Você está na página 1de 14

Austenite Formation in Plain Low-Carbon Steels

HAMID AZIZI-ALIZAMINI, MATTHIAS MILITZER, and WARREN J. POOLE


In this study, austenite formation from hot-rolled (HR) and cold-rolled (CR) ferrite-pearlite
structures in a plain low-carbon steel was investigated using dilation data and microstructural
analysis. Dierent stages of microstructural evolution during heating of the HR and CR
samples were investigated. These stages include austenite formation from pearlite colonies,
ferrite-to-austenite transformation, and nal carbide dissolution. In the CR samples, recrys-
tallization of deformed ferrite and spheroidization of pearlite lamellae before transformation
were evident at low heating rates. An increase in heating rate resulted in a delay in spheroidi-
zation of cementite lamellae and in recrystallization of ferrite grains in the CR steel. Further-
more, a morphological transition is observed during austenitization in both HR and CR
samples with increasing heating rate. In HR samples, a change from blocky austenite grains to a
ne network of these grains along ferrite grain boundaries occurs. In the CR samples, austenite
formation changes from a random spatial distribution to a banded morphology.
DOI: 10.1007/s11661-010-0551-5
The Minerals, Metals & Materials Society and ASM International 2010
I. INTRODUCTION
AUSTENITE formation occurs during many indus-
trial heat treatments of steels. However, the importance
of this phase transformation has been undervalued
because austenite usually is not found in the micro-
structure of nal steel products. Furthermore, it is
challenging to characterize austenite microstructures
that are present at high temperatures. These limitations
resulted in few studies of austenite formation compared
with austenite decomposition. Nevertheless, a signicant
body of work on austenite formation is available in the
literature.
[15]
However, with the development of ad-
vanced high-strength steels such as dual phase (DP),
transformation-induced plasticity (TRIP), and complex
phase steels, there has been renewed interest in studying
austenite formation. For example, intercritical anneal-
ing is an essential processing step for these steels when
manufactured as cold-rolled and coated sheets, primar-
ily for automotive applications. In addition, austenite
formation is of major interest for microstructure evolu-
tion in the heat-aected zone of welds. Microstructures
from which austenite formation has been investigated
include hot-rolled (HR) and cold-rolled (CR) ferrite-
pearlite,
[6,7]
ferrite-spheroidized carbide,
[8]
and ferrite-
martensite
[9]
structures. Investigating the austenite for-
mation from ferrite-pearlite structures, Speich et al.
[1]
observed three major transformation stages, i.e., (1)
rapid pearlite-to-austenite transformation, (2) austenite
formation from proeutectoid ferrite, and (3) nal
equilibrium via partitioning of Mn in austenite. Yang
et al.
[7]
investigated the eect of initial cold reduction on
the kinetics of austenite formation and its morphology.
They observed that recrystallization of deformed ferrite
and spheroidization of pearlite lamellae can take place
prior to austenite formation in the CR sample. Using
dilatometry data, de Cock et al.
[10]
showed that ferrite
recrystallization prior to phase transformation results in
dimension changes prior to austenite formation in low
and ultra-low carbon steels. The sequence of these
microstructural changes is dependent strongly on the
employed heating rate in both HR and CR struc-
tures.
[6,11,12]
San Mart n et al.
[6]
showed that there can
be an overlap between the rst two stages of austenite
formation at suciently low heating rates (e.g., 0.05 K/s)
in a HR low-carbon Nb microalloyed steel. Savran
et al.
[11]
also reported that lamellar ferrite and cementite
phases in pearlite colonies either can transform simul-
taneously or consecutively depending on heating rate. In
the CR structures, however, rapid heating results in an
overlap between ferrite recrystallization and austenite
formation. This interaction aects the morphology of
austenite directly and, consequently, mechanical prop-
erties of intercritically annealed multiphase steels.
Huang et al.
[12]
investigated systematically the eect of
heating rate on the microstructure of a Mo-alloyed DP
steel. They showed that an overlap between ferrite
recrystallization and austenite formation with increasing
the heating rate from 1 K/s to 100 K/s resulted in a
morphological transition from a randomly distributed
to a banded structure of martensite. As a result, a
signicant change in mechanical properties was
recorded.
[13]
The same morphological transition, from
random to brous distribution, was observed by
Grange.
[14]
The extent of this overlap depends on the
chemical composition of the steel and amount of cold
reduction. Petrov et al.
[15]
and Huang et al.
[12]
reported
that an ultrafast heating rate, i.e., in excess of 1000 K/s,
is needed to view the overlap in the CMnSi TRIP steels
used in their studies. Kestens et al.
[16]
showed that this
overlap is not viable even at 3000 K/s in an interstitial
HAMID AZIZI-ALIZAMINI, PhD Student, MATTHIAS
MILITZER and WARREN J. POOLE, Professors, are with the
Centre for Metallurgical Process Engineering, The University of
British Columbia, Vancouver, British Columbia V6T 1Z4, Canada.
Contact e-mail: hazizi@interchange.ubc.ca
Manuscript submitted March 29, 2010.
Article published online December 3, 2010
1544VOLUME 42A, JUNE 2011 METALLURGICAL AND MATERIALS TRANSACTIONS A
free (IF) steel. Overall, there are a limited number of
studies devoted to the eect of heating rate on the
interaction between ferrite recrystallization and austen-
ite formation and its impact on the microstructural
evolution in CR steel products
[12,1517]
The aim of the current work is to investigate
systematically the eect of initial structure, both HR
and CR, and heating rate on the dilation response, and
microstructural changes during austenite formation in a
low-carbon steel. In particular, this study is designed to
quantify experimentally individual stages of austenite
formation and their potential interaction with ferrite
recrystallization and spheroidization of cementite lamel-
lae. Advancing knowledge in this area is critical to
evaluate intercritical annealing strategies for advanced
high-strength steels.
II. EXPERIMENTAL
A plain, low-carbon steel received as industrially hot-
rolled material was used for this study. The detailed
chemical composition is presented in Table I. The HR
steel was then 80 pct cold rolled, i.e., from 9.8 mm to
1.8 mm using a laboratory rolling mill (roll diameter:
130 mm). For austenite formation studies, test coupons
of 10 9 60 9 1.8 mm were cut from the HR and CR
sheets with longitudinal direction of the test coupon being
aligned with the rolling direction. A Gleeble 3500
(Dynamic Systems Inc., Poestenkill, NY) thermome-
chanical simulator was employed for all heat treatments.
The temperature was controlled using a type K thermo-
couple spot welded on the center of the sample. Dilatom-
etry tests were conducted under high vacuum, 0.26 Pa
(2.0 9 10
3
Torr). Continuous heating tests were per-
formed with heating rates ranging from1 K/s to 900 K/s.
Adilatometer was attached to the center of the samples to
measure the change in width during heating. The volume
fraction of austenite was determined via analyzing the
dilatometric data using the lever rule. Details of the
procedure for this measurement can be found else-
where.
[18]
To analyze the microstructure during heating,
additional samples were then subjected to interrupted
heating tests and water quenching. The cooling rate was
approximately 1000 K/s to ensure complete austenite-to-
martensite transformation after quenching. For these
tests, the test chamber was back lled with inert Ar gas
after a high vacuum had been achieved.
A microstructural analysis was carried out along the
transverse direction. The microstructures were charac-
terized using optical and electron microscopy. A Hitachi
S2300 (Hitachi Science Systems Ltd., Tokyo, Japan)
scanning electron microscope (SEM) with a secondary
electron detector and an energy-dispersive X-ray system
for chemical analysis was used. AHitachi H-800 (Hitachi
Science Systems Ltd., Tokyo, Japan) transmission
electron microscope (TEM) operated at 200 kV was
employed for TEM observation. Scanning Auger micro-
scopy was used for the characterization of carbide
particles using a Microlab 350 system (Thermo Electron
Corp.) equipped with eld emission source (10 keV and
3.5 nA) and hemispherical energy analyzer in a vacuum
of 2910
7
Pa. A secondary electron detector attached to
the equipment was used to characterize selected carbide
particles. LePera etching
[19]
was employed to reveal
martensite and to measure the volume fraction of
austenite (martensite at room temperature) from opti-
cal micrographs. To reveal prior austenite grain bound-
aries, the following procedure was followed: First,
as-quenched samples were tempered in a tube furnace
at 823 K (550 C) for 15 hours in an Ar atmosphere
followed by water quenching. Then, an etching solution
composed of aqueous picric acid with sodium dodecyl-
benzene and a few droplets of Triton X-100 (Sigma-
Aldrich, St. Louis, MO) as a surface active agent at a
temperature range between 333 K and 353 K (60 C and
80 C) were used to reveal austenite grain boundaries.
Grain size measurements were based on the equivalent
area diameter approach and at least 500 grains were
analyzed using SEM. The quantitative measurements
were conducted using Clemex image analysis software
(Clemex Technologies Inc., Longueuil, PQ, Canada).
For SEM analyses, samples were electropolished in
95 pct acetic acid and 5 pct perchloric acid solution, and
then etched with 3 pct Nital. To reveal cementite
particles inside martensite islands, two-step etching
was employed. After light etching with 2 pct Nital, deep
etching using 4 pct Picral for 10 to 15 seconds was
performed.
[20]
These samples were also used for Auger
measurements. For TEM observations, thin foils were
prepared by twin-jet polishing technique using a mixture
of 95 pct acetic acid and 5 pct perchloric acid at an
applied potential of 40 V at 293 K (20 C).
III. RESULTS
A. Initial Structures
Figure 1(a) shows the initial HR ferrite-pearlite struc-
ture consisting of approximately 20 vol pct pearlite and
80 vol pct ferrite. Ferrite grain size and pearlite lamellar
spacing are approximately 7 lm and 240 nm, respec-
tively. Individual cementite particles were also observed
at ferrite grain boundaries (indicated by arrows in the
inset in Figure 1(a)). The HR microstructure is banded,
which can be related to the segregation of Mn. Energy-
dispersive X-ray analysis conrmed that concentration
of Mn in ferrite region was close to the nominal value in
the steel, however, it indicated segregation inside pearl-
ite colonies with an average Mn content of approxi-
mately 1.0 wt pct and an average distance of 12 lm
between pearlite bands. Figure 1(b) shows the CR
Table I. Chemical Composition of the Steel Used in this Study in Weight Percent
Element Fe C Mn P S Si Al N
Wt pct Balanced 0.17 0.74 0.009 0.008 0.012 0.04 0.0047
METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 42A, JUNE 20111545
structure, where the alignment of ferrite grains and
pearlite colonies into the rolling direction is evident in
the microstructure. The higher magnication inset in the
gure shows fragmentation and bending (arrows) of
carbide plates. After 80 pct cold reduction, there was an
inhomogeneous distribution of deformation in ferrite
and pearlite. A measurement of the average thickness of
pearlite colonies before and after deformation showed
that the amount of cold reduction is only 70 pct in
pearlite. The details of the procedures for strain mea-
surements can be found elsewhere.
[13]
B. Dilation Response and Microstructural
Characteristics in the HR and CR Materials
Figure 2 shows dilation curves for the HR and CR
samples (black and gray lines, respectively) heated at
1 K/s into the single austenite phase region. Using the
rst derivative of the dilation curves, several evolution
steps can be distinguished. In region (1), lattice expan-
sion in ferrite-pearlite structures in both HR and CR
samples takes place at heating. The recovery of
deformed structure can also occur in the CR sample.
In region (2), deviation from linear thermal expansion
can be observed in the CR sample, whereas the dilation
response in the HR steel remains unaected. Both of
these two stages are prior to austenite formation.
A sharp drop in the dilation curve is evident in both
HR and CR samples in stage (3), which is related to
pearlite-to-austenite transformation. Austenite forma-
tion continues in region (4) for both steels. The dilation
in this region can be related to ferrite-to-austenite
transformation. Finally, in region (5), linear lattice
expansion of austenite after completion of transforma-
tion is observed. In detail, these stages can be rational-
ized in the following sections.
1. Thermal expansion of ferrite-pearlite structure
In region (1), below 773 K (500 C), lattice expan-
sion of ferrite-pearlite structure takes place in both HR
and CR structures after heating. In the CR sample,
recovery of the deformed structure can proceed via
rearrangement and annihilation of dislocations, but this
cannot be observed with dilatometry. The microstruc-
tural features of the CR steel remain unaected (com-
pare Figures 3(a) and 1(b)). The measured linear
thermal expansion coecient of ferrite-pearlite structure
is 16.0 9 10
6
K
1
that is in good agreement with the
data available in the literature.
[21]
2. Recrystallization of ferrite in the CR sample
In region (2), the linear thermal expansion continues in
the HR sample, but it deviates from linearity in the CR
steel. This deviation starts at 773 K (500 C) and
reaches up to 0.1 pct at 923 K (650 C). Microstructural
observations, which are shown in Figure 3(b), indicate
that there are two major changes during this process: (1)
recrystallization of deformed ferrite grains and (2)
Fig. 1(a) Initial HR structure and (b) after 80 pct cold reduction.
The insets show higher magnication images. ND, normal direction;
RD, rolling direction. (F: ferrite, P: pearlite, C: cementite).
Fig. 2Dilation curves and their rst derivatives for the HR and
CR steels during heating at 1 K/s.
1546VOLUME 42A, JUNE 2011 METALLURGICAL AND MATERIALS TRANSACTIONS A
spheroidization of cementite particles. The deviation
from a linear coecient of thermal expansion could arise
from several mechanisms, i.e., (1) volume change caused
by the loss of dislocation density during recovery and
recrystallization, (2) change in crystallographic texture,
(3) dissolution/spheroidization of carbides, and (4) relax-
ation of residual stresses from cold rolling. Subsequent
examinations, which are presented in detail in the
Appendix, revealed that ferrite recrystallization is the
main mechanism responsible for the observed deviation.
3. Pearlite-to-austenite transformation
The rst stage of austenite formation in the HR and
CR steel consists of pearlite-to-austenite transformation
(stage (3)). The sharp drop in the rst derivative of the
dilation curve in Figure 2 indicates a relatively fast
phase transformation rate. Closer observation reveals
that the transformation started earlier in the CR
samples. The lower level of the rst derivative curve
for the CR sample in this stage suggests a slower
transformation rate as compared with the HR sample.
Figure 4(a) shows martensite (formerly austenite at high
temperature) sweeping lamellar pearlite colonies in the
HR material. There is almost no sign of spheroidization
of cementite lamellae inside pearlite colonies prior to
austenitization. It has been reported that austenite
nucleates mainly at ferrite-pearlite interfaces as well as
the interface between pearlite colonies.
[22]
In addition,
simultaneous nucleation of austenite at ferrite grain
boundaries, especially at carbide particles, has also been
reported at low heating rates.
[6]
However, the situation is dierent in the CR steel.
Figure 4(b) represents the early stages of austenite
formation in the CR steel at 1003 K (730 C). It can
be observed that cementite lamellae were mostly sphero-
idized (90 to 95 pct) even though some small fraction
of the lamellar structures are still preserved in the
microstructure (shown by arrow P in Figure 4(b)). A
closer examination of these latter regions reveals an
interesting observation. The bright eld TEM image
shown in Figure 4(c) provides an example for a ferrite
recrystallization front moving into the deformed lamel-
lae leaving behind rows of spheroidized carbides. This
provides evidence that this is one mechanism to produce
spheroidized carbides, which is not available in the
undeformed case. These carbides are visible clearly in
Figure 4(b) and (d) (indicated by arrows C). At the same
time, deformed ferrite layers inside pearlite colonies
recrystallize and form ne equiaxed grains that are
pinned by carbide particles (Figure 4(d)). Recrystalliza-
tion is complete prior to austenite formation except for
the small regions in the pearlite colonies (e.g., P in
Figure 4(b)). Nucleation and growth of austenite occur
from the ferrite-spheroidized cementite aggregates.
Growth of austenite continues until complete consump-
tion of the aggregate. Figure 5(a) shows the presence of
particles inside martensite (austenite at high tempera-
ture) in the CR sample at 1013 K (740 C). The presence
of the particles inside martensite after completion of
pearlite-to-austenite transformation was also observed
for the HR steel. To conrm that these particles are
carbides, Auger electron microscopy was used. Prior to
taking Auger spectra, surface contaminations were
removed by sputtering with Ar ions. Figure 5(b) pro-
vides an example of Auger electron spectra where the
carbon content of the particle (black line) can be
compared with that in the martensite matrix (gray line).
A sharp increase in carbon content is observed in the
particle compared to the matrix. A subsequent analysis
on the particles using dierentiation spectrum compared
with the reference data
[23]
indicates that the carbon is
present in the form of a carbide that is most probably
cementite.
Figure 6(a) shows the microstructure of the CR
structure quenched at 1013 K (740 C). It can be
observed that subsequent growth of austenite grains
takes place at ferrite grain boundaries. A closer exam-
ination using the inset reveals that undissolved carbide
particles are mostly spheroidized inside the ferrite
matrix. The volume fraction of austenite at this tem-
perature is 0.22, which is almost equivalent to the initial
pearlite content in the steel. The temperature at which
the rst contraction, which results from pearlite-to-
austenite transformation, nishes was dened as Ac
h
, as
indicated in the derivative curve in Figure 2. It was
shown by San Mart n et al.
[6]
that the Ac
h
temperature
was a good metric to separate the pearlite-to-austenite
and ferrite-to-austenite transformations at low heat-
ing rates, e.g., 0.05 K/s, but at higher heating rates,
Fig. 3Microstructural evolution of the CR structure quenched at
(a) 783 K (510 C) (point A in Fig. 2(b)) and (b) 943 K (670 C)
(point B in Fig. 2(b)) during heating at 1 K/s.
METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 42A, JUNE 20111547
austenite formation at ferrite grain boundaries can also
occur below the Ac
h
temperature.
[6]
4. Ferrite-to-austenite transformation
Another increase in the austenite volume fraction
occurs by austenite growth into proeutectoid ferrite
grains, stage (4) in Figure 2. Dissolution of carbide
particles is completed during this stage, and no carbide
particles were evident at 1033 K (760 C) in ferrite or
inside austenite.
Figure 6(b) shows the ferrite-martensite DP structure
resulting when quenching from 1053 K (780 C) in the
CR sample. Here, martensite islands are distributed
randomly. It can be observed that a necklace shape
structure of these islands covers most ferrite grain
boundaries. The volume fraction of austenite in both
HR and CR structures calculated via lever rule using
dilatometry curves together with metallographic mea-
surements is presented in Figure 7. In addition to the
experimental results, the paraequilibrium (PE) austenite
fraction calculated from Thermo-Calc (Thermo-Calc
Software, Stockholm, Sweden) using the FE-2000 data-
base is shown for comparison (Note: paraequilibrium
indicates a constraint equilibrium without partitioning
of substitutional alloying elements). It can be observed
that the volume fraction obtained from dilatation data is
in good agreement with the metallographic observations
except at about 20 vol pct of austenite. At this point, the
dilation data underestimates the actual transformed
fraction. This point also coincides with complete pearl-
ite-to-austenite transformation. This discrepancy was
also observed by Oh et al.
[24]
in the early stage of
austenite formation experiments. They attributed this to
the redistribution of carbon atoms in austenite grains.
The carbon content in the austenite phase in the steel
used in this study can reach up to 0.75 wt pct at the
beginning of austenite formation based on paraequilib-
rium calculation. The lattice parameter and, thus, the
specic volume of austenite is a function of carbon
content.
[21]
This eect was not considered in the simpli-
ed lever-rule calculations based on the linearized
thermal expansion of austenite with 0.17 wt pct carbon
content. However, a correction to take into account
the carbon content of austenite using the proposed
approaches
[24]
can be misleading in the current case
because of the remnant of carbide particles inside
austenite (Figure 5(a)). One would need to determine
the volume fraction of cementite particles inside aus-
tenite grains and thereby estimate the actual carbon
content of austenite.
5. Thermal expansion of austenite
Subsequent heating into the single austenite phase
region, stage (5), consists of linear thermal expansion of
Fig. 4Microstructure of HR steel heated at 1 K/s to 1008 K (735 C) (a) and of CR steel heated at 1 K/s to 1003 K (730 C) (b), (c) and
(d) (F: ferrite, M: martensite, P: pearlite, C: cementite).
1548VOLUME 42A, JUNE 2011 METALLURGICAL AND MATERIALS TRANSACTIONS A
austenite grains. Both HR and CR materials expand
linearly in single austenite phase region with a thermal
expansion coecient of 23.0 9 10
6
K
1
that is in
agreement with the reported data for low-carbon
austenite.
[21,25]
C. Effect of Heating Rate on Dilation Response
and Microstructure Evolution
In this section, the eect of heating rate on the
dilation curves and microstructural evolution is studied.
Dierent heating rates of 1 K/s, 10 K/s, 100 K/s,
300 K/s, and 900 K/s were employed for the investiga-
tion. For dilation experiments, all samples were heated
into the single-phase austenite region.
Figure 8 shows dilation curves and their rst deriv-
atives for both HR and CR materials at dierent heating
rates. Table II summarizes the critical temperatures in
both HR and CR samples at dierent heating rates. Ae
1
(equilibrium austenite formation start temperature) and
Ae
3
(equilibrium austenite formation nish temperature)
are 977 K and 1096 K (704 C and 823 C), respec-
tively. The eect of Mn segregation on the equilibrium
temperatures remains marginal as it decreases Ae
1
and
Ae
3
by 9 K and 7 K, respectively. It is evident that
raising the heating rate will increase the Ac
1
temperature
(start of austenite formation upon heating), the Ac
3
temperature (nish of austenite formation at heating),
and the Ac
h
temperature in the HR material (Fig-
ure 8(a)). For example, the Ac
1
and Ac
3
temperatures
increase from 1003 K to 1063 K (730 C to 790 C) and
from 1127 K to 1200 K (854 C to 927 C), respectively,
when the heating rate is raised from 1 K/s to 900 K/s.
However, the overall trend in the shape of dilation
curves and their derivatives remains similar. From the
derivative of the dilation curves, two distinct and
perhaps overlapping stages can be observed, pearlite-
to-austenite transformation with a relatively sharp slope
followed by ferrite-to austenite transformation for all
the heating rates. Dilation curves for the CR structures,
however, reveal two distinct and rather interesting
observations in comparison with their HR counterparts
(Figure 8(b)): First, it can be observed that increasing
the heating rate from 1 K/s to 900 K/s resulted in the
disappearance of the deviation from linear thermal
expansion before phase transformation. This eect is
more pronounced in the derivative curves in which the
cusp for 1 K/s at approximately 923 K (650 C) is
gradually shifting up toward the line representing the
linear thermal expansion coecient of the ferrite-pearl-
ite structure (16.0 9 10
6
K
1
). Second, as the heating
rate increases, a negligible change occurs in the start
Fig. 5(a) Microstructure of the CR steel heated at 1 K/s to reveal
cementite particles inside martensite; arrows show carbide particles
(F: ferrite, M: martensite) and (b) Auger spectra for carbide particles
and martensite.
Fig. 6Microstructure of the CR steel continuously heated with
1 K/s heating rate followed by water quenching at (a) 1013 K
(740 C) (the arrows inside the inset depict spheroidized carbide par-
ticles) and (b) 1053 K (780 C) (F: ferrite, M: martensite).
METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 42A, JUNE 20111549
temperature of austenite formation represented by the
sharp drop in the derivative of dilation curves. It can be
observed that all these sharp drops collapse essentially
onto the same line. A similar behavior is observed for
the Ac
h
temperature. However, the change in the Ac
3
temperature is similar to that observed for the HR
material, i.e., the Ac
3
temperature increases from about
1125 K (852 C) to 1183 K (910 C) when the heating
rate ramps up from 1 K/s to 900 K/s.
Figure 9 shows austenite fractions in both HR and
CR steels as a function of temperature at dierent
heating rates measured using dilatometry. Prediction by
paraequilibrium calculation is also included in the
graphs. It can be observed that increasing the heating
rate increases both the Ac
1
and Ac
3
temperatures, and
shifts the transformation to higher temperatures in the
HR steel below equilibrium as explained previously.
Figure 9(b) also represents the lever-rule results for
austenite fractions for dierent heating rates in the CR
steel. A comparison between the lever-rule results and
metallographic measurements for 300 K/s heating rate is
given in Table III. It is evident that the lever-rule
analysis of dilation data for the early stages of austenite
formation is unsatisfactory, but there is a good agree-
ment between the lever-rule calculation and metallo-
graphic measurements for austenite fractions above 0.2
at 300 K/s. Similar observations were made for 1 K/s
heating rate; see Section IIIB. Furthermore, the dila-
tion measurements suggest that up to 20 vol pct aus-
tenite fraction, the transformation is independent of the
heating rate. However, beyond this point, increasing the
heating rates shifts the transformation gradually to
higher temperatures. Table IV represents the tempera-
ture for the formation of 0.5 volume fraction of aus-
tenite (T
0.5
) for dierent heating rates for both HR and
CR samples. These data indicate a reduced temperature
dependency of austenite fraction as a function of heating
rate in the CR steel. For example, the shift in T
0.5
when
Fig. 7Austenite fraction in the HR and CR steels as a function
of temperature for continuous heating at 1 K/s and PE austenite
fraction.
Fig. 8Dilation curves and their rst derivatives for (a) HR and (b) CR steels during heating with dierent heating rates.
1550VOLUME 42A, JUNE 2011 METALLURGICAL AND MATERIALS TRANSACTIONS A
increasing the heating rate from 1 K/s to 900 K/s is
92 K in the HR steel, whereas this shift is 33 K for the
CR sample.
Figure 10 shows the microstructures for the HR steel
heated at 1 K/s and 300 K/s to 1023 K (750 C)
followed by water quenching. It can be observed that
increasing the heating rate has changed the morphology
and distribution of austenite grains from a large blocky
one to a network of ne austenite grains along ferrite
grain boundaries.
Figure 11 shows the microstructural evolution of the
CR structure heated at 300 K/s followed by immediate
water quenching at 973 K, 1013 K, and 1053 K (700 C,
740 C, and 780 C), respectively. Several major dier-
ences exist between the microstructures of rapid-heated
samples in comparison with their low-heating-rate
counterparts in the CR steel shown in Figure 6.
(a) Figure 11(a) shows the microstructure at 973 K
(700 C). This temperature is just before the start
of austenite formation, but some regions in the
ferrite structure are not yet recrystallized. These
regions are distributed mostly between closely
spaced pearlite colonies, depicted by arrows.
Approximately 12 pct of ferrite remained unrecrys-
tallized at this stage. At the same time, the lamellar
pattern of cementite particles in the pearlite colo-
nies is almost unaected, and the spheroidization
of cementite lamellae in the pearlite colonies is at
early stages in contrast to the 1 K/s heating rate
experiment.
(b) At 1013 K (740 C) (Figure 11(b)), it can be
observed that the elongated austenite grains are
formed that are not observed during intercritical
annealing at 1 K/s (Figure 6(a)). A closer observa-
tion of Figure 11(c) reveals that austenite forma-
tion starts at pearlite colonies and ferrite grain
boundaries. The latter nucleation sites are shown
Table II. Summary of the Critical Transformation Temperatures (Ac
1
, Ac
h,
and Ac
3
in K (C)) at Dierent Heating Rates
Heating Rate (K/s) 1 10 100 900
HR sample Ac
1
1003 (730) 1012 (739) 1033 (760) 1063 (790)
Ac
h
1036 (763) 1051 (778) 1065 (792) 1085 (812)
Ac
3
1127 (854) 1143 (870) 1188 (915) 1223 (950)
CR sample Ac
1
992 (719) 993 (720) 997 (724) 1007 (734)
Ac
h
1044 (771) 1050 (777) 1033 (760) 1043 (770)
Ac
3
1125 (852) 1127 (854) 1160 (887) 1183 (910)
Fig. 9Austenite fraction in (a) HR and (b) CR steels as a function of temperature for dierent heating rates and paraequilibrium (PE) austen-
ite fraction.
Table III. Comparison of Austenite Volume Fraction
Measured Using Metallographic Analysis and the Lever Rule
for the CR Steel Heated at 300 K/s
Temperature,
[K (C)]
993
(720)
1033
(760)
1053
(780)
1073
(800)
1103
(830)
Lever rule 0.02 0.19 0.30 0.42 0.67
Metallography 0.14 0.26 0.30 0.38 0.70
Table IV. Eect of Heating Rate on T
0.5
in the HR
and CR Samples
Heating
rate (K/s) 1 10 100 900
T
0.5
(HR),
K (C)
1067 (794) 1073 (800) 1113 (840) 1159 (886)
T
0.5
(CR),
K (C)
1067 (794) 1071 (798) 1083 (810) 1100 (827)
METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 42A, JUNE 20111551
by arrows in Figure 11(c). The austenite grains
formed at ferrite grain boundaries are substantially
smaller than those formed at pearlite colonies
elongated in rolling direction. Figure 11(d) shows
the microstructure at 1053 K (780 C) with
0.3 volume fraction martensite. It can again be
observed that banded martensite islands are elon-
gated in rolling direction. This is in contrast to the
random distribution observed for lower heating
rates (Figure 6(b)). Similar to the sample quenched
from 1013 K (740 C), large numbers of isolated
ne austenite grains are distributed at the ferrite
grain boundaries. Unlike in the HR samples and
the CR samples heated at 1 K/s, no network of
austenite grains formed along ferrite grain bound-
aries. Furthermore, measurements of austenite
grain sizes on samples heated at 1 K/s and 300 K/s,
respectively, to just above the Ac
3
temperature fol-
lowed by water quenching revealed a grain size
reduction from 11 lm to 6 lm as a result of
increasing the heating rate (Figure 12).
(c) Figure 13(a) shows the distribution of carbide par-
ticles inside martensite islands in the sample
quenched from 1013 K (740 C) after heating at
300 K/s. It can be observed that the process of
spheroidization of cementite lamellae is almost
complete. A few remaining elongated particles are
not yet spheroidized; these particles are indicated
by arrows in Figure 13. A closer examination
shows that the size of the carbide particles is ner
than in the samples heated at 1 K/s to 1013 K
(740 C). The average size for carbide particles
decreases from 190 nm to 130 nm when increas-
ing the heating rate from 1 K/s (Figure 5(a)) to
300 K/s (Figure 13).
D. Cold-Rolled and Recrystallized Structures
Starting with the CR structures, it was shown that the
initial structure before austenite formation can be
dierent depending on the heating rate (Figures 3(b)
and 11(a)), which makes it dicult to compare the
results systematically. Thus, in a next step, all CR
samples were heated to 943 K (670 C) with 1 K/s
heating rate (point B in Figure 2(b)), and then heated
above the Ac
3
temperature with dierent heating rates
of 1 K/s, 10 K/s, 100 K/s, and 900 K/s. In this scenario,
the initial structure will be a recrystallized (REX) ferrite
matrix with partially spheroidized pearlite colonies
(Figure 3(b)), which is referred to hereafter as CR+
REX structure. Per denition, for heating at 1 K/s the
CR+REX sample coincides with the CR one. Fig-
ure 14(a) shows the dilation curves and their rst
derivatives for the CR+REX samples. Unlike for the
CR structures (Figure 8(b)), the shift in the Ac
1
tem-
peratures via increasing the heating rate is now evident.
Figure 14(b) represents the austenite fraction vs tem-
perature at dierent heating rates for CR+REX
structures. With increasing the heating rate, the curves
shift toward higher temperature, resembling the trend
observed in the HR structures.
This trend is summarized in Figure 14(c) in which the
Ac
1
temperature represented by a sharp drop in the
derivative of the dilation curves is shown for the HR,
CR, and CR+REX samples. As discussed, the shift in
the Ac
1
temperature for the CR samples is just 15 K
when increasing the heating rate by almost three orders
of magnitude. In contrast, this shift is approximately
60 K for both HR and CR+REX samples, which
conrms the importance of the overlap between recrys-
tallization and austenite formation. The start tempera-
tures for austenite formation in the CR+REX samples
are approximately 10 K lower than those for the HR
samples at all employed heating rates.
IV. DISCUSSION
The observations made during high-heating-rate
experiments can be explained in terms of the kinetics
of the events such as recrystallization, spheroidization,
and austenite formation, and their interactions. In the
HR steel, austenite formation is the main process,
whereas in the CR samples, spheroidization of cementite
lamellae, recrystallization of deformed ferrite grains,
and austenite formation are the major events that take
place during heating. The heating rate dictates available
time for these events and their interactions.
Fig. 10Microstructures of the HR steel continuously heated at
(a) 1 K/s and (b) 300 K/s followed by water quenching from 1023 K
(750 C) (F: ferrite, M: martensite).
1552VOLUME 42A, JUNE 2011 METALLURGICAL AND MATERIALS TRANSACTIONS A
In the HR samples, an increase in austenite formation
start and nish temperatures, the Ac
1
and Ac
3
temper-
atures (Figure 8(a)), can then be understood in the
aforementioned context. When increasing the heating
rate, increasingly less time is available for austenite
formation as a thermally activated process to take place,
which leads to a shift to higher transformation temper-
atures that is well documented in the literature.
[26]
This
increase of the superheating for austenite formation
results in an increased nucleation site density for
austenite grains such that a ner network of austenite
grains forms (Figure 10).
In the CR steel, at low heating rates, e.g., 1 K/s, there
is sucient time for recrystallization of ferrite to take
place before austenite formation starts. This can be
observed in Figure 3(b) in which fully recrystallized
grains exist without any sign of austenite formation, i.e.,
presence of martensite at room temperature. In this
scenario, austenite forms at ferritecementite interfaces
and then grows into ferrite grains. Nucleation at ferrite
grain boundaries as a competitive event results in
formation of a network of austenite at higher temper-
atures. This trend suggest similarities between micro-
structural evolution during intercritical annealing in
both HR and CR samples heated at lower heating rate,
e.g., 1 K/s (Figures 6 and 10).
However, with increasing the heating rate in the CR
steel, the possibility of having unrecrystallized ferrite
grains at the beginning of austenite formation increases.
The extent of this increase is proportional to the
employed heating rate. These unrecrystallized grains
with a rather high stored energy will then be suitable
places for austenite formation. Furthermore, frag-
mented pearlite lamellae and/or spheroidized carbides
provide an increased nucleation density compared with
HR samples. Similar to the HR samples, increasing the
heating rate in the CR samples results in a delayed
austenite formation as a thermally activated process.
The balance of these eects leads to a similar dilation
response independent of heating rate in the CR samples
at the early stages of austenite formation, i.e., the Ac
1
temperature is essentially independent of heating rate
(Table II). Because a network of austenite grains does
not exist at ferrite grain boundaries, the dominant
growth mechanism at high heating rates will be the one
that requires short-range carbon redistribution, i.e.,
austenite nucleates at the interfaces of pearlite colonies
that are elongated in the rolling direction, transforms
these colonies, and then grows laterally into the unre-
crystallized ferrite grains that are mostly located
between these colonies. Growth of austenite grains
nucleated at ferrite grain boundaries will then be
hampered because of limited carbon supply. In this
scenario, austenite grains will inherit the distribution of
pearlite colonies elongated in the rolling direction
(Figure 11(d)). This leads to a transition from a random
Fig. 11Microstructural evolution of the CR steel continuously heated at 300 K/s followed by water quenching from (a) 973 K (700 C), (b, c)
1013 K (740 C) and (d) 1053 K (780 C) (F: ferrite, M: martensite, P: pearlite).
METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 42A, JUNE 20111553
distribution of martensite islands for lower heating rates
to a banded distribution for higher heating rates,
compare Figures 6(b) and 11(d). As shown in Figure 15,
the formation of banded austenite structures from the
pearlite colonies is pronounced particularly at the
highest heating rate employed, i.e., 900 K/s.
Huang et al.
[12]
also observed a similar morphological
shift in a Mo-alloyed DP steel in which increasing the
heating rate from 1 K/s to 100 K/s resulted in banded
structures. Their explanation for the transition was
based on concurrent recrystallization of ferrite grains
and austenite formation. They speculated that recrys-
tallization of ferrite grains during austenite formation
encourages austenite grains to nucleate and grow on
deformed pearlite colonies elongated in the rolling
direction. Subsequent growth takes place via rapid
lengthening and thickening of the austenite grains rather
than nucleation on nonstationary ferrite grain bound-
aries. However, at low heating rates, austenite has the
possibility of competitive formation on both ferrite
grain boundaries as well as pearlite colonies leading to
an equiaxed distribution of austenite grains. They
showed that in their steel, approximately 90 pct of
ferrite grains remained unrecrystallized at 100 K/s at the
beginning of austenite formation, whereas this volume
fraction is much lower, 15 pct, in the steel used in this
study even at 900 K/s because of the relatively lean steel
chemistry. Perhaps a heating rate of several thousand
degrees Kelvin per second is needed to view a signicant
overlap of recrystallization and austenite formation in
the current steel. A morphological change via rapid
heating was also reported by Grange.
[14]
They achieved
brous DP steels via heating of initially CR ferrite-
pearlite/martensite structures.
The kinetics of spheroidization of cementite lamellae
is controlled mainly by the diusion of carbon atoms.
Essentially, the spheroidization of pearlite colonies can
be divided into fragmentation of carbide particles,
rounding o the sharp edges and then growth of
particles via Ostwald ripening and shape coarsening.
These stages can proceed simultaneously.
[27]
In the HR
structure, the time available for the range of heating
rates employed in this study is not sucient for these
processes to take place. Annealing times of the order of
hours just below the eutectoid temperature are needed
for complete spheroidization of cementite lamellae in the
HR steel.
[28]
But, in the CR samples, the morphology of
cementite lamellae has been modied by cold rolling.
Fragmentation and bending of cementite lamellae as
well as introducing crystal defects such as vacancies and
dislocations into the ferrite grains and ferritecementite
interfaces as fast diusion paths stimulate the sphero-
idization process. Thus, it is likely to observe this
process to take place in the CR samples. At low heating
rates, e.g., 1 K/s, sucient time is available for almost
complete spheroidization to occur before austenite
formation. Cementite particles can even coarsen to
larger carbide particles during the process, whereas at
high heating rates, cementite lamellae remain frag-
mented without any subsequent coarsening. This
explains the ner distribution of carbide particles at
300 K/s as compared with 1 K/s.
Austenite grain renement in the CR samples via
rapid heating can be related to an increase in austenite
nucleation site density. These nucleation sites include
pearlite colonies and ferrite grain boundaries
(Figures 11(b) and (c)). Deformed pearlite colonies
provide suitable nucleation sites for austenite grains.
Fig. 12Revealing austenite grains in the CR samples heated at
(a) 1 K/s and (b) 300 K/s to just above the corresponding Ac
3
tem-
peratures followed by water quenching.
Fig. 13Revealing cementite particles inside martensite islands in
the CR sample heated at 300 K/s to 1013 K (740 C).
1554VOLUME 42A, JUNE 2011 METALLURGICAL AND MATERIALS TRANSACTIONS A
Furthermore, a large population of ferrite grain bound-
aries provides suitable sites for austenite nucleation.
Simultaneously, rapid heating prevents extensive growth
of austenite grains in the intercritical annealing region.
The renement of austenite grains via rapid heating was
also reported by Andrade-Carozzo and Jacques.
[17]
Lesch et al.
[29]
employed this idea to develop ultrane
grained ferrite structure via rapid transformation
annealing in low carbon steels. The process is essentially
based on rapid heating of CR structures just above the
Ac
3
temperature followed by rapid cooling to room
temperature.
The relative independence of austenite fraction in the
CR steel on heating rate in comparison with the HR
counterpart (Figure 9) can be understood based on the
initial structures prior to austenite formation in both
materials. In the HR structure, this initial structure
remains unchanged. Thus, increasing the heating rate
raises superheating until a network of austenite grains
forms at ferrite grain boundaries. In the CR sample,
however, the situation is dierent. Similarly, increasing
the heating rate tends to increase the superheat, but the
initial structure prior to austenite formation will also be
a function of the heating rate. In ferrite, it changes from
a fully recrystallized structure to a partially recrystal-
lized one as the heating rate increases, and in pearlite, it
delays the spheroidization of carbide particles. Thus,
nucleation and growth scenarios for austenite formation
will change accordingly. The balance of these two eects
will then aect the kinetics of austenite formation
as a function of heating rate. However, when rst
recrystallizing the CR sample, i.e., with the CR+REX
Fig. 14(a) Dilatation curves and their rst derivatives for CR+REX samples (heated at 1 K/s heating rate up to 943 K (670 C) followed by
heating into austenite single-phase region with dierent heating rates), (b) austenite fraction as a function of temperature for dierent heating
rates, and PE austenite fraction, (c) comparison of Ac
1
temperatures at dierent heating rates for HR, CR and CR+REX samples.
Fig. 15Microstructure of the CR steel continuously heated at
900 K/s followed by water quenching from 983 K (710 C) (F: fer-
rite, M: martensite, P: pearlite).
METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 42A, JUNE 20111555
structure, the similarity with the dilation response of the
HR sample is restored (Figure 14), as the initial struc-
ture in both materials mainly consists of recrystallized
ferrite. The slightly dierent transformation start tem-
peratures (Figure 14(c)) might be attributable to par-
tially spheroidized pearlite lamellae in the CR+REX
samples. Because austenite nucleation commences at the
interface of ferrite and cementite in the pearlite colonies,
the broken lamellae provide more interfaces, and
consequently, the possibility for austenite nucleation
increases resulting in lower Ac
1
temperatures in
the CR+REX samples compared with their HR
counterparts.
V. CONCLUSIONS
In this study, austenite formation in a HR and
CR plain low-carbon steel was investigated systemati-
cally using dilation experiments and microstructural
characterization. A variety of scenarios can be rational-
ized to explain the overall rate of austenite formation
considering the overlap between austenite formation,
recrystallization of ferrite-pearlite structure, and sphero-
idization of Fe
3
C. The following can be concluded from
this study.
The dilation response can, in the HR material, be
subdivided into several stages: (1) thermal expansion of
ferrite/pearlite structure, (2) pearlite-to-austenite trans-
formation, (3) ferrite-to-austenite transformation, and
(4) austenite thermal expansion. Additionally, recrystal-
lization of ferrite and spheroidization of pearlite can
take place in the CR sample. Recrystallization of ferrite
results in a change in the dilation response of the CR
steel.
Increasing the heating rate shifts the austenite
formation to higher temperatures in the HR sample
and results in formation of a network of ner austenite
grains. Increasing the heating rate in the CR material,
however, results in formation a variety of initial
structures ranging from fully recrystallized to partially
recrystallized structures prior to austenite formation.
This transition can be monitored using the dilation
data. Furthermore, the degree of spheroidization of
carbide lamellae in pearlite colonies is reduced with
increasing the heating rate. As a result, a morphological
shift is observed from randomly distributed to a
banded structure of austenite. It can be noted that
the banded feature of austenite could be controlled
theoretically by the processing parameters such as the
level of cold work and the heating rate. A higher
heating rate also resulted in austenite grain renement
in CR steels.
ACKNOWLEDGMENTS
The authors would like to acknowledge the Natu-
ral Sciences and Engineering Research Council of
Canada (NSERC) for their nancial support. We are
grateful to J.D. Embury for valuable discussions and
suggestions.
APPENDIX
To examine the eect of dierent factors on the
deviation from linearity in the dilation curves observed
before austenite formation, see Section IIIB; additional
investigations were carried out. Cold rolling can intro-
duce residual stresses that are concentrated on the
surface of the samples.
[30]
Relieving residual stresses
imposed by cold rolling can occur prior to austenite
formation in the CR steel. It can also contribute to the
dimensional changes in the samples.
[31]
However, reduc-
tion of residual stresses via thinning the samples down to
half thickness by polishing shows no sign of change in
the deviation. Furthermore, dilatometry of 80 pct CR
prespheroidized samples (annealed for 24 hours at
963 K [690 C] before cold rolling) revealed a dilation
response similar to that of the CR sample. Thus,
spheroidization was then ruled out to contribute to
dilation changes. A 80 pct CR Ti-added IF steel with the
chemical composition given in Table A1 was used to
supplement the dilation study during heating for a case
where no cementite is present, i.e., any eect from
pearlite can be excluded. The initial microstructure
consists of elongated ferrite grains. To examine the start
temperature of ferrite-to-austenite transformation, a
sample was heated into the single austenite phase region
at 1 K/s, and an Ac
1
temperature of 1193 K (920 C)
was measured. As shown in Figure A1, a thermal cycle
was then employed to investigate the dilation before
phase transformation. It includes a heating stage (solid
black line), fast He quenching (black dashed line), and
reheating stage of the sample (gray line). The CR sample
was heated rst at 1 K/s to 1123 K (850 C) followed by
rapid He quenching down to 473 K (200 C) and
immediate reheating at 1 K/s to 1123 K (850 C). The
contraction in the dilation curve in the rst heating path
is evident in Figure A1 (similar to the CR low-carbon
steel in Figure 2). The relative change length (DL/L
o
) is
approximately 0.11 pct, which is similar to the CR low-
carbon steel, i.e., 0.1 pct. However, in the next reheating
stage, the dilation curve shows no sign of contraction.
Microstructural observations at 873 K (600 C) and
1123 K (850 C) showed that the microstructure chan-
ged from the CR to a completely recrystallized one
Table A1. Chemical Composition of the IF Steel in Weight Percent
Element Fe C Mn P S Si Ti B N
Wt pct Balanced 0.0026 0.16 0.011 0.008 0.01 0.068 0.0005 0.003
1556VOLUME 42A, JUNE 2011 METALLURGICAL AND MATERIALS TRANSACTIONS A
during the rst heating stage. Thus, a reduction in
dislocation density and texture changes from an alpha
ber to a gamma dominant texture
[10,32]
that occur
during ferrite recrystallization remain as possible expla-
nation for the deviation. A reduction in specic volume
of 0.37 pct for 80 pct cold-drawn steel wires during
annealing and before austenite formation was reported
by Gridnev et al.
[33]
With an assumption of isotropic
change in length during dilation, one can estimate DL/
L
o
through DL/L
o
DV/3V
o
for a small relative volume
change in which DL is the length change, L
o
is the initial
length, DV is the volume change, and V
o
is the initial
volume of the sample. Thus, the value for DL/L
o
concluded from the work by Gridnev et al.
[33]
is
0.12 pct, which is in good agreement with the obser-
vations made in this study, i.e., 0.1 pct for the low-
carbon steel and 0.11 pct for the IF steel. Gridnev
et al.
[33]
attributed this observation to annihilation of
point and line defects as well as microcracks. De Cock
et al.
[10,32]
also showed that the reduction in dislocation
density during recrystallization can account for the
amount of dilation (0.05 pct) that is observed during
heating of 80 pct CR low and ultralow carbon steels
before austenitization at 10 K/s. Thus, it is concluded
that the deviation from linear thermal expansion is
related to recrystallization of the sample during heating
and is consistent with the volume change caused by a
reduction in the average dislocation density.
REFERENCES
1. G.R. Speich, V.A. Demarest, and R.L. Miller: Metall. Trans. A,
1981, vol. 12A, pp. 141928.
2. N.C. Law and D.V. Edmonds: Metall. Mater. Trans. A, 1980,
vol. 33A, pp. 3346.
3. M. Hillert, K. Nilsson, and L.-E. To rndahl: J. Iron Steel Res. Int.,
1971, vol. 209, pp. 4965.
4. C.I. Garcia and A.J. Deardo: Metall. Mater. Trans. A, 1981,
vol. 12A, pp. 52130.
5. R.R. Judd and H.W. Paxton: Trans. TMS-AIME, 1968, vol. 242,
pp. 20614.
6. D. San Mart n, T. de Cock, A. Garc a-Junceda, F.G. Caballero, C.
Capdevilla, and C. Garc a de Andre s: Mater. Sci. Technol, 2008,
vol. 24, pp. 26672.
7. D.Z. Yang, E.L. Brown, D.K. Matlock, and G. Krauss: Metall.
Trans. A, 1985, vol. 16A, pp. 138592.
8. C.I. Garcia and A.J. Deardo: Metall. Trans. A, 1981, vol. 12A,
pp. 52130.
9. J.J. Yi, I.S. Kim, and H.S. Choi: Metall. Trans. A, 1985, vol. 16A,
pp. 123745.
10. T. de Cock, C. Capdevila, F.G. Caballero, and C. Garc a de
Andre s: Scripta Mater, 2006, vol. 54, pp. 94954.
11. V.I. Savran, Y. Van Leeuwen, D.N. Hanlon, C. Kwakernaak,
W.G. Sloof, and J. Sietsma: Metall. Mater. Trans. A, 2007,
vol. 38A, pp. 94655.
12. J. Huang, W.J. Poole, and M. Militzer: Metall. Mater. Trans. A,
2004, vol. 35A, pp. 336375.
13. M. Mazinani and W.J. Poole: Metall. Mater. Trans. A, 2007,
vol. 38A, pp. 32839.
14. R.A. Grange: 2nd Int. Conf. on the Strength of Metals and Alloys,
ASM, Materials Park, OH, 1970, pp. 86176.
15. R. Petrov, L. Kestens, and Y. Houbaert: ISIJ Int., 2001, vol. 41,
pp. 88390.
16. L. Kestens, A.C.C. Ries, W.J. Kaluba, and H. Houbaert: Mater.
Sci. Forum, 2004, vols. 467470, pp. 28792.
17. V. Andrade-Carozzo and P.J. Jacques: Mater. Sci. Forum, 2007,
vols. 539543, pp. 464954.
18. D. Quidort and Y.J.M. Brechet: ISIJ Int., 2002, vol. 42, pp. 1010
17.
19. F.S. LePera: J. Met., 1980, vol. 32, pp. 3839.
20. J.R. Vilella: Metallographic Techniques for Steels, ASM, Materials
Park, OH, 1938, p. 36.
21. M. Onink, C.M. Brakman, F.D. Tichelaar, E.J. Mittemeijer, S.
van der Zwaag, J.H. Root, and N.B. Konyer: Scripta Metall.
Mater., 1993, vol. 29, pp. 101116.
22. G.R. Speich and A. Szirmae: Trans. TMS-AIME, 1969, vol. 245,
pp. 106374.
23. C.C. Chang: Characterization of Solid Surfaces, P.F. Kane and
G.G. Larrabee, eds., Plenum Press, New York, NY, 1974,
pp. 53437.
24. C-S Oh, H.N. Ham, C.G. Lee, T-H. Lee, and S-J. Kim: Met.
Mater. Int., 2004, vol. 10, pp. 399406.
25. C-M. Li, F. Sommer, and E. Mittemeijer: Z. Metall., 2000, vol. 91,
pp. 59.
26. C.R. Brooks: Principles of the Austenitization of Steels, University
Press, Cambridge, UK, 1992, p. 101.
27. Y.L. Tian and W. Kraft: Metall. Trans. A, 1987, vol. 18A,
pp. 140314.
28. S. Chattopadhyay and M. Sellars: Acta Mater., 1982, vol. 30,
pp. 15770.
29. C. Lesch, P. Alvarez, W. Bleck, and J.G. Sevillano: Metall. Mater.
Trans. A, 2007, vol. 38A, pp. 188290.
30. I. Nikitin and M. Besel: Scripta Mater., 2008, vol. 58, pp. 239
42.
31. T. Hasegawa and T. Yakou: Acta Metall., 1982, vol. 30, pp. 235
43.
32. T. de Cock, C. Capdevila, J.P. Ferrer, F.G. Caballero, J.A.
Jime nez, and C. Garc a de Andre s: Mater. Sci. Technol., 2008,
vol. 24, pp. 83236.
33. V.N. Gridnev, V.G. Gavrilyuk, and Y.Y. Meshkov: Metall. Sci.
Heat Treat., 1971, vol. 13, pp. 2022.
Fig. A1Dilatation curves for IF steel during heating-cooling-heat-
ing cycles. The cooling rate during rapid cooling was approximately
200 K/s.
METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 42A, JUNE 20111557

Você também pode gostar