You are on page 1of 14

Effect of Composition and Deformation on Coarse-Grained

Austenite Transformation in Nb-Mo Microalloyed Steels


N. ISASTI, D. JORGE-BADIOLA, M.L. TAHERI, B. LÓPEZ, and P. URANGA

Thermomechanical processing of microalloyed steels containing niobium can be performed to


obtain deformed austenite prior to transformation. Accelerated cooling can be employed to
refine the final microstructure and, consequently, to improve both strength and toughness. This
general rule is fulfilled if the transformation occurs on a quite homogeneous austenite micro-
structure. Nevertheless, the presence of coarse austenite grains before transformation in dif-
ferent industrial processes is a usual source of concern, and regarding toughness, the coarsest
high-angle boundary units would determine its final value. Sets of deformation dilatometry tests
were carried out using three 0.06 pct Nb microalloyed steels to evaluate the effect of Mo alloying
additions (0, 0.16, and 0.31 pct Mo) on final transformation from both recrystallized and
unrecrystallized coarse-grained austenite. Continuous cooling transformation (CCT) diagrams
were created, and detailed microstructural characterization was achieved through the use of
optical microscopy (OM), field emission gun scanning electron microscopy (FEGSEM), and
electron backscattered diffraction (EBSD). The resultant microstructures ranged from polygo-
nal ferrite (PF) and pearlite (P) at slow cooling ranges to bainitic ferrite (BF) accompanied by
martensite (M) for fast cooling rates. Plastic deformation of the parent austenite accelerated
both ferrite and bainite transformation, moving the CCT curves to higher temperatures and
shorter times. However, an increase in the final heterogeneity was observed when BF packets
were formed, creating coarse high-angle grain boundary units.

DOI: 10.1007/s11661-011-0624-0
Ó The Minerals, Metals & Materials Society and ASM International 2011

I. INTRODUCTION ical processes to microalloyed grades. When controlled


rolling is applied to Nb microalloyed steels, the austenite
VALUE-ADDED applications of microalloyed steels is rolled at temperatures low enough to produce strain-
continue to be developed to meet the increasing material induced precipitation of carbonitrides during deforma-
demands of a variety of structural applications. Most of tion to inhibit recrystallization.[5] As is well known, the
the required properties can be obtained via two main addition of Nb as an alloying element can retard or
strategies: chemical alloying and optimization of the inhibit recrystallization due to two mechanisms: the
processing parameters. Suitable combinations of alloy- solute drag effect owed to Nb atoms in solid solution
ing element additions, such as Nb, Ti, Mo, or B, and the pinning effect due to strain-induced precipita-
contribute to an increase in strength directly through tion, the latter usually exerting the strongest effect.[6,7]
microstructural refinement, solid solution strengthening, Some authors reported that the reduction of austenite
and precipitation hardening, as well as, indirectly, grain size and an increase in retained strain provide a
through enhanced hardenability and associated modifi- higher density of nucleation sites, which leads to a
cation of the resultant microstructure.[1] Nevertheless, refinement of the final transformed microstructure[8,9]
toughness may be impaired depending on the selected and a modification of the balance between the different
processing strategy followed to achieving the strength ferrite morphologies.
requirements. Nowadays, combinations of high resis- Occasionally, rolled products show heterogeneous
tance and high toughness are required for applications austenite structure prior to transformation. For exam-
such as gas and oil transportation pipes, offshore ple, the presence of microstructural banding in rolled
facilities, and naval technologies.[2–4] plates was reported as an origin for final heterogeneities
Reducing grain size improves both strength and in the microstructure in low-carbon microalloyed
toughness, and the best way to obtain fine microstruc- steels.[10] The microstructural bands are coarse bainitic
tures in the final product is by applying thermomechan- phases formed in a uniform matrix of ferrite and are
transformed from coarse austenite grains, which are
N. ISASTI, Graduate Student, and D. JORGE-BADIOLA, developed when conditions favor abnormal grain
B. LÓPEZ and P. URANGA, Researchers, are with CEIT and growth or strain-induced grain boundary migration in
TECNUN (University of Navarra), Donostia-San Sebastian, 20018, the rolling passes.[11,12] It is also possible to have
Basque Country, Spain. Contact e-mail: puranga@ceit.es M.L. inappropriate austenite refinement before strain begins
TAHERI, Hoeganaes Assistant Professor of Metallurgy, is with
Materials Department, Drexel University, Philadelphia, PA 19104. accumulating, in some cases, as a consequence of a low
Manuscript submitted November 16, 2010. soaking temperature[13] or in others due to microalloy-
Article published online February 12, 2011 ing segregation.[14] Also, in thin slab direct rolling

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 42A, DECEMBER 2011—3729


technologies, the refinement and conditioning of aus- morphology, the difference is attributed to the insuffi-
tenite prior to transformation may be limited. The initial cient diffusion of carbon to develop continuous lamel-
as-cast coarse grain sizes are present at the entry of the lae. Also, when degenerated P is formed during
first stand. Together with the small total reductions, it continuous cooling, it involves transformation of M/A
makes it difficult to reach enough refinement before island to degenerated P along with massive ferrite
transformation;[15] this problem is most notorious when transformation.
final gages thicker than 10 mm are produced. In all these The transformation temperatures for the formation of
cases, the transformation behavior of these coarse grains BF, which consists of packets of parallel ferrite laths or
will be completely different from those corresponding to plates separated by low-angle boundaries and contain-
the surrounding finer grains. In a recent work, it was ing very high dislocation densities, are in the interme-
observed that a higher fraction of coarse grain sizes in diate temperature range. Although the austenite
the microstructure increases the chance of finding those decomposition is only to ferrite, coexisting with retained
grains at the cleavage origin, leading to a wider scatter in austenite or M/A constituent, the microstructural
fracture stress in comparison to more homogeneous arrangement of acicular-shaped ferrite crystals in groups
grain size distributions.[14] of parallel laths is included in bainite classification. The
Various terms were proposed to name the transfor- most prominent features of the microstructure are the
mation microstructures. Those terms mostly used are by aligned, elongated, parallel islands of retained austenite
Bramfitt and Speer[16] and the ISIJ Bainite Commit- or M/A constituent within the prior austenite grains and
tee.[17] In the present article, the ISIJ Bainite Committee the conservation of the austenite grain boundary. The
notation will be adopted, as phases observed fit better to structure consists of many fine, elongated ferritic crys-
proposed ones. The phases are identified as polygonal tals with a high dislocation density.[25]
ferrite (PF), lamellar pearlite (P), degenerated pearlite Dilatometry tests have been widely used in order to
(DP), quasi-polygonal ferrite (QF), granular ferrite obtain in-situ information concerning austenite decom-
(GF), bainitic ferrite (BF), and martensite (M). position kinetics in steels.[29,30] Dilatometry registers
PF nucleates as grain boundary allotriomorphs and is length changes that occur during heat treatment of a
characterized by very low dislocation densities and the sample. This information combined with microstruc-
absence of substructure. Growth of polygonal equiaxed tural analysis is the basis for CCT diagram determina-
ferrite is controlled by rapid substitutional atom transfer tion. Taking into account all the information gathered
across partially coherent boundaries and long-range from dilatometry tests and microstructural character-
diffusion of carbon atoms, which are rejected from the ization, this article undertakes a detailed study of the
growing ferrite.[18–20] transformation products in coarse austenite grains,
QF grains have irregular grain boundaries and often analyzing the effect of composition and austenite defor-
show etching evidence of substructure. Similar to PF, mation in low carbon niobium microalloyed steels
QF nucleates heterogeneously at the boundaries of the containing different levels of molybdenum.
austenite grains. The transformation can be accom-
plished by short-range diffusion across transformation
interfaces.[21] However, interstitial or substitutional II. MATERIALS AND EXPERIMENTAL
atom partitioning may occur at the migrating interfaces, PROCEDURE
causing irregular growth and jagged boundaries of
massive ferrite crystals.[22] QF also contains high dislo- The chemical compositions of the steels studied are
cation densities, dislocation sub-boundaries, and even listed in Table I. Uniaxial compression tests were
martensite-austenite (M/A) constituents.[23,24] performed in a Bähr DIL805A/D quenching and defor-
GF, or granular bainite, consists of sheaves of mation dilatometer (Bähr-Thermoanalyse GmbH, Hull-
elongated ferrite crystals with low misorientations and horst, Germany), where solid cylinders (5-mm diameter,
a high dislocation density,[25] generally containing 10-mm long) were employed. The steels were subjected
roughly equiaxed islands of M/A microconstituent. to simulated recrystallization (cycle A) and nonrecrys-
The carbon that is partitioned from the BF stabilizes tallization (cycle B) deformation schedules depicted
the residual austenite so that the final microstructure schematically in Figure 1, followed by controlled cool-
contains both retained austenite and some high carbon ing at constant rates in the range 0.1 K/s to a maximum
M;[26] the dispersed particles have granular or equiaxed cooling of 200 K/s. Both schedules (cycles A and B)
morphology. There is evidence that the GF crystals included a simulated roughing step performed at a
grow by a diffusion-controlled ledge mechanism.[27] The temperature of 1423 K (1150 °C) using a strain of 0.30
formation of different second phases is directly con- and a strain rate of 1 s1, followed by postdeformation
nected with the effect of alloying elements on the ferrite holding at the same temperature for 12 seconds to
reaction, which in turn determines the size and carbon ensure complete recrystallization. Furthermore, the
content of the last transformed austenite. M/A constit- simulated nonrecrystallization schedule (cycle B) con-
uents are the most common feature of granular bai- tained an additional finishing step performed at a
nite.[28] M/A either forms small islands, distributed temperature of 1173 K (900 °C), which was applied in
between ferrite grains, or covers the surfaces of larger the nonrecrystallization temperature region using a
bainitic grains, which develop a steep carbon gradient. strain of 0.4 at 1 s1.
Similar to lamellar P, degenerated P is formed at low The microstructures generated at the end of the cycles
cooling rates by diffusion processes, and because of its were characterized after etching in 2 pct Nital by optical

3730—VOLUME 42A, DECEMBER 2011 METALLURGICAL AND MATERIALS TRANSACTIONS A


Table I. Chemical Compositions of Steels (Weight Percent)

Steel C Mn Si Nb Mo Al N
6NbMo0 0.05 1.56 0.05 0.06 0.01 0.028 0.004
6NbMol6 0.05 1.6 0.05 0.061 0.16 0.03 0.005
6NbMo31 0.05 1.57 0.05 0.059 0.31 0.031 0.005

Fig. 1—Schematics of thermomechanical schedules performed at the


dilatometer for both cycle A (transformation from recrystallized aus-
tenite) and cycle B (transformation from deformed austenite).

microscopy (OM, Olympus IX2-UCB, Olympus Cor-


poration, Tokyo, Japan), SEM (PHILIPS* XL30CP),
Fig. 2—Equivalent strain distribution in the central cross section of
the dilatometry sample for the cycle A deformation pass
*PHILIPS is a trademark of FEI Company, Hillsboro, OR. (T = 1423 K (1150 °C), e = 0.3, and e_ ¼ 1 s1 ). The section at
which the microstructural observation and hardness tests were per-
formed is also marked.
and field emission gun scanning electron microscopy
(FEGSEM, JEOL** JSM-7000F). The dilatometry
Austenite grain structure was analyzed in samples
quenched at 1173 K (900 °C). Samples were etched
**JEOL is a trademark of Japan Electron Optics Ltd. using saturated picric acid and HCl. Average recrystal-
lized austenite grain size prior to transformation was
samples were sectioned along their longitudinal axis, measured using the mean intercepted length method.
selecting the region corresponding to a maximum area Recrystallized austenite mean grain sizes of Dc =
fraction of nominal strain and reduced strain gradient. 119 ± 8 lm, Dc = 125 ± 8 lm, and Dc = 84.8 ±
Figure 2 shows the equivalent strain distribution in 5 lm were obtained for steels 6NbMo0, 6NbMo16,
the central cross section calculated using Abaqus/ and 6NbMo31, respectively.
Explicit[31] for a dilatometry sample deformed 0.3 at
1 s1, under the same conditions as for the cycle A
schedule. The observation plane is also drawn on the III. RESULTS
image.
Selected samples were prepared for electron backscat- A. Microstructures
tered diffraction (EBSD) observations. The specimens The microstructures of the specimens cooled at every
were polished down to 1 lm and the final polishing was cooling rate were analyzed by OM for both schedules
with colloidal silica. Orientation imaging was carried and three steels. As an example, and for the case of steel
out on the PHILIPS XL30CP scanning electron micro- 6NbMo0, the optical micrographs of the specimens
scope with W-filament, using TSL (TexSEM Laborato- cooled at 0.1, 0.5, 10, and 100 K/s are shown in Figure 3
ries, Provo, UT) equipment. For some specific samples, for both recrystallized (schedule A) and unrecrystallized
higher resolution orientation imaging analysis was (schedule B) conditions of the parent austenite phase
carried out on the JEOL JSM-7000F FEGSEM using prior to transformation.
HKL Channel5 EBSD. Different scan step sizes were A gradual shift from low cooling rate phases such as
defined depending on the resolution needed, varying PF and P (0.1 K/s) to high rate phases such as BF and
from 0.2 lm for high resolution scans to 2 lm for M (100 K/s) is observed with a transition through
general microstructural characterization. Vickers hard- microstructures composed by QF and GF (10 K/s). For
ness measurements were made using a 1-kg load and are the slowest cooling rate of 0.1 K/s, the microstructure is
reported as an average of 10 measurements. formed by PF and P. The significant refinement of the

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 42A, DECEMBER 2011—3731


Fig. 3—Light micrographs of the microstructures transformed from (a) through (d) recrystallized and (e) through (h) deformed austenite at steel
6NbMo0. 0.1 K/s: (a) PF + GF and (e) PF + P. 0.5 K/s: (b) PF + QF + GF and (f) PF + GF. 10 K/s: (c) QF + GF and (g) QF + GF. 100 K/s:
(d) BF + M and (h) BF + M + QF + GF.

3732—VOLUME 42A, DECEMBER 2011 METALLURGICAL AND MATERIALS TRANSACTIONS A


ferrite grain size when transformation has occurred from
unrecrystallized austenite should be noted. The opposite
condition corresponds to the highest cooling rate of
100 K/s, where a mixture of bainite and M is obtained in
both conditions, although the amount of M formed
from the recrystallized austenite is larger. For interme-
diate cooling rates, 0.5 K/s, for example, the micro-
structure is a mixture of PF and GF in the
unrecrystallized austenite; an important fraction of
GF, together with an increase in quasi-polygonal ferrite
fraction, is formed in the recrystallized condition.
After a 10 K/s cooling rate, the microstructure is
constituted by QF and GF, for both schedules. The
morphology of bainite is changed as a function of the
cooling rate. By increasing the cooling rate, the domi-
nant morphology of bainitic structures progressively
changes from granular ferrite (GF) to BF. It should be
noted that, in the deformed austenite, the GF morphol-
ogy predominates over the BF and that the grains are
distributed in a more random manner; i.e., the prior
austenite grain boundary networks are less evident
compared to recrystallized austenite.
Figure 4 shows several FEGSEM micrographs of
some microstructural features, which cannot be cor-
rectly shown using optical microscopic images. In
Figure 4(a), which corresponds to deformed austenite
cooled at 0.5 K/s, a mixture of PF and GF is shown.
The characteristic equiaxed shape of ferrite crystals and
a mixture of M/A islands in GF and cementite carbides
are observed. Degenerated P is also detected in some of
the PF/GF interphases. Figure 4(b) shows a mixture of
quasi-polygonal ferrite and GF in a sample correspond-
ing to deformed austenite cooled at 10 K/s. The
presence of M/A islands, instead of carbides, has been
confirmed by means of diffraction patterns using EBSD
technique. In Figure 4(c), cooled at 100 K/s, a mixture
of M and bainite (mainly BF but still with an important
fraction of GF) is represented. This BF microstructure
shows the characteristic elongated ferrite laths separated
by elongated M/A islands.

B. Dilatometry Tests
For the dilation curves to be interpreted, a careful and
systematic analysis of resulting data is needed. This
analysis has to be made in conjunction with optical
micrographs and detailed scanning electron microscopic
pictures. Figure 5(a) depicts the measured dilation
curves, represented as DL/L0 vs temperature, for several
cooling rates (0.1, 0.5, 10, and 100 K/s) obtained from
cycle B tests performed with steel 6NbMo31. Fig- Fig. 4—FEGSEM micrographs of microstructures corresponding to
ure 5(b) shows the evolution of the transformed fraction steel 6NbMo0 and schedule B: (a) 0.5 K/s, (b) 10 K/s, and (c)
derived from dilatometry curves in Figure 5(a) once the 100 K/s.
lever rule is applied.[32] This rule is based on extrapo-
lating the linear expansion behavior from the temper-
ature regions where no transformation occurs, and cooling rate, the transformed fraction curve exhibits
subsequently assuming proportionality between the several slope changes related to transitions from PF to
fraction of decomposed austenite and the observed GF and DP phases. For other cooling rates, and even if
length change. Depending on the cooling rate, the shape the optical micrographs show that the microstructures
of the curve and the transformation temperatures vary, are composed by a mixture of different phases, the
lowering the transformation start temperature as the evolution of the transformed fraction does not show
cooling rate increases. For the particular case of 0.5 K/s clearly the beginning and end of each transformation.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 42A, DECEMBER 2011—3733


Fig. 5—(a) Dilatometry curves and (b) evolution of transformed
fraction in steel 6NbMo31, cycle B (deformed austenite) and selected Fig. 6—(a) and (b) Transformation rate as a function of temperature
cooling rates. calculated from tests in Fig. 5 for steel 6NbMo31, cycle B (deformed
austenite) and selected cooling rates.

Therefore, for the temperature range and fraction of reduces the transformation start-finish range as the
each phase formation to be determined, additional cooling rate increases. The transformation rate is much
information can be extracted if the slope of the lower for PF than for BF, which indicates the higher
transformed fraction (dftrans/dt) is analyzed. Figures 6(a) growth rate of the bainitic laths, characteristic for
and (b) show transformation rates as a function of the displacive transformations.
temperature for the tests plotted in Figure 5. At a
cooling rate of 0.1 K/s (Figure 6(a)), the transformation
rate curve has a single peak related to the formation of C. Hardness Measurements
PF. For the 0.5 K/s cooling rate, two different peaks Measurements of the bulk Vickers hardness of each
corresponding to the formation of PF and GF are specimen were made and each value is the average of 10
detected. In the case of 10 K/s, the curve has a single individual measurements performed in the longitudinal
peak corresponding to the formation of the mixture of section marked in Figure 2. These data are summarized
QF and GF. Comparing these four curves, similar start in Figure 7, where hardness values are plotted for each
temperatures are observed for PF formation (0.1 and steel and the deformation schedule studied as a function
0.5 K/s curves) and GF formation (0.5 and 10 K/s of the transformation start temperature. As expected,
cooling rates). For higher cooling rates (e.g., 100 K/s, the lowest hardness values correspond to the micro-
Figure 6(b)), the transformation rate curve shows structures composed by PF, which are formed at
a sharp peak corresponding to the bainite and M the highest transformation temperatures. Conversely,
formations. This peak increases in value, i.e., higher the highest hardness values were displayed by the
transformation rates, shifts to lower temperatures, and microstructures composed by BF and M, formed at

3734—VOLUME 42A, DECEMBER 2011 METALLURGICAL AND MATERIALS TRANSACTIONS A


GF. For these structures, the hardness in samples of cycle
B is higher than for samples of cycle A. Finally, for the
highest cooling rates, samples transformed from recrys-
tallized austenite show higher hardness values due to the
higher M fractions present.

IV. DISCUSSION
A. CCT Diagrams
The information obtained from dilatometry tests,
micrographs, and hardness measurements was used to
build continuous cooling transformation (CCT) dia-
grams for both schedules and three steels. In addition to
phase stability regions, cooling rates and Vickers hard-
ness values are presented in those CCT diagrams
(Figure 8). The transformation start and finish temper-
Fig. 7—Vickers hardness values plotted as a function of the trans- atures were determined as the 5 and 95 pct transformed
formation temperature for the three steels studied. Open and filled
symbols correspond to the transformation products formed from the
fractions, defined from curves such as the ones shown in
recrystallized and unrecrystallized austenite, respectively. Figure 5(b).
In all CCT diagrams, similar phases are observed. At
the lowest cooling rates, the transformed microstruc-
the lowest transformation temperatures. The hard- tures are composed by PF, GF, and P. At intermediate
ness data obtained for QF and GF combinations, cooling rates, a mixture of QF and GF is formed; and
characterized by intermediate transformation tempera- the transformation sequence during continuous cooling
tures, lay between the aforementioned limits. For the is QF formation followed by GF. Finally, at the highest
three steels examined, there is no significant influence of cooling rates, the transformed microstructures are
steel composition when similar microstructures are composed by BF and M.
compared. In the case of 6NbMo0 and transformation from
The comparison of hardness values plotted as a undeformed austenite (cycle A), Figure 8(a), cooling
function of the cooling rate is not always evident as rates lower than 0.5 K/s promote the formation of a
different microstructures are being compared. Based on mixture of PF and GF. The transformed microstruc-
this reasoning, values in Figure 7 are plotted using tures containing GF and QF could be achieved at a
transformation temperatures, where comparable micro- cooling rate range from 1 to 10 K/s. BF is formed at
structures are roughly approximated by similar trans- cooling rates higher than 20 K/s simultaneously with M.
formation start temperatures. Hardness values for For steel 6NbMo0, the austenite deformation causes an
samples produced from deformed austenite are reported expansion of the transformation field for PF in the CCT
to be systematically higher than those from recrystal- diagram (Figure 8(b)) and causes the formation of
lized austenite.[33] Analyzing the data in Figure 7 (filled lamellar P at the lowest cooling rate (0.1 K/s). Defor-
symbols for deformed austenite and open symbols for mation in the austenite causes a shifting in the trans-
recrystallized austenite), this conclusion is not valid for formation to higher temperatures. The PF formation
the entire range of microstructures generated. occurs at 1080 K (807 °C) at 0.1 K/s (Ar3 displaced
Three different ranges can be defined depending on 34 K compared to undeformed austenite). On the other
transformation start temperature and hardness values. hand, the formation of GF + QF increases until 20 K/s.
For the highest transformation start temperatures A mixture of BF, GF, QF, and M is obtained at the
(>973 K (700 °C), i.e., low cooling rates), most of the cooling rates from 20 to 200 K/s.
points in the graph are from samples of deformed The CCT diagram from recrystallized austenite of
austenite, full symbols. This fact implies that ferritic steel 6NbMo16 (Figure 8(c)) shows some similarities
structures are formed mainly in deformed samples and compared to the CCT of 6NbMo0 from the same
not in the ones transforming from recrystallized austenite. thermomechanical process. In that case, PF ceases to be
Various effects of the deformation of austenite, such as formed from 1 K/s. The addition of Mo causes a
austenite grain refinement, increase in grain boundary shifting in the transformation to lower temperatures and
area per unit volume, Sv, presence of microalloy precip- shorter times. CCT corresponding to 6NbMo16 from
itates, and loss of microalloying elements due to precip- deformed austenite (Figure 8(d)) also shows an increase
itation in austenite, ultimately result in promoting the in PF field at the cooling range from 0.1 to 5 K/s.
austenite/ferrite transformation and a decrease in harde- Moreover, Ar3 shifts to higher temperatures (is displaced
nability. As a consequence, the hardness of the deformed 5 to 15 K). At the cooling rate of 0.1 K/s, DP appears
samples is found to be lower than that of the undeformed instead of lamellar P and the first is favored by coarse-
samples when the structure is mainly ferritic.[34] For grained austenite.[34] For intermediate cooling rates
intermediate transformation start temperatures, 823 K corresponding to the formation of GF, QF, and BF,
to 973 K (550 °C to 700 °C), phases are mainly QF and the diagram exhibits shifting in Bs temperature to higher

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 42A, DECEMBER 2011—3735


Fig. 8—CCT diagrams for the (a), (c), and (e) recrystallized and (b), (d), and (f) deformed austenite for the three steels studied: (a) and (b) steel
6NbMo0, (c) and (d) steel 6NbMo16, and (e) and (f) steel 6NbMo31.

temperatures compared to the case of recrystallized austenite (Figure 8(e)), PF formation along austenite
austenite (Bs is displaced 10 to 50 K). grain boundaries is limited at lower cooling rates
For steel 6NbMo31, the transformed phases and CCT (0.1 K/s and 0.5 K/s) and GF is also formed. The
diagram shape are similar compared to the previously mixture of GF, QF, and BF could be achieved at cooling
analyzed diagrams. For the case of undeformed rates ranging from 10 to 50 K/s. Finally, the formation

3736—VOLUME 42A, DECEMBER 2011 METALLURGICAL AND MATERIALS TRANSACTIONS A


of M is observed at cooling rates higher than 50 K/s. In affected if the Mo level is increased. A useful parameter
the case of 6NbMo31 and deformed austenite (Fig- to evaluate the length of the tail of a distribution is to
ure 8(f)), the diagram shows the same trend previously calculate a critical grain size for a fixed area fraction.
mentioned. Deformation increases the transformation The grain sizes for which 10 pct of the area fraction of
start and finish temperatures, as well as accelerating grains have a size greater than that value (denoted as
bainitic and ferritic transformations. Dc10 pct) are 65, 67, and 77 lm 6NbMo0, 6NbMo16,
Comparing both CCT diagrams, in the case of and 6NbMo31 steels, respectively. Alloying elements are
recrystallized austenite, both the transformation start reported to decrease the nucleation rates of all the ferrite
and finish temperatures were clearly lower than in the morphological types nucleating heterogeneously on the
case of deformed austenite. In the same way, deforma- austenite grain or deformation-induced boundaries.[33]
tion accelerates bainitic and ferritic transformation.
There is no significant change in morphology of the
C. Effect of Deformation
transformation products derived from deformed aus-
tenite compared to the microstructures formed from the In Figure 10, the CCT diagrams developed for steel
recrystallized austenite except for the observed refine- 6NbMo31 for both recrystallized and deformed austen-
ment of the PF and QF grains (Figure 2) and the GF ite are plotted together for comparison; the cooling rates
and BF assemblies. The effect of the strain accumulation are also indicated. The deformation of the prior
on the ferrite refinement is related to the increase of the austenite forces an expansion of the ferrite transforma-
specific grain boundary area and is attributed to a tion field in the CCT diagram. An increase in the ferrite
significant increase in the density of ferrite nucleation transformation start temperature was observed (Ar3).
sites introduced by deformation.[35–38] Similarly, an increase of the bainite transformation start
temperature was also observed in the range of high
cooling rates.
B. Effect of Composition While the increase of Ar3 temperature with deforma-
The effect of molybdenum on the transformed phases tion is well documented, the behavior of Bs is more
and temperatures can be evaluated by comparing the complex. It has been widely accepted that the enhanced
CCT diagrams shown in Figure 8. On one hand, it can PF formation, brought about by the deformation of
be affirmed that there is no clear effect on Ar3, Bs, or Ms austenite, may be largely attributed to an increase in the
temperature, even if a decrease in Ms is observed for the ferrite nucleation rate, while the ferrite growth rate
two steels containing Mo. On the other hand, a slight seems to be less affected. This suggestion leads to a
shift of the PF formation region to lower cooling rates is significantly finer PF grain size obtained in the present
observed when Mo is added, due to the increase in study after transformation from the deformed austenite,
hardenability. Cizek et al.[33] also observe a relatively compared to the case of the recrystallized austenite.[33]
weak contribution of molybdenum to the hardenability In contrast to what occurred with ‘‘reconstructive’’
enhancement. The authors suggest that the static recrys- type transformations, such as allotriomorph ferrite, for
tallization occurring after deformation at a high tem- displacive transformations such as bainite and M,
perature of 1423 K (1150 °C) for the tests reported in plastic deformation retards the decomposition of the
the present article and accompanied by the movement of austenite; this effect is known as ‘‘mechanical stabiliza-
the austenite grain boundaries might prevent molybde- tion.’’[39] The mechanism appears to be that the growth
num from segregating to these boundaries. In addition of bainite/M laths is retarded by the deformation in the
to this effect, the steels analyzed in the current study austenite. The transformation of deformed austenite
have bigger austenite grain sizes and higher levels of might be accelerated at initial stages, although the
niobium. Therefore, the total grain boundary area per overall rate of transformation would be reduced com-
unit volume, Sv, is reduced and the niobium in solution pared to nondeformed austenite. As a consequence, a
may have reached an important hardenability effect, smaller quantity of bainite may be formed in the former
saturating any further effect attributable to molybde- case, although this behavior seems to depend on the
num. degree of deformation. Lightly deformed austenite
Nevertheless, an important effect of Mo addition on transforms more rapidly relative to undeformed austen-
the ferritic structure size distribution is observed for the ite because of the increase in the defect density (nucle-
low cooling rates. Figure 9 shows three image quality ation sites, higher Sv). However, although the nucleation
maps obtained after the application of schedule B for rate will be larger in heavily deformed austenite, the
the three steels studied and the same cooling rate of overall rate of transformation will be reduced; the
0.1 K/s. PF is the predominant phase in the three growth will be retarded.
microstructures with a banded secondary phase. This Similarly to what is observed in the present study,
secondary phase is lamellar P for the 6NbMo0 steel and Kazimierz and Lis recently observed that bainitic trans-
degenerated P for the steels containing Mo. A decrease formation was also accelerated in deformed austenite in
in the banding is also noticed as Mo content increases. the range of high cooling rates.[40] These authors sug-
As shown in Figure 9(d), a considerable coarsening is gested that a higher dislocation density in the deformed
measured as Mo content increases. The mean ferritic austenite leads to a faster nucleation of the bainite laths.
grain sizes are 16.4, 16.8, and 19.1 lm for 6NbMo0, In contrast, at slower cooling rates, they observed a
6NbMo16, and 6NbMo31 steels, respectively. Addition- retardation of the bainitic transformation compared to
ally, the homogeneity of the final structure is clearly the kinetics found in nondeformed austenite.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 42A, DECEMBER 2011—3737


Fig. 9—Microstructural details obtained by EBSD for a cooling rate of 0.1 K/s and cycle B: image quality maps for (a) steel 6NbMo0, (b) steel
6NbMo16, and (c) steel 6NbMo31. (d) Unit size distributions (defined using the 15 deg misorientation criterion) plotted in terms of accumulated
area fraction for the three steels.

and both schedules in steel 6NbMo16. Different imaging


options were used to define parameters that can be
related to the microstructure, among others, the 4 and
15 deg misorientation crystallographic unit sizes. The
first one is related to the strength level of the resulting
microstructure, while the second defines toughness
properties related to microstructural homogeneity.
Figure 11 shows the influence of cooling rates on the
mean unit size for the microstructures transformed from
recrystallized (schedule A) and deformed (schedule B)
austenite in steel 6NbMo16, considering both the afore-
mentioned misorientation criteria of 4 and 15 deg. As
expected, using a 15 deg threshold misorientation, larger
unit sizes are quantified at all conditions. On the other
hand, it is found that, as a general trend, the average
unit size decreases with increasing the cooling rate. For
mainly ferritic structures, this fact is related to an
Fig. 10—Comparison between the CCT diagrams obtained for re-
crystallized and deformed austenite in steel 6NbMo31. increase in the nucleation rate and a decrease in grain
growth with increasing the cooling rate.[41–43] If the
4 deg criterion is chosen, the average unit size decreases
In order to better characterize the effect of stored from 16.9 to 8.5 lm in the case of recrystallized
deformation in austenite on transformed structures, austenite and from 13.8 to 8.1 lm in the case of
EBSD scans were performed for different cooling rates unrecrystallized austenite. Meanwhile, in the case of

3738—VOLUME 42A, DECEMBER 2011 METALLURGICAL AND MATERIALS TRANSACTIONS A


Fig. 11—Influence of cooling rate on the average unit size for both
recrystallized (schedule A) and unrecrystallized (schedule B) austenite
using 4 and 15 deg threshold misorientation criteria in steel
6NbMo16.

the 15 deg criterion, mean unit sizes decrease from 25.4


to 11.8 lm and from 17.6 to 12 lm for recrystallized
and deformed austenite, respectively.
In contrast to mainly ferritic structures, a different
behavior is observed when the microstructures show an
important fraction of BF. An increase in the mean unit
size value is observed for the 15 deg criterion, and the
increase is more important in the case of schedule A (5
and 10 K/s cooling rates). For schedule B, this increase Fig. 12—Grain size (15 deg) distributions plotted in terms of accu-
mulated area fraction for steel 6NbMo16 and cycles (a) A and (b) B.
shifts to higher cooling rates (20 K/s). The further
reduction in mean unit size at higher cooling rates is
related to the formation of M. This variation is not For schedule B (Figure 12(b)), a gradual increase in
observed for the 4 deg criterion, due to the fact that only heterogeneity is measured as the cooling rate increases.
high-angle grain boundary density is affected when the Relatively fine-grained and homogeneous distributions
bainite fraction increases; the low-angle boundary den- are obtained for mainly ferritic structures (0.1 to 1 K/s
sity remains approximately constant. range). When GF becomes the main phase, an increase
The results shown in Figure 9 define the evolution of in heterogeneity is clearly noticed, reaching a maximum
the mean values of several microstructural features, but level of heterogeneity for the 20 and 50 K/s cooling rate,
they do not provide information concerning the influ- exactly where BF appears. For the highest cooling rate
ence of accumulated strain and cooling rate on micro- measured, 100 K/s, a slight increase in homogeneity is
structural heterogeneity that could affect the toughness observed as the M fraction is increased.
behavior. In order to take this into account, the 15 deg Previously published studies suggest that, to quantify
unit size distributions obtained at different cooling rates the ductile-brittle transition, instead of the mean unit
are represented in Figure 12. Grain area fractions were size, a parameter able to properly catch the relevance of
considered in order to better identify the presence of coarse grain fractions is required.[44,45] Following that,
coarse grains. the unit size for which 10 pct of the area fraction of
In the case of schedule A (Figure 12(a)), different grains have a size greater than that value was selected
trends may be highlighted for the 15 deg crystallo- (denoted as Dc10 pct).
graphic grain distributions. For the lowest cooling rates The corresponding Dc10 pct values have been drawn
of 0.1 and 0.5 K/s, the mainly ferritic structures lead to in Figure 13 as a function of the cooling rate. A similar
homogeneous distributions. As the cooling rate is trend in Dc10 pct is observed for microstructures
increased to the 1 to 10 K/s range, an increase in transformed from recrystallized and deformed austenite.
heterogeneity is observed and the cooling rate of 10 K/s Dc10 pct values increase as the cooling rate increases up
is the one with the widest unit size distribution. to a maximum located at 10 K/s for cycle A and 50 K/s
Nevertheless, for the 20 and 50 K/s cooling rates, tied for cycle B. For higher cooling rates, the Dc10 pct value
to M formation, an improvement in homogeneity is decreases sharply in both cases.
reached obtaining the finest and most homogeneous At the lowest cooling rates, with mainly ferritic
distribution for the 50 K/s cooling rate. structures, the accumulation of strain provides a higher

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 42A, DECEMBER 2011—3739


strain was not sufficient to promote additional intra-
granular bainite nucleation on austenite defects.[51]
For a better understanding of the transition in
homogeneity between mainly bainitic structures and
the formation of some M, additional EBSD scans were
performed using the FEGSEM with a 0.2-lm step and
for a 140 9 140 lm2. Figure 14 shows two scans relative
to the 6NbMo16 steel and two cooling rates of 20 and
100 K/s transformed from deformed austenite (cycle B).
The first one is composed by a mixture of GF and BF,
while in the latter, some M is formed together with the
previous phases. Misorientation angle maps (Fig-
ures 14(c) and (d)) are plotted considering 4 and
15 deg criteria using gray and black colors, respectively.
Even if image quality maps suggest that the structure
in Figure 14(a) is finer and more homogeneous when
compared to the one in Figure 14(b), misorientation
image maps show that the high-angle boundary density
Fig. 13—Change of Dc10 pct value as a function of the cooling rate in the first one is low (Figure 14(c)), and therefore the
in steel 6NbMo16 and both cycles A and B. The crystallographic toughness properties of this microstructure would be
grains with misorientations of 15 deg were considered.
affected by the coarse units formed. Analysis of the
100 K/s phases (Figures 14(b) and (d)) shows that BF
density of nucleation sites resulting in a refined and packets are formed in parent austenite grain bound-
more homogeneous microstructure compared to the aries. Areas with phases such as BF and GF (left side
nondeformed material, following a tendency similar to of the image) show a low high-angle grain boundary
that observed for the mean unit size (Figure 11). At density, while regions with M formation (right side of
these cooling conditions, nucleation density is high the image) reflect an important refinement in the
enough to achieve a refinement in the entire unit size 15 deg unit sizes.
distribution and, therefore, results in low Dc10 pct
values.
However, in the case of higher cooling rates (>1 K/s), V. CONCLUSIONS
where the microstructure is constituted mainly by QF
and GF, the increase in the cooling rate seems not to The transformed microstructures for the three Nb-Mo
have an important effect on the refinement of the tail of microalloyed steels contain PF, QF, and P, bainite with
the unit size distribution given by high-angle grain different morphologies such as GF or BF, and M. The
boundaries. The increase in heterogeneity, reflected in accumulation of strain in the austenite accelerates both
the peak in the Dc10 pct, is tied to the maximum the ferritic and bainitic transformations, displacing the
fraction of BF in the microstructure. For the case of CCT diagrams to higher temperatures and shorter
high-temperature bainite transformations, Furuhara times.
et al.[46,47] observed that a small undercooling promotes A slight effect of molybdenum is observed in the
variant selection leading to coarse bainite packet sizes, ferrite stability region shifting it to longer times when
and Lambert-Perlade et al.,[48] also in conditions corre- the alloying level is increased. Additionally, microstruc-
sponding to initial coarse austenite grains, confirmed tural coarsening was quantified for low cooling rates
that the austenite grain size can play an important role (0.1 K/s) when Mo content increases.
in the nucleation of bainite variants. The formation of Mean unit size, measured by means of 4 or 15 deg
M at high cooling rates promotes an important micro- misorientation criteria using EBSD scans, decreases as
structural homogenization reflected in the Dc10 pct the cooling rate increases; the refining effect is more
value drop and general refinement (distributions in notorious in the range of low cooling rates where the
Figure 12), for structures transformed from both re- microstructures are mainly formed by ferrite. Similarly,
crystallized and unrecrystallized austenite. This refine- the accumulation of strain in the austenite results in
ment, related to M formation and bainite packet size smaller mean unit sizes. Nevertheless, this general trend
reduction, is also observed for normal austenite grain is not valid for cooling rates where BF fraction becomes
sizes and high undercooling.[49,50] important, and an increase in the 15 deg mean unit size
The influence of deformed austenite in microstruc- is observed until M is formed at higher cooling rates.
tural heterogeneities at high cooling rates must be taken Structure homogeneity varies depending on the cool-
into account. The beneficial effect observed in ferritic ing rate and the phases formed. In the range of low
microstructures, and also for QF and GF combinations, cooling rates, the microstructure is relatively finely
completely disappears when BF becomes the main grained and homogeneous. As the cooling rate increases
phase. In contrast to what happens with ferrite, results and the GF fraction becomes the major constituent, a
suggest that, for coarse austenite grains, bainite packet coarsening in the structure together with a more
refinement is not enhanced by the accumulated strain. In heterogeneous distribution is measured, reflected in the
the present case, probably the limited accumulated 15 deg mean unit size, unit distribution, and Dc10 pct

3740—VOLUME 42A, DECEMBER 2011 METALLURGICAL AND MATERIALS TRANSACTIONS A


Fig. 14—Microstructural details obtained in steel 6NbMo16 and cycle B using FEGSEM EBSD. Image quality maps: (a) 20 K/s, (b) 100 K/s
and misorientation angle maps (4 deg units: gray; 15 deg units: black), (c) 20 K/s, and (d) 100 K/s.

parameter. This increase reaches a peak for the maxi- REFERENCES


mum BF amount in both structures transformed from 1. G.I. Garcia: Int. Conf. Microalloying ‘95, ISS, Warrendale, PA,
recrystallized and deformed austenite. This suggests that 1995, pp. 365–75.
the coarse austenite grains transform into closely aligned 2. S.G. Jansto: New Developments on Metallurgy and Applications of
low misorientation BF units. These results suggest that High Strength Steels Conf., Buenos Aires, 2008, TMS, Warren-
dale, PA, pp. 1313–26.
the austenite grain size may play an important role in 3. N.A. McPherson: Ironmaking and Steelmaking, 2009, vol. 36,
the nucleation of bainite variants. The formation of M pp. 193–200.
promotes a refined and more homogeneous structure; 4. E.J. Czyryca, D.P. Kihl, and R. DeNale: AMPTIAC Q., 2003, vol.
this homogenization is more important as the M 7, pp. 63–70.
5. B. Dutta, E. Valdés, and C.M. Sellars: Acta Metall. Mater., 1992,
fraction increases. vol. 40, pp. 652–62.
6. M.G. Akben, I. Weiss, and J.J. Jonas: Acta Metall., 1981, vol. 29,
pp. 111–21.
7. O. Kwon and A.J. DeArdo: Acta Metall. Mater., 1991, vol. 39,
ACKNOWLEDGMENTS pp. 529–38.
8. D.N. Hanlon, J. Sietsma, and S. van der Zwaag: ISIJ Int, 2001,
Financial support of this work by the Spanish Sci- vol. 41, pp. 1028–36.
ence and Innovation Department (MAT2009-09250 9. Y. van Leeuwen and J. Sietsma: Mater. Sci. Forum, 2007,
project) is gratefully acknowledged. One of the authors vols. 539–543, pp. 4572–77.
10. S. Cai and J.D. Boyd: Mater. Sci. Forum, 2005, vols. 500–501,
(NI) acknowledges a research grant from the Univer- pp. 171–78.
sity of Navarra. PU is grateful to NSF and TMS for 11. H. Asahi, A. Yagi, and M. Ueno: Metall. Mater. Trans. A, 1998,
the MS&T’10 Conference registration fee funding. vol. 29A, pp. 1375–81.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 42A, DECEMBER 2011—3741


12. T. Tanaka: Int. Met. Rev., 1981, vol. 26, pp. 185–212. 32. R. Petrov, L. Kestens, and Y. Houbaert: Mater. Charact., 2004,
13. H. Meuser, F. Grimpe, S. Meimeth, C.J. Heckmann, and C. vol. 53, pp. 51–61.
Träger: Mater. Sci. Forum, 2005, vols. 500–501, pp. 565–72. 33. P. Cizek, B.P. Wynne, C.H.J. Davies, B.C. Muddle, and P.D.
14. D. Chakrabarti, M. Strangwood, and C. Davis: Metall. Mater. Hodgson: Metall. Mater. Trans. A, 2002, vol. 33A, pp. 1331–49.
Trans. A, 2009, vol. 40A, pp. 780–95. 34. P.A. Manohar, T. Chandra, and C.R. Killmore: ISIJ Int., 2006,
15. P. Uranga, A.I. Fernández, B. López, and J.M. Rodriguez-Ibabe: vol. 36, pp. 1486–93.
43rd Mechanical Working and Steel Processing Conf., ISS, War- 35. I. Tamura: Int. Conf. Thermec ‘88, ISIJ, Tokyo, 1988, pp. 1–10.
rendale, PA, 2001, vol. 33, pp. 511–29. 36. T. Tanaka: Int. Conf. Microalloying 95, M. Korchynsky, A.J.
16. B.L. Bramfitt and J.G. Speer: Metall. Trans. A, 1990, vol. 21A, DeArdo, P. Repas, and G. Tither, eds., Pittsburgh, ISS, Warren-
pp. 817–29. dale, PA, 1995, pp. 165–81.
17. T. Araki, I. Kozasu, H. Tankechi, K. Shibata, M. Enomoto, and 37. R. Bengochea, B. Lopez, and I. Gutierrez: Metall. Mater. Trans.
H. Tamehiro, eds., Atlas for Bainitic Microstructures, ISIJ, Tokyo, A, 1998, vol. 29A, pp. 417–26.
1992, vol. 1. 38. R. Bengochea, B. Lopez, and I. Gutierrez: in Microalloying
18. H.I. Aaronson and H.A. Domian: Trans. AIME, 1966, vol. 236, in Steels (l-as 98), J.M. Rodriguez-Ibabe, I. Gutierrez, and B.
pp. 781–96. Lopez, eds., San Sebastian, Spain, 1998, pp. 201–08.
19. M. Hillert: in Solid-Solid Phase Transformations, H.I. Aaronson, 39. H.K.D.H. Bhadeshia: Bainite in Steels, Transformations, Micro-
D.E. Laughlin, R.F. Sekerka, and C.M. Wayman, eds., TMS, structure and Properties, 2nd ed., The Institute of Materials,
Warrendale, PA, 1982, pp. 789–806. London, 2001, pp. 201–24.
20. D.E. Coates: Metall. Trans., 1973, vol. 4, pp. 2313–25. 40. A. Kazimierz and J. Lis: Mater. Sci. Forum, 2007, vols. 539–543,
21. T.B. Massalski: Phase Transformations, ASM, Metals Park, OH, pp. 4620–25.
1970, pp. 433–95. 41. M. Umemoto, Z.H. Guo, and I. Tamura: Mater. Sci. Technol.,
22. M. Hillert: Metall. Trans. A, 1984, vol. 15A, pp. 411–19. 1987, vol. 3, pp. 249–55.
23. J. Cawley, C.F. Harris, and E.A. Wilson: New Aspects of Micro- 42. S. Zajac, T. Siwecki, B. Hutchinson, and M. Attlegard: Metall.
structures in Modern Low Carbon High Strength Steels Symp, ISIJ, Trans. A, 1991, vol. 22A, pp. 2681–94.
Tokyo, 1994, pp. 11–14. 43. J.H. Beynon and C.M. Sellars: High Strength Low Alloy Steels
24. K. Shibata and K. Asakura: New Aspects of Microstructures in Conf., Wollongong, 1984, TMS, Warrendale, PA, 1984, pp. 142–
Modern Low Carbon High Strength Steels Symp, ISIJ, Tokyo, 50.
1994, pp. 31–34. 44. A. From and R. Sandström: Mater. Charact., 1999, vol. 42,
25. G. Krauss and S.W. Thompson: ISIJ Int., 1995, vol. 35, pp. 937– pp. 111–22.
45. 45. T. Hanamura, F. Yin, and K. Nagai: ISIJ Int., 2004, vol. 44,
26. H.K.D.H. Bhadeshia: Bainite in Steels, Transformations, Micro- pp. 610–17.
structure and Properties, 2nd ed., The Institute of Materials, 46. T. Furuhara, H. Kawata, S. Morito, and T. Maki: Mater. Sci.
London, 2001, pp. 277–79. Eng. A, 2006, vol. A431, pp. 228–36.
27. H.J. Lee, G. Spanos, G.J. Shiflet, and H.I. Aaronson: Acta Me- 47. T. Furuhara, N. Takayama, and G. Miyamoto: Mater. Sci.
tall., 1988, vol. 36, pp. 1129–40. Forum, 2010, vols. 638–642, pp. 3044–49.
28. S. Zajac, V. Schwinn, and K.H. Tacke: Mater. Sci. Forum, 2005, 48. A. Lambert-Perlade, A.F. Gourgues, and A. Pineau: Acta Mater.,
vols. 500–501, pp. 387–94. 2004, vol. 52, pp. 2337–48.
29. R.F. Speyer: Thermal Analysis of Material, Marcel Dekker, Inc, 49. K. Fujiwara, S. Okaguchi, and H. Ohtani: ISIJ Int., 1995, vol. 35,
New York, NY, 1994. pp. 1006–12.
30. C. de Garcı́a Andrés, F.B. Caballero, C. Capdevila, and H.K.D.H. 50. K. Fujiwara and S. Okaguchi: Mater. Sci. Forum, 1998, vols. 284–
Bhadeshia: Scripta Mater, 1998, vol. 39, pp. 791–96. 286, pp. 271–78.
31. ABAQUS Reference Manuals, Dassault Systèmes, Providence, RI, 51. R.Y. Zhang and J.D. Boyd: Metall. Mater. Trans. A, 2010,
2009. vol. 41A, pp. 1448–59.

3742—VOLUME 42A, DECEMBER 2011 METALLURGICAL AND MATERIALS TRANSACTIONS A

You might also like