johnson WA, Mehl RF. Trans Am Inst Min (Metall) Engs 1939;
135:1.
[2] Avrami M. J Chem Phys 1939;7:1109.
[3] Avrami M. J Chem Phys 1940;8:212.
[4] Avrami M. J Chem Phys 1941;9:177.

1639

[5] Christian JW. The theory of transfomation in metals and alloys, Part
1 equilibrium and general kinetics theory. Oxford: Pergamon Press;
1975.
[6] Mittemeijer EJ. J Mater Sci 1992;27:3977.
[7] Kempen ATW, Sommer F, Mittemeijer EJ. J Mater Sci 2002;37:1321.
[8] Liu F, Sommer F, Mittemeijer EJ. J Mate Sci 2004;39:1621.
[9] Weinberg MC, Kapral R. J Chem Phys 1989;91:7146.
[10] Bradley RM, Strenski PN. Phys Rev B 1989;40:8967.
[11] Kashchiev D. Surf Sci 1969;14:209.
[12] Kelton KF, Greer AL, Thompson CV. J Chem Phys 1983;79:6261.
[13] Andrienko YA, Brilliantov NV, Krapivsky PL. Phys Rev A 1992;
45:2263.
[14] Sekimoto K. Physica A 1986;135:328.
[15] Sekimoto K. Phys Lett A 1984;105:390.
[16] Pusztai T, Granasy L. Phys Rev B 1998;57:14110.
[17] Kooi BJ. Phys Rev B 2004;70(22):224108.
[18] Ruitenberg G, Petford-Long AK, Doole RC. J Appl Phys 2002;92:
3116.
[19] Liu F, Sommer F, Mittemeijer EJ. J Mater Sci. Available from: <http:
//dx.doi.org/10.1007/s10853-006-0802-4>.
[20] Ranganathan S, Heimendahl M. J Mater Sci 1981;16:2401.
[21] Liu F, Sommer F, Mittemeijer EJ. Int Mater Rev, in press.
[22] Starink MJ. J Mater Sci 2001;36:4433.
[23] Liu F, Sommer F, Mittemeijer EJ. J Mater Res 2004;19:2586.
[24] Kempen ATW, Sommer F, Mittemeijer EJ. Acta Mater 2002;50:1319.
[25] Liu F, Sommer F, Mittemeijer EJ. Acta Mater 2004;52:3207.
[26] Shneidman VA, Weinberg MC. J Non-Cryst Sol 1993;160:89.
[27] Weinberg MC, Birnie DP, Shneidman III VA. J Non-Cryst Sol 1997;
219:89.
[28] Liu F, Sommer F, Mittemeijer EJ, unpublished work.
[29] Liu YC, Sommer F, Mittemeijer EJ. Acta Mater 2003;51:507.
[30] Malek J. Thermochimica Acta 2000;355:239.

Вам также может понравиться