ArticlePDF Available

Dynamic Ferrite Transformation Behaviors in 6Ni-0.1C Steel

Authors:

Abstract

Phase transformation from austenite to ferrite is an important process to control the microstructures of steels. To obtain finer ferrite grains for enhancing its mechanical property, various thermomechanical processes followed by static ferrite transformation have been carried out for austenite phase. This article reviews the dynamic transformation (DT), in which ferrite transforms during deformation of austenite, in a 6Ni-0.1C steel recently studied by the authors. Softening of flow stress was caused by DT, and it was interpreted through a true stress-true strain curve analysis. This analysis predicted the formation of ferrite grains even above the Ae3 temperature (ortho-equilibrium transformation temperature between austenite and ferrite), where austenite is stable thermodynamically, under some deformation conditions, and the occurrence of DT above Ae3 was experimentally confirmed. Moreover, the change in ferrite grain size in DT was determined by deformation condition, i.e., deformation temperature and strain rate at a certain strain, and ultrafine ferrite grains with a mean grain size of 1 μm were obtained through DT with subsequent dynamic recrystallization of ferrite.
1 23
JOM
The Journal of The Minerals, Metals &
Materials Society (TMS)
ISSN 1047-4838
JOM
DOI 10.1007/s11837-014-0913-3
Dynamic Ferrite Transformation Behaviors
in 6Ni-0.1C Steel
Nokeun Park, Lijia Zhao, Akinobu
Shibata & Nobuhiro Tsuji
1 23
Your article is protected by copyright and all
rights are held exclusively by The Minerals,
Metals & Materials Society. This e-offprint is
for personal use only and shall not be self-
archived in electronic repositories. If you wish
to self-archive your article, please use the
accepted manuscript version for posting on
your own website. You may further deposit
the accepted manuscript version in any
repository, provided it is only made publicly
available 12 months after official publication
or later and provided acknowledgement is
given to the original source of publication
and a link is inserted to the published article
on Springer's website. The link must be
accompanied by the following text: "The final
publication is available at link.springer.com”.
Dynamic Ferrite Transformation Behaviors in 6Ni-0.1C Steel
NOKEUN PARK,
1
LIJIA ZHAO,
2
AKINOBU SHIBATA,
1,2
and NOBUHIRO TSUJI
1,2,3
1.—Elements Strategy Initiative for Structural Materials (ESISM), Kyoto University, Yoshida
Honmachi, Sakyo-ku, Kyoto 606-8501, Japan. 2.—Department of Materials Science and Engi-
neering, Kyoto University, Yoshida Honmachi, Sakyo-ku, Kyoto 606-8501, Japan. 3.—e-mail:
nobuhiro-tsuji@mtl.kyoto-u.ac.jp
Phase transformation from austenite to ferrite is an important process to
control the microstructures of steels. To obtain finer ferrite grains for
enhancing its mechanical property, various thermomechanical processes fol-
lowed by static ferrite transformation have been carried out for austenite
phase. This article reviews the dynamic transformation (DT), in which ferrite
transforms during deformation of austenite, in a 6Ni-0.1C steel recently
studied by the authors. Softening of flow stress was caused by DT, and it was
interpreted through a true stress–true strain curve analysis. This analysis
predicted the formation of ferrite grains even above the Ae
3
temperature
(ortho-equilibrium transformation temperature between austenite and fer-
rite), where austenite is stable thermodynamically, under some deformation
conditions, and the occurrence of DT above Ae
3
was experimentally confirmed.
Moreover, the change in ferrite grain size in DT was determined by defor-
mation condition, i.e., deformation temperature and strain rate at a certain
strain, and ultrafine ferrite grains with a mean grain size of 1 lm were
obtained through DT with subsequent dynamic recrystallization of ferrite.
INTRODUCTION
One of the most effective ways to enhance
strength without losing toughness in metallic
materials is grain refinement, especially in steels.
The fundamental idea to obtain finer ferrite grains
in low-carbon steels is to use phase transformation
phenomenon from austenite (c: high-temperature
phase) to ferrite (a: low-temperature phase). Many
innovative techniques have been developed to
achieve fine ferrite grains, for example, thermome-
chanical controlled processing (TMCP)
1
. Figure 1a
is a schematic illustration showing a continuous-
cooling-transformation (CCT) diagram and the
change in microstructures by TMCP in low-C steels.
Coarse austenite grains are first deformed at
recrystallization temperatures of austenite, result-
ing in fine austenite grains through recrystalliza-
tion of austenite. As a result, the austenite grain
boundary area per unit volume, which is one of the
effective nucleation sites for ferrite transformation,
increases. Once austenite is deformed at a nonre-
crystallization region, a large amount of lattice
defects, such as dislocations, which can be potential
nucleation sites for ferrite transformation, are
introduced into the austenite. Finally, ferrite
transformation occurs at a relatively lower tem-
perature by accelerating the cooling process, where
the nucleation rate of ferrite is enhanced by a large
driving force (under large supercooling). Through
the process explained in Fig. 1a, fine ferrite with
mean grain size of around 58lm could be obtained
through the TMCP.
Recent developments of facilities’ capacity (such
as that in rolling mills) can make it possible to apply
a larger amount of one-pass deformation at a much
lower temperature. As a result, ultrafine ferrite
structures with a mean grain size of approximately
1lm have been achieved.
25
As shown in a sche-
matic illustration of Fig. 1b, ferrite transformation
could occur during the plastic deformation of aus-
tenite that is conducted in larger one-pass defor-
mation at lower temperatures because the kinetics
of ferrite transformation are greatly accelerated by
deformation. The ferrite transformation occurring
during deformation of austenite has been termed as
dynamic (ferrite) transformation (DT), strain-induced
ferrite transformation,
612
deformation-induced
JOM
DOI: 10.1007/s11837-014-0913-3
Ó2014 The Minerals, Metals & Materials Society
Author's personal copy
ferrite transformation,
1317
and so on, and it has
received great attention as another mechanism of
ferrite grain refinement.
One controversial issue about DT is whether the
DT occurs at temperature above Ae
3
(ortho-equi-
librium transformation temperature between aus-
tenite and ferrite), where ferrite is unstable so that
austenite single phase should be stable thermody-
namically. Choi et al.
9
denied the possibility of the
occurrence of DT at temperatures near or above Ae
3
because the critical strain for the occurrence of DT
was higher than that of dynamic recrystallization of
austenite so that dynamic recrystallization of aus-
tenite preferentially occurred compared to DT dur-
ing deformation of austenite. On the other hand,
Yada et al.
2
, who were pioneers of studies on DT,
proposed a strain-temperature-phase map and in-
sisted that DT took place even above Ae
3
when
plastic deformation of austenite was large enough.
Recently, we systematically studied a thermome-
chanical process under various deformation condi-
tions, i.e., various deformation temperatures and
strain rates, for clarifying the occurrence of DT in a
6Ni-0.1C steel, and it was revealed that DT could
take place even at temperatures higher than Ae
3
when the combination of strain rate and deforma-
tion temperature was appropriately selected.
1820
Therefore, we discuss the deformation condition for
the occurrence of DT as the first topic in this article.
Another important subject of DT is the ferrite
grain refinement by DT depending on deformation
conditions, such as deformation temperature, strain
rate, austenite grain size, and so on. There have
been several reports discussing grain refinement
mechanisms by DT:
2,5,11,2124
transformation from
dynamically recrystallized austenite, continuous
nucleation of ferrite during orientation rotation of
austenite, dynamic recrystallization of ferrite during
DT, etc. However, a comprehensive understanding
has not been clearly revealed yet because it is hard to
consider the effect of each contribution in DT inde-
pendently. In the second section of this report, we try
to discuss the effects of deformation parameters on
the grain size of dynamically transformed ferrite.
EXPERIMENTAL PROCEDURES
Material
The material used in this study was a 6Ni-0.1C
steel (weight percent), which was designed to widen
austenite temperature range and to have a long
incubation time for the onset of static ferrite
transformation at given temperatures. The detailed
composition of the material is shown in Table I.
The Ae
3
and Ap
3
(paraequilibrium transformation
temperature between austenite and ferrite) in the
6Ni-0.1C steel are calculated by Thermo-Calc soft-
ware (Thermo-Calc Software, Inc., McMurray, PA,
USA) to be 728°C and 684°C, respectively.
18
Thermomechanical Process
As-received plates were homogenized at 1100°C
for 24 h, followed by water quenching. Cylindrical
specimens 12 mm in height and 8 mm in diameter
were cut from the homogenized plate by electric
discharge machining. Physical simulations of a
thermomechanical process were carried out using
a thermomechanical processing simulator Therm-
acmastor-Z (Fuji Electronic Industrial Co., Ltd.,
Saitama, Japan). Here, two austenitization condi-
tions were chosen to have different austenite grain
sizes; 1200°C for 180 s for obtaining coarse aus-
tenite grains (mean grain size of 400 lm) and
800°C for 180 s for having fine austenite grains
(mean grain size of 15 lm). The austenitized
Fig. 1. Schematic illustrations showing CCT diagrams and change in microstructures for (a) static ferrite transformation from deformed austenite
and (b) dynamic ferrite transformation during deformation of austenite.
Park, Zhao, Shibata, and Tsuji
Author's personal copy
specimens were cooled to deformation temperature
by N
2
gas at a cooling rate of 30°Cs
1
and then
kept for 60 s to ensure uniform temperature dis-
tribution within the specimen. It has been con-
firmed that no static ferrite transformation
occurred during holding for 60 s at the whole
temperature range in the case of coarse austenite
grain size (see the time–temperature-transforma-
tion diagram of this alloy shown in Ref. 18). An
uniaxial compression was applied to the austeni-
tized specimens at four constant strain rates
ranging from 10
3
s
1
to 10
0
s
1
to a true strain of
0.96. After completing the deformation, the speci-
mens were immediately quenched by a precisely
controlled water-injection system.
To study the deformation behavior of ferrite at
elevated temperatures, one homogenized plate was
slowly cooled in the furnace to obtain a mixed
microstructure composed of ferrite and a small
amount of pearlite. The specimens composed of
ferrite and pearlite were heated up to the defor-
mation temperature ranging from 400°C to 700°C
(below Ae
3
), followed by an isothermal holding for
60 s, and then an uniaxial compression was applied
at a strain rate of 10
1
s
1
. True stress–true strain
data were obtained from the load–displacement
data taken during the hot deformation of both aus-
tenite and ferrite.
Microstructural Observation
Microstructures of the deformed specimens
were observed by optical microscopy (OM), elec-
tron backscattering diffraction (EBSD) analysis
in a field-emission scanning electron microscope
(FE-SEM; FEI XL30S FEG; FEI Company,
Hillsboro, OR, USA), and transmission electron
microscopy (TEM; Philips CM200 FEG; Philips,
Amsterdam, The Netherlands). The cross-sections
parallel to the compression direction of the deformed
specimens were mechanically polished and then
electropolished in a solution of 10% HClO
4
and
90% CH
3
COOH at 25°C. A 3% nital solution was
used to reveal microstructures for OM. The mean
grain sizes of ferrite were measured by the linear
interception method on the EBSD boundary
maps, which were similar to those presented in
our previous article.
18
Boundaries with misori-
entation above 2°were used for the measure-
ment. Thin foils for TEM observations were
prepared by twin-jet electropolishing using the
identical solution as that used for the EBSD
observations.
RESULTS AND DISCUSSION
Analysis of True Stress–True Strain Curves
for the Occurrence of DT
Because ferrite is softer than austenite at elevated
temperatures in low-carbon steels,
25
the occurrence of
DT influences the flow stress, which would decrease
with increasing the fraction of dynamically trans-
formed ferrite. Whether DT occurs or not can be de-
tected through the analysis of stress–strain behaviors
during hot deformation. We examined the change in
true stress–true strain behaviors of the specimens
deformed at various deformation temperatures and
strain rates. Figure 2a shows the true stress–true
strain curves of the specimens deformed at a strain
rate of 10
2
s
1
at various temperatures after au-
stenitization at 1200°C for 180 s. The mean grain size
of austenite was 400 lm. The shape of the stress
strain curves can be classified into three different
types. At higher temperatures, the curves show the
shapes typical for the dynamic recrystallization of
austenite. With increasing the amount of strain, the
flow stress increases until the peak stress is reached.
After attaining the peak stress, the flow stress de-
creases and then keeps a constant value. At interme-
diate temperatures, the curves show the dynamic
recovery type where the flow stress initially increases
and then keeps a constant value without showing any
peak and stress drop. At lower temperatures, the
stress–strain curves exhibit a significant softening
again after the peak stress. The shape of these curves
is similar to that for dynamic recrystallization of
austenite at higher temperatures. It has been con-
firmed that the observed softening at lower tempera-
tures is attributed to the occurrence of DT.
9
Figure 2b
summarizesthe change in the maximum flow stress of
the specimens deformed at four different strain rates
as a function of deformation temperature. It should be
noted that the softening of the maximum flow stress
due to the occurrence of DT is significant when the
strain rate is slow, for example 10
3
or 10
2
s
1
,and
the deformation temperature is low.
The Zener–Hollomon (Z) parameter, described in
Eq. 1, is a well known parameter that unifies the
effects of deformation temperature and strain rate
in high-temperature deformations,
26
Z_
eexp Q=RTðÞ (1)
Table I. Chemical composition of the 6Ni-0.1C steel used in the current study (wt.%)
CNiMnSi Al P S Fe
0.112 5.97 0.003 0.008 0.037 0.002 0.0015 bal.
Dynamic Ferrite Transformation Behaviors in 6Ni-0.1C Steel
Author's personal copy
where _
eis the strain rate, Qis the activation energy
for high-temperature deformation of austenite, Ris
the gas constant, and Tis the absolute temperature.
The constitutive equation for the hot deformation of
metallic materials, which can cover wide range of Z
parameter, is described in Eq. 2.
27
Z¼Asinh arðÞ (2)
where Aand aare constants, and ris the maxi-
mum flow stress of the true stress–true strain
curve in this study. The maximum flow stress val-
ues at higher temperatures in Fig. 2b were plotted
as a function of temperature in Fig. 2c to obtain the
activation energy for the high-temperature defor-
mation of austenite, and the estimated value was
310 kJ mol
1
. The maximum flow stress values in
Fig. 2b are plotted as a function of Zparameter in
Fig. 2d.AsshowninEq.2, the maximum flow
stress in a form of hyperbolic sine function is
proportional to the Zparameter when the Z
parameter is low, i.e., at high temperatures or at
low strain rates, which is generally accepted as a
typical hot-deformation behavior of metallic mate-
rials.
27
However, the maximum flow stress in a
form of hyperbolic sine function gradually deviates
from the extrapolated line when the Zparameter
is higher than a critical value (1.7 910
14
s
1
).
Because the deviation of the maximum flow stress
from the extrapolated line above the critical Z
parameter was attributed to the occurrence DT, the
deformation conditions for the occurrence of DT
could be described by a combination of deformation
temperature and strain rate. Table II shows the
deformation parameters above the critical Z
parameter, indicating that DT could occur even
above Ae
3
(728°C) when the strain rate is high
enough.
It is also found from Fig. 2d that the deviation of
the maximum flow stress from the extrapolated line
Fig. 2. Procedures of the true stress–true strain analysis for confirming the occurrence of dynamic ferrite transformation.
18
(a) True stress–true
strain curves of the specimens deformed at a strain rate of 10
2
s
1
at various temperatures. (b) Change in the maximum flow stress at different
strain rates as a function of the deformation temperature. (c) Determination of the activation energy (310 kJ mol
1
) for calculating the Zener–
Hollomon parameter about austenite deformation. (d) Change in the maximum stress in a form of hyperbolic sine function as a function of the
Zener–Hollomon parameter. Z
C
indicates the critical Zener–Hollomon parameter for the occurrence of stress softening.
Park, Zhao, Shibata, and Tsuji
Author's personal copy
in the lower Zregion is more significant at lower
strain rates. In the case of a lower strain rate, the
deformation time itself, which is another important
parameter describing kinetics of DT, becomes much
longer so that a larger fraction of ferrite can form
during deformation, resulting in the significant
deviation of the maximum flow stress.
In order to confirm the occurrence of DT above
Ae
3
, we used the specimen having finer austenite
grains (mean grain size of 15 lm) to accelerate
kinetics of DT. The specimens austenitized at 800°C
and the specimens having a ferrite and pearlite
microstructure were deformed at various tempera-
tures at a strain rate of 10
1
s
1
. The true stress–
true strain curves of the austenitized specimens and
those of the ferrite specimens are shown in Fig. 3a
and b, respectively. The maximum flow stress val-
ues obtained from the stress–strain curves are
plotted as a function of the deformation tempera-
ture in Fig. 3c (square: the austenitized speci-
mens, circle: the ferrite specimens). Figure 3disan
enlarged part of Fig. 3c close to Ae
3
. The maximum
flow stress values at temperatures higher than
800°C were taken from the Fig. 2b because austenite
Table II. Deformation Conditions Above the Critical Zener–Hollomon Parameter
18
Parameters Conditions
Strain rate, e(s
1
)10
0
10
1
10
2
10
3
Temperature, T(°C) <865 <790 <725 <667
Fig. 3. (a) True stress–true strain curves of the specimens deformed in austenite states.
19
(b) True stress–true strain curves of the specimens
having a ferrite and pearlite microstructure. (c) Change in the maximum stress of the austenite specimens obtained from (a) and that of the ferrite
specimens obtained from (b), as a function of the deformation temperature. (d) An enlarged area in (c). Ap
3
and Ae
3
temperatures are marked as
dashed lines in (c) and (d).
Dynamic Ferrite Transformation Behaviors in 6Ni-0.1C Steel
Author's personal copy
grain size did not influence the maximum flow
stress but only accelerated kinetics of dynamic
softening. The dash-dotted line in Fig. 3c and d is
the extrapolation of the fitting curve derived from
Eq. 2, and the dotted lines in Fig. 3d indicate upper
and lower 95% confidence intervals. It is clearly
seen that the flow stress deviates from the extrap-
olated curve at temperatures above Ae
3
in Fig. 3d,
and it comes close to the maximum flow stress of the
ferrite with decreasing the deformation tempera-
ture in Fig. 3c.
Although the analysis in Fig. 3implies that DT is
likely to occur above Ae
3
, two other possibilities
must be checked: One is the static transformation to
ferrite from deformed austenite during the cooling
procedure, and the other is that the Ae
3
calculated
by Thermo-Calc software (728°C) might be lower
than the true Ae
3
of the 6Ni-0.1C steel. Figure 4a
and b are OM images of the austenitized specimens
that were deformed at 750°C, i.e., above the calcu-
lated Ae
3
(728°C), to strains of (Fig. 4a) e= 0.33 and
(Fig. 4b) e= 0.96 at a strain rate of 10
1
s
1
. The
measured fractions of ferrite in Fig. 4a and b are
15% and 55%, respectively. Let us assume here that
DT did not happen during the deformation, but the
ferrite grains observed in Fig. 4a and b were stati-
cally transformed from the deformed austenite
during water quenching. The flow stress of the
specimen deformed to e= 0.96 (166 MPa) is lower
than that at e= 0.33 (188 MPa), as shown in
Fig. 3a. The flow stress can be regarded as a func-
tion of the dislocation density in the austenite, so
that the density of lattice defects in the austenite
deformed to e= 0.33 is expected to be higher than
that in the austenite deformed to e= 0.96. Accord-
ingly, the fraction of statically transformed ferrite
from the austenite deformed to e= 0.33 should be
larger than that that from the austenite deformed to
e= 0.96 because the larger amount of lattice defects
would accelerate the kinetics of static ferrite
transformation. However, the result shown in
Fig. 4a and b is opposite; i.e., the fraction of ferrite
at e= 0.33 (15%) is smaller than that at e= 0.96
(55%), suggesting that the observed ferrite grains
were not statically transformed during postdefor-
mation cooling but formed during deformation, i.e.,
dynamically transformed ferrite. Therefore, the
softening of the flow stress can be interpreted as the
increase in the fraction of dynamically transformed
ferrite, i.e., kinetics of DT.
20
Figure 4c is an OM image of the specimen
deformed to a strain of 0.96 and subsequently held
for 600 s at the identical temperature (750°C) after
the deformation. The fraction of ferrite decreased
from 55% in Fig. 4b to less than 2% during sub-
sequent holding at 750°C. The result demonstrates
that the dynamically transformed ferrite reversely
transformed to austenite during statically keeping
at 750°C after the deformation, and the deformation
temperature (750°C) was surely above Ae
3
. There-
fore, it can be concluded that DT can occur even
above Ae
3
when the deformation conditions are
above the critical value of the Zparameter.
Lee et al.
28
qualitatively explained how ferrite
formed during deformation of austenite at temper-
atures above Ae
3
. Once deformation is applied to
austenite, the extra energy generated by deforma-
tion is added into the austenite phase so that the
free energy of austenite increases much more than
that of ferrite. Therefore, DT can occur even at
temperatures above Ae
3
because newly formed fer-
rite is relatively stable compared to the deformed
austenite. Figure 5shows a part of the phase dia-
gram of a Fe-6Ni-C system calculated by Thermo-
Calc software. The change in Ae
3
lines with differ-
ent additional energies on austenite phase is pre-
sented in Fig. 5a. The dashed line corresponds to
the chemical composition of carbon, 0.1 wt.%, used
in this study. For the occurrence of DT at 750°C, a
large amount of additional energy in austenite
phase (50 J mol
1
) is required in Fig. 5a. Here, it is
necessary to consider the contributing factors that
increase the free energy of austenite. One major
factor is a stored energy of dislocations. Figure 5b
exhibits the required dislocation density to vary the
Ae
3
lines corresponding to the additional energy in
Fig. 5a. Here, it was assumed that the stored energy
[E(J mol
1
)] by dislocations is E¼lb2qVm, where l
Fig. 4. Optical microscope images of the austenite specimens deformed to different strains:
19
(a) e= 0.33, (b) e= 0.96 at a strain rate of
10
1
s
1
at 750°C. (c) The specimen deformed to a strain of 0.96 at 750°C and subsequently held at 750°C for 600 s. Arrows indicate ferrite
grains, of which fractions are shown in the figures as well.
Park, Zhao, Shibata, and Tsuji
Author's personal copy
is the shear modulus, bis the magnitude of Burgers
vector, qis the dislocation density, and V
m
is the
molar volume. As shown in Fig. 5b, however, the
calculated dislocation density (2 910
15
m
2
)isnot
realistic (seems too high for high-temperature
deformation of austenite). Some possible contribut-
ing factors, such as the heterogeneity of dislocation
distributions within austenite grains, austenite
grain boundary energy, elastic energy under defor-
mation, etc., have been proposed, and further
quantitative understanding about driving force for
DT above Ae
3
is required.
2934
Change in Grain Size of Dynamically Trans-
formed Ferrite
It has been reported that the grain size of dynam-
ically recrystallized ferrite grains is a function of Z
parameter,
35
and the ferrite grain size obtained by
thermomechanical process using DT is much finer
compared with that obtained through conventional
TMCP processes.
3,5
Figure 6shows the change in the
grain size of ferrite dynamically transformed from
the austenite with a mean grain size of either 15 lm
or 400 lm, which was deformed to a strain of 0.96 in
the 6Ni-0.1C steel, plotted as a function of the Z
parameter. The activation energy for determining
the Zparameter used in Fig. 6is 254 kJ mol
1
,
which is the energy for self-diffusion of iron atom in
body-centered cubic structure. The grain size of the
dynamically transformed ferrite decreased with
increasing the Zparameter (i.e., at a high strain rate
or low deformation temperature), of which Z-depen-
dence was similar to the previous reports about
dynamic recrystallization of ferrite.
35
The reason
why finer ferrite grains were obtained at higher Z
parameter could be explained in the following ways.
A higher strain rate introduces a larger number of
lattice defects in austenite, and such a large density
of defects in the deformed austenite acts as the
nucleation site for ferrite transformation. That is, the
nucleation density of dynamically transformed fer-
rite is high. In addition, when the deformation tem-
perature is lower, the growth rate of newly formed
ferrite becomes slower. Consequently, the mean
grain size of the dynamically transformed ferrite
reaches to 1 lm when the austenite grain size is fine
(15 lm) and the Zparameter is high. The result also
indicates the effect of the grain boundary density of
austenite, which plays a role of preferential nucle-
ation site for ferrite transformation. A large density
of austenite grain boundaries in the specimens hav-
ing the austenite grain size of 15 lm also enhances
the number of nucleated ferrite grains per unit
volume.
There have been several explanations about grain
refinement in DT. Cellular automaton calculation
revealed that the finer ferrite grain size in DT could
be obtained owing to continuous nucleation of fer-
rite around austenite/ferrite phase boundaries as
well as at austenite grain interiors because retained
austenite is continuously deformed.
11
Adachi et al.
22
proposed that a crystal rotation of austenite and DT
occurred simultaneously during deformation so that
the orientation of dynamically transformed ferrite
in the earlier stage of DT was different with that in
the later stage of DT. Although Hurley et al.
36
denied any possibility of dynamic recrystallization
of ferrite after DT, Matsumura and co-workers
24
insisted dynamic recrystallization of dynamically
transformed ferrite. In the current study, the rep-
resentative TEM image of dynamically transformed
Fig. 5. A part of the phase diagram in a Fe-6Ni-C system showing the change in Ae
3
lines with respect to (a) the additional energy to austenite
and (b) the corresponding dislocation density in austenite phase. The dashed line corresponds to the chemical composition of carbon, 0.1 wt.%,
used in this study.
Dynamic Ferrite Transformation Behaviors in 6Ni-0.1C Steel
Author's personal copy
ferrite shown in Fig. 7, for which the specimen was
deformed to a strain of 0.96 at a strain rate of
10
1
s
1
at 650°C and exhibited a mixed micro-
structure composed of fine and equiaxed ferrite
grains and ferrite grains elongated perpendicular to
the compression direction (double-headed arrow).
The fine equiaxed grains in Fig. 7are probably the
ferrite grains dynamically recrystallized during
subsequent deformation after DT.
37
Accordingly, it
can be considered that the refinement of ferrite
grains through DT is attributed to several factors:
the continuous nucleation of ferrite nuclei, limited
grain growth of ferrite, and dynamic recrystalliza-
tion of transformed ferrite. Subsequent systematic
research on the grain refinement mechanisms of DT
has been carried out using a 10Ni-0.1C steel in the
authors’ group, from which the results will soon be
published elsewhere.
SUMMARY AND CONCLUSION
We summarized the recent studies about dynamic
ferrite transformation in a 6Ni-0.1C steel in this
article. The major results are listed as follows:
1. The occurrence of the dynamic transformation
could be confirmed from the stress–strain ana-
lysis. The deformation conditions for the occur-
rence of dynamic transformation were clearly
determined by Zener–Hollomon parameter.
When the Zener–Hollomon parameter at a given
deformation condition was higher than the crit-
ical value (1.7 910
14
s
1
), softening of the flow
stress occurred, resulting from dynamic trans-
formation.
2. Dynamic softening of the flow stress due to
dynamic ferrite transformation was observed
even at 750°C (above Ae
3
). The fraction of the
dynamically transformed ferrite increased with
increasing the strain. The ferrite grains were
reversely transformed to austenite during sub-
sequent and static annealing at 750°C after the
deformation, which confirmed the occurrence of
the dynamic transformation above Ae
3
.
3. The grain size of the dynamically transformed
ferrite decreased with increasing the Zener–
Hollomon parameter, i.e., at lower deformation
temperatures or higher strain rates. The effect of
austenite grain size was significant on the grain
size of the dynamically transformed ferrite, and
the finer austenite grain size resulted in the finer
ferrite. Finally, the mean grain size of ferrite
could be smaller than 1 lm, which was consid-
ered to be achieved through dynamic transfor-
mation and subsequent dynamic recrystallization
of ferrite.
ACKNOWLEDGEMENTS
The authors would like to gratefully thank Prof.
Hiroyuki Yasuda of Osaka University for his con-
siderable support in thermomechanical experi-
ments. This study was financially supported by the
Grant-in-Aid for Scientific Research on Innovative
Area, ‘‘Bulk Nanostructured Metals’’ (Area
No.2201), the Grant-in-Aid for Scientific Research
(A) (No.24246114), and the Elements Strategy Ini-
tiative for Structural Materials (ESISM), all
through the Ministry of Education, Culture, Sports,
Science and Technology (MEXT), Japan (Contact
No. 22102002). N.P. was supported also by the
Japan Society for Promotion of Science (JSPS) as a
JSPS postdoctoral fellow. All the support is grate-
fully appreciated.
REFERENCES
1. K. Nishioka and K. Ichikawa, Sci. Technol. Adv. Mater. 13,
023001 (2012).
2. H. Yada, Y. Matsumura, and K. Nakajima, U.S. Patent
4466842 (August 1984).
3. Y. Matsumura and H. Yada, Trans. ISIJ 27, 492 (1987).
Fig. 6. Change in the grain size of dynamically transformed ferrite as
a function of the Zener–Hollomon parameter. Two austenite grain
sizes, 400 lm (square) and 15 lm (circle), were used for the
experiments. The activation energy used for calculating the Zener–
Hollomon parameter was 254 kJ mol
1
.
Fig. 7. A TEM micrograph of ferrite grains formed through a ther-
momechanical process at 650°C. The austenite grain size was
15 lm, the strain rate was 10
1
s
1
, and the strain was 0.96. The
double-headed arrow indicates the compression direction.
Park, Zhao, Shibata, and Tsuji
Author's personal copy
4. H. Yada, T. Matsumura, and T. Senuma, Steels Other Met.
ISIJ 88, 200 (1988).
5. P. Hodgson, M.R. Hickson, and R.K. Gibbs, Scr. Mater. 40,
1179 (1999).
6. M.R. Hickson, P.J. Hurley, R.K. Gibbs, G.L. Kelly, and P.D.
Hodgson, Metall. Mater. Trans. 33, 1019 (2002).
7. S.C. Hong, S.H. Lim, H.S. Hong, K.J.K.S. Lee, and D.H.
Shin, Mater. Sci. Eng. 355, 241 (2003).
8. S.C. Hong, S.H. Lim, K.J. Lee, and D.H. Shin, ISIJ Int. 43,
394 (2003).
9. J.K. Choi, D.H. Seo, J.S. Lee, K.K. Um, and W.Y. Choo, ISIJ
Int. 43, 746 (2003).
10. T.H. Ahn, K.K. Um, J.K. Choi, D.H. Kim, K.H. Oh, M. Kim,
and H.N. Han, Mater. Sci. Eng. 523, 173 (2009).
11. C. Zheng, N. Xiao, L. Hao, D. Li, and Y. Li, Acta Mater. 57,
2956 (2009).
12. M.H. Cai and H.D.Y.K. Lee, Mater. Trans. 52, 1722 (2011).
13. S.C. Hong and K.S. Lee, Mater. Sci. Eng. 323, 148 (2002).
14. W. Choo, Trends Met. Mater. Eng. 16, 93 (2003).
15. J.K. Park, K.H. Kim, J.H. Chung, and S.Y. Ok, Metall.
Mater. Trans. 39, 235 (2008).
16. Y.Q. Weng, X.J. Sun, and H. Dong, Symposium on Ultrafine
Grain Structures (Beijing: Metallurgical Industry Press,
2005), pp. 9–15.
17. Y. Weng, Ultra-Fine Grained Steels (New York: Springer, 2009).
18. N. Park, A. Shibata, D. Terada, and N. Tsuji, Acta Mater. 61,
163 (2013).
19. N. Park, S. Khamsuk, A. Shibata, and N. Tsuji, Scr. Mater.
68, 538 (2013).
20. N. Park, S. Khamsuk, A. Shibata, and N. Tsuji, Scr. Mater.
68, 611 (2013).
21. R.Z. Wang and T.C. Lei, Scr. Metall. Mater. 31, 1193 (1994).
22. Y. Adachi, P.G. Xu, and Y. Tomota, ISIJ Int. 48, 1056 (2008).
23. C. Zheng, D. Li, S. Lu, and Y. Li, Scr. Mater. 58, 838 (2008).
24. S.Y. Ok and J.K. Park, Scr. Mater. 52, 1111 (2005).
25. A.Z. Hanzaki, R. Pandi, P.D. Hodgson, and S. Yue, Metall.
Trans. 24, 2657 (1993).
26. T. Sakai and J. Jonas, Acta Metall. 32, 189 (1984).
27. A. Cingara and H.J. McQueen, J. Mater. Process. Technol.
36, 17 (1992).
28. S.W. Lee, D.H. Seo, and W.Y. Choo, Korean J. Met. Mater.
36, 1966 (1998).
29. M. Tong, D. Li, Y. Li, J. Ni, and Y. Zhang, Metall. Mater.
Trans. 35, 1565 (2004).
30. M. Tong, J. Ni, Y. Zhang, D. Li, and Y. Li, Scr. Mater. 50,
909 (2004).
31. C. Ghosh, V.V. Basabe, J.J. Jonas, Y.M. Kim, I.H. Jung, and
S. Yue, Acta Mater. 61, 2348 (2013).
32. Nokeun Park (Ph.D. Dissertation, Kyoto University, 2013).
33. J.J. Jonas and C. Ghosh, Acta Mater. 61, 6125 (2013).
34. H. Dong and X. Sun, Curr. Opin. Solid State Mater. Sci. 9,
269 (2005).
35. A. Ohmori, S. Torizuka, K. Nagai, N. Koseki, and Y. Kogo,
Metall. Mater. Trans. 45, 2224 (2004).
36. P.J. Hurley, P.D. Hodgson, and B.C. Muddle, Scr. Mater. 40,
433 (1999).
37. L.Zhao,N.Park,A.Shibata,andN.Tsuji,TMS 2014
Annual Meeting Supplemental Proceedings, 2014, p. 919.
Dynamic Ferrite Transformation Behaviors in 6Ni-0.1C Steel
Author's personal copy
... Besides the double differentiation method, analysis of the variation in maximum stress of the stress-strain curves can prove the occurrence of DT as well. Park et al. [99] conducted compression tests using a 6Ni-0.1C steel and compared the deformation behavior of austenite with that of ferrite under various deformation conditions. ...
... Evidence for the occurrence of DT by analysis of the maximum stress of stress-strain curves. [99] a), b) True stress-true strain curves of the 6Ni-0.1C (wt%) steel specimens a) with full austenite phase and b) with mostly ferrite phase deformed at various temperatures at a strain rate of 10 À1 s À1 . ...
... When austenite and ferrite are present simultaneously, the flow stress can be derived by the method from the ref. [108] Park et al. [99] calculated the phase diagram of a Fe-6Ni-C system using Thermo-Calc software, as shown in Figure 8. The change in Ae 3 lines with different additional energies on austenite phase is presented in Figure 8a. ...
Article
Full-text available
Ultrafine grained (UFG) steels with grain sizes around 1 micron exhibit an excellent strength-ductility combination and have been extensively studied worldwide. Among the different grain refinement strategies, thermomechanical controlled processing (TMCP) employing dynamic transformation (DT), that is, ferrite transformation during deformation of austenite, is considered as the simplest and commercially exploitable approach to produce ultrafine ferrite (UFF) with grain size of a couple of microns or below. The present paper reviews the research history of DT and highlights the major aspects of continuous interest including the methods and evidences for identifying DT, thermodynamics and kinetics of DT, mechanism for UFF formation and the effects of some key thermomechanical parameters on DT (and UFF formation), together with an outlook for the future research, and new TMCP design for industrial application. This paper also discusses some areas remaining under debate such as the diffusional or displacive mechanism, thermodynamic modeling, and the mechanism for UFF formation, etc.
... In this scenario, the kinetics of the bainitic transformation is accelerated by increasing the initial number density of potential nucleation sites for bainite [6,14], besides increasing the local carbon diffusivity in austenite [39]. Increased nucleation rate is also suggested as the main mechanism which boosts dynamic transformations (e.g., [37,[39][40][41][42]). According to the literature, when the stress exceeds the yield strength of austenite, the defects introduced by deformation enhance the nucleation rate, and hence the overall rate of bainite reaction increases [43]. ...
Article
Full-text available
In this work, the microstructural evolution during the dynamic transformation of austenite to bainite was directly observed by in-situ high energy synchrotron X-ray diffraction measurements during warm uniaxial compression performed at the P07 beamline of PETRA III, DESY (Deutsches Elektronen-Synchrotron). Plastic deformation triggers the phase transformation, which is continuously stimulated by the introduction of dynamic dislocations into the austenite. This scenario accelerates the kinetics of bainite formation in comparison with conventional isothermal treatment. No mechanical stabilization of austenite was observed during dynamic transformation. Evidence of carbon partitioning between phases during plastic deformation was obtained. Further post-process investigations suggest that the bainitic microstructure developed during compression is oriented perpendicular to the loading direction. The findings open up new possibilities to design carbide-free bainitic microstructures directly via thermomechanical processing.
... The fine-grained austenitic microstructure also leads to a reduction in the critical deformation for a ferrite formation by dynamic strain-induced transformation (DSIT) and to a relatively homogeneous size distribution of ferritic grains. Conversely, in the case of a coarse-grained prior structure, the newly formed ferritic grains are primarily located along the boundaries of the austenitic grains [25][26][27][28]. However, this conclusion is only valid in the case of allotriomorphic ferrite. ...
Article
Full-text available
The combined effect of deformation temperature and strain value on the continuous cooling transformation (CCT) diagram of low-alloy steel with 0.23% C, 1.17% Mn, 0.79% Ni, 0.44% Cr, and 0.22% Mo was studied. The deformation temperature (identical to the austenitization temperature) was in the range suitable for the wire rolling mill. The applied compressive deformation corresponded to the true strain values in an unusually wide range. Based on the dilatometric tests and metallographic analyses, a total of five different CCT diagrams were constructed. Pre-deformation corresponding to the true strain of 0.35 or even 1.0 had no clear effect on the austenite decomposition kinetics at the austenitization temperature of 880 °C. During the long-lasting cooling, recrystallization and probably coarsening of the new austenitic grains occurred, which almost eliminated the influence of pre-deformation on the temperatures of the diffusion-controlled phase transformations. Decreasing the deformation temperature to 830 °C led to the significant acceleration of the austenite → ferrite and austenite → pearlite transformations due to the applied strain of 1.0 only in the region of the cooling rate between 3 and 35 °C·s−1. The kinetics of the bainitic or martensitic transformation remained practically unaffected by the pre-deformation. The acceleration of the diffusion-controlled phase transformations resulted from the formation of an austenitic microstructure with a mean grain size of about 4 µm. As the analysis of the stress–strain curves showed, the grain refinement was carried out by dynamic and metadynamic recrystallization. At low cooling rates, the effect of plastic deformation on the kinetics of phase transformations was indistinct.
... Several researchers have found that the grain size of ferrite decreases down to approximately 1 μm through thermomechanical processing at relatively low temperature [3][4][5][6][7]. In this process, phase transformation might occur during deformation of parent austenite phase because transformation kinetics is greatly accelerated by a large amount of lattice defects introduced into austenite [8]. In order to distinguish with normal ferrite transformation, i.e., static ferrite transformation (ST), which occurs during cooling or static holding, the ferrite transformation occurring during deformation of austenite has been termed as dynamic ferrite transformation [3,[9][10][11][12][13], strain-induced ferrite transformation [4,5,14,15], deformation-induced ferrite transformation [16,17], and so on. ...
Article
Full-text available
Nowadays, a new concept of process utilizing dynamic ferrite transformation, which can achieve ultrafine-grained structure with a mean grain size of approximately 1 μm, has been proposed. This paper reports transformation mode of dynamic ferrite transformation and formation mechanism of ultrafine-grained structure revealed by our novel technique of in-situ neutron diffraction analysis during thermomechanical processing. Dynamic ferrite transformation occurs in a diffusional manner, whose partitioning behavior changes from para- to ortho-equilibrium with the progress of transformation. Moreover, we propose that dynamic recrystallization of dynamically-transformed ferrite is the main mechanism for the formation of ultrafine-grained structure.
... static transformation, ST) [9,10] and (2) phase transformation during deformation (i.e. dynamic transformation, DT) [4,5,[11][12][13][14][15]. The deformation applied before or during transformation determines how and where the defects are introduced, and therefore could lead to different scenarios of transformation kinetics affected by nucleation density and growth rate related to atomic diffusion. ...
Article
Full-text available
Comparison on the kinetics of two different phase transformations, including phase transformation after deformation and phase transformation during deformation (i.e. dynamic transformation, DT), reveals a new discovery that the transformation kinetics can be significantly enhanced in DT even under low driving forces. DT enables continuous generation of defects (e.g. dislocations) near the phase boundary, which can act as short-circuiting diffusion paths for atoms. The diffusivity of atoms is enhanced and the activation energy for the atom jump across the phase boundary is lowered under stress during DT, resulting in more pronounced grain growth as well as accelerated transformation kinetics.
Article
Full-text available
Double-hit hot compression tests were carried on medium-carbon low-alloy steels using Gleeble 3800® thermomechanical simulator. The experiments were performed at strain rates of 0.25 and 0.5 s−1 and temperatures of 1150 and 1200 °C with interpass times of 5, 15, and 25 s. The onset of critical stresses for dynamic transformation (DT) for both first and second hit were detected using the double-differentiation method. It was found that the critical stress for DT increased with a decrease in temperature and an increase in strain rate. The presence of dynamically transformed ferrite was observed and quantified using electron-backscatter diffraction, kernal average misorientation, and grain boundary maps. Then, a thermodynamic analysis was carried out using JmatPro software. A method of determining the change in Gibbs energy during DT phenomenon is proposed for double hit deformation.
Article
The solute partitioning under dynamic strain-induced austenite reversion in a Fe-11Mn-1Al-0.07C (in wt.%) steel was studied by atom probe tomography. It was found that the concentrations of Mn and Al fluctuate in austenite at the nanoscale during warm deformation at 650 °C, leading to less stable chemical zones. It is proposed that stored energy in ferrite can change the thermodynamic equilibrium, leading to the formation of austenite with lower Mn and higher Al contents. This paper reveals evidence that stored strain can deviate the thermodynamic equilibrium, and hence could be a new path for engineering the nanostructures in austenite.
Article
Full-text available
There have been previous attempts to observe the occurrence of dynamic ferritic transformation at temperatures even above Ae3 in a low-carbon steel, and not only in steels, but recently also in titanium alloys. In this study, a new approach is proposed that involves treating true stress-true strain curves in uniaxial compression tests at various temperatures, and different strain rates in 0.1C-6Ni steel, which is a model alloy used to decelerate the kinetics of ferrite transformation from austenite. The initial flow stress up to peak stress was used to analyze the change in dynamic softening phenomena, such as dynamic recovery, dynamic recrystallization, and dynamic transformation. It is worth mentioning that for predicting the occurrence of dynamic transformation, flow stress before reaching peak stress is much more sensitive to the change in the dynamic softening rate due to dynamic transformation, compared to peak stress. It was found that the occurrence of dynamic ferritic transformation could be successfully obtained even at temperatures above Ae3 once the deformation condition was satisfied. This deformation condition is a function of both the strain rate and the deformation temperature, which can be described as the Zener - Hollomon parameter. In addition, the driving force of dynamic ferritic transformation might be much less than that of the dynamic recrystallization of austenite at a given deformation condition. By applying this technique, it is possible to predict the occurrence of dynamic transformation more sensitively compared with the previous analysis method using peak stress during deformation.
Article
Full-text available
Plate rolling simulations were carried out on an X70 Nb and a low C steel by means of torsion testing. A seven-pass rolling schedule was employed where the last pass was always applied above the respective Ae3 temperature of the steel. Interpass intervals of 10 and 30s were employed, which corresponded to cooling rates of 1.5 and 0.5 C/s. The mean flow stresses (MFS`s) applicable to each schedule increased less rapidly than expected from the decreases in temperature due to the dynamic transformation (DT) that took place during straining. The amounts of ferrite that retransformed into austenite during holding were determined by optical metallography. These increased with length of the interpass intervals and were reduced in the X70 steel due to the presence of Nb. The holding times after rolling required to increase the amount of austenite available for microstructure control on subsequent cooling were also determined for the two steels.
Article
Full-text available
The effect of strain on the microstructural evolution of ferrite grains during dynamic transformation was investigated using a 10Ni-0.1C steel uniaxially compressed at a strain rate of 10-2 s-1 and temperature of 520 °C. The deformation of the ferrite formed at relatively early stages of dynamic transformation led to the formation of subgrains inside the ferrite grains. The misorientation between subgrains changed from low angles to high angles with increasing strain. The formation of equiaxed grains surrounded by high angle boundaries was confirmed, of which volume fraction increased with increasing compression strain. The results indicated that the grain refinement in the process was not only due to dynamic transformation but also due to the deformation and dynamic recovery/recrystallization of ferrite grains transformed at relatively early stages of dynamic transformation.
Article
Full-text available
The water-cooled thermomechanical control process (TMCP) is a technology for improving the strength and toughness of water-cooled steel plates, while allowing control of the microstructure, phase transformation and rolling. This review describes metallurgical aspects of the microalloying of steel, such as niobium addition, and discusses advantages of TMCP, for example, in terms of weldability, which is reduced upon alloying. Other covered topics include the development of equipment, distortions in steel plates, peripheral technologies such as steel making and casting, and theoretical modeling, as well as the history of property control in steel plate production and some early TMCP technologies. We provide some of the latest examples of applications of TMCP steel in various industries such as shipbuilding, offshore structures, building construction, bridges, pipelines, penstocks and cryogenic tanks. This review also introduces high heat-affected-zone toughness technologies, wherein the microstructure of steel is improved by the addition of fine particles of magnesium-containing sulfides and magnesium- or calcium-containing oxides. We demonstrate that thanks to ongoing developments TMCP has the potential to meet the ever-increasing demands of steel plates.
Article
Large deformation for a common C-Mn steel at the temperatures around Ar3 results in an ultrafine-grained ferrite structure. This is due to dynamic transformation and recrystallization of ferrite during deformation. To apply the new finding to an actual plant rolling, multi-pass deformations which substitute for a single pass heavy deformation have been studied.It was realized that when the interpass time during successive deformations is shorter than 2s, the effect of accumulative strain can make ferrite grains fine.A trial of strip rolling in plant provided a mean ferrite grain size of ASTM No. 13.5.
Article
When austenite is deformed above the equilibrium transformation temperature Ae(3), it is dynamically transformed into Widmanstatten ferrite by a displacive mechanism. On removal of the load it is slowly retransformed into austenite by diffusional processes. The forward transformation has recently been explained in terms of a thermodynamic model in which the lower free energy of austenite is raised above that of normally unstable ferrite as a result of the additional stored energy associated with the dislocations introduced by straining. This model is here shown to be unable to account for the initiation of transformation at critical strains of about 0.1, at which only low densities of dislocations are present. Of particular importance is the observation that dynamic transformation can be initiated at temperatures 100 C and more above the Ae3 and that the critical strain actually decreases with increasing temperature and increasing chemical free energy barrier. This discrepancy is removed by allowing for mechanical (stress-based) activation of the transformation. The latter provides the energy required to accommodate the shear of the parent austenite into Widmanstatten plates, as well as the volume change or dilatation accompanying ferrite formation. The work of dilatation and the shear accommodation work, omitted from the previous analysis, are introduced here as barriers to the transformation that are overcome by the applied stress. This modified approach is able to account for the very rapid forward (mechanically activated) transformation compared with the much slower reverse transformation that takes place in the absence of stress.
Article
In previous studies related to the austenite to ferrite transformation of controlled rolled steels it was shown that there is a limiting ferrite grain size of approximately 5 {micro}m regardless of the level of retained strain introduced into the austenite. This is at least partly due to ferrite coarsening during transformation. If, however, a 1 {micro}m grain size -- here called ultrafine ferrite (UFF) -- could be produced and retained then this would increase the yield strength by almost 350 MPa compared to a 5 {micro}m ferrite. Most recently, the current authors have reported the development of a new thermomechanical process which produces UFF in hot rolled steel strip. This grain refinement appears to be the result of a strain induced transformation reaction activated over a significant volume of the austenite. This requires high undercooling and high effective rolling strains. This paper describes how this combination of austenite conditioning, deformation conditions and temperature control can lead to the formation of significant volumes of UFF in low carbon steel.
Article
The effect of austenite grain size on kinetics of dynamic ferrite transformation above Ae(3) in a 6Ni-0.1C steel was studied. As the austenite grain size decreased, the onset of dynamic transformation was accelerated. The increase in the fraction of dynamically transformed ferrite was in good agreement with the change in flow stress, i.e. dynamic softening. The kinetics of dynamic transformation could be evaluated by an Avrami-type formula. (c) 2012 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Article
Dynamic ferrite transformation behavior was investigated over a wide temperature range using a 6Ni–0.1C steel. Softening in flow stress due to dynamic transformation was observed at temperatures above Ae3, the ortho-equilibrium austenite-to-ferrite transformation temperature. Microstructural observation revealed that ferrite grains formed at temperatures above Ae3 showed deformation microstructures, and those grains were reversely transformed to austenite during subsequent holding at the same temperature. Therefore, we concluded that dynamic ferrite transformation could certainly occur even above Ae3.
Article
Recent observations regarding the dynamic transformation of deformed austenite at temperatures above the Ae3 are reviewed. Experimental results obtained on four different steels over the temperature range from 743 to 917 °C and at strains up to ε = 5 are described. It is shown that there is a critical strain for the formation of superequilibrium ferrite and that the volume fraction of transformed ferrite increases with the strain. The structures observed are Widmanstätten in form and appear to have nucleated displacively. The effect of deformation on the Gibbs energy of austenite is estimated by assuming that the austenite continues to work-harden after initiation of the transformation and that its flow stress and dislocation density can be derived from the experimental flow curve by making suitable assumptions about two-phase flow. By further taking into account the inhomogeneity of the dislocation density, Gibbs energy contributions (driving forces) are derived that are sufficient to promote transformation as much as 100 °C above the Ae3. The C diffusion times required for the dynamic formation of the cementite particles observed are estimated. These range from ∼25 to 100 μs and are therefore consistent with the times available during rolling. The Gibbs energy calculations suggest that growth of the Widmanstätten ferrite is followed by C diffusion at the lower carbon contents, while it is accompanied by C diffusion at the higher carbon levels.
Article
In order to clarify the occurrence of dynamic ferrite transformation in a 6Ni–0.1C steel, the stress–strain behavior in uniaxial compression was analyzed for a wide range of temperatures and strain rates. Significant softening of flow stress for austenite was observed at lower temperatures at a constant strain rate, which seemed to correspond with the occurrence of dynamic transformation to ferrite. Analysis of the maximum stress in the stress–strain curves indicated that dynamic ferrite transformation occurred above a certain value of the Zener–Hollomon parameter (Z). The critical deformation condition (ZC) for the occurrence of dynamic transformation was determined. Increasing the amount of softening resulted in an increase in the fraction of ferrite, and the maximum flow stress came close to the flow stress of ferrite. Microstructural observations revealed that the specimens exhibiting softening consisted of ferrite grains with typical characteristics of deformation microstructure, such as a change in crystal orientation within the ferrite grain, inhomogeneity in ferrite morphology and dislocation substructures inside the grains. All these characteristics confirmed the occurrence of ferrite transformation during deformation, i.e. dynamic ferrite transformation.
Book
Steel are the basic materials in human civilization. Sufficient quantity of high quality steel materials is necessary for realization of industrialization of countries in the world and for providing conditions for the modern life of mankind.