Sie sind auf Seite 1von 11

Materials Science & Engineering A 564 (2013) 389–399

Contents lists available at SciVerse ScienceDirect

Materials Science & Engineering A


journal homepage: www.elsevier.com/locate/msea

Cold-rolling and inter-critical annealing of low-carbon steel: Effect of initial


microstructure and heating-rate
A. Karmakar a, M. Ghosh b, D. Chakrabarti a,n
a
The Department of Metallurgical and Materials Engineering, Indian Institute of Technology (I.I.T.), Kharagpur 721302, Kharagpur, West Bengal, India
b
Council of Scientific and Industrial Research, National Metallurgical Laboratory, CSIR-NML, Jamshedpur 831007, Jharkhand, India

a r t i c l e i n f o abstract

Article history: The effects of starting microstructure and heating rate on the inter-critical annealing treatment have
Received 23 August 2012 been investigated by 80% cold-rolling of different initial microstructures namely ferrite–pearlite and
Received in revised form ferrite–martensite with blocky and fibrous martensitic morphologies, followed by inter-critical
27 November 2012
annealing treatment using two different heating rates (  0.5 1C/s and  300 1C/s). Sub-critical annealing
Accepted 28 November 2012
has also been carried out to understand the effect of starting structures on cold-rolled and completely
Available online 3 December 2012
recrystallized microstructures. Due to the fine-scale lamellar structure comprised of alternate layers of
Keywords: ferrite and martensite, ferrite-fibrous martensite starting structure showed the finest ferrite grain sizes
Cold-rolling (3–6 mm) and more uniform distribution of martenisitc islands compared to ferrite–pearlite and ferrite-
Sub-critical annealing
blocky martensite after the inter-critical annealing. Slow heating resulted in a coarser ferrite grain size
Inter-critical annealing
with more uniform distribution of martensite, compared to rapid heating. Recrystallization–
Heating rate
Starting microstructures transformation interaction and the avoidance of ferrite grain growth contributed to the finer grain
Low-carbon steel sizes after rapid annealing. Nature and distribution of y-particles and the austenite islands played an
important role in controlling the ferrite recrystallization and grain growth. Finer microstructural
constituents offered superior combination of strength, ductility and strain-hardening ability to ferrite-
fibrous martensite starting structures after rapid annealing, compared to other structures. Rapidly
heated samples showed higher strength than the slowly heated samples.
& 2012 Elsevier B.V. All rights reserved.

1. Introduction annealing treatment, in place of conventional ferrite–pearlite


structure to develop ultrafine ferrite grains [6–14]. Tsuji and co-
Conventional, sub-critical annealing of the cold-rolled steel is workers [6–8] produced ultra-fine ferrite (UFF) grained steel with
either carried out in batch-annealing furnace or in continuous- grain size less than 1 mm by 50% cold-rolling of martensite and
annealing furnace. This annealing schedule comprised of slow annealing at 500 1C. Grange [9] performed short and rapid
heating (heating rate o10 1C/s) to the annealing temperature austenitizing of undeformed martensite, which has been recently
below Ac1 (  550–700 1C), prolonged isothermal holding ( 41 h) repeated by Azizi-Alizamini et al. [10] and Nakada et al. [11] on
followed by slow-cooling [1]. In general sub-critical (recrystalli- both undeformed and cold-rolled martensite. However, the
zation) annealing offers yield strength (YS)  200–300 MPa [1], deformation of martensite is difficult as it requires high-flow
whilst, inter-critical annealing above Ac1 and water-cooling (cool- stress [8] and suffers from transverse cracking problem at the
ing rate, RC Z50 1C/s) can raise YS to 600 MPa in cold-rolled, edge of the specimen [12]. Considering the problems associated
dual-phase steels [2–4]. Development of ultra-fine grained steels with cold-deformation of martensite, rolling of mixed phase
(average grain size o3 mm) by applying severe plastic deforma- structures can be preferred for the development of ultra-fine
tion (SPD) processes can increase the strength even higher ferrite grain structures [8–15].
(YS 4800 MPa), but affect the strain-hardening ability and uni- Ultra-fine dual-phase (UFDP) steels show better strength and
form elongation (eu r 5%) [5]. SPD techniques are also difficult to uniform elongation than the UFF, carbide steels [5,10,16], which
apply in the industrial scale. encouraged the development of rapid-transformation annealing
Considering the above limitations, different starting micro- (RTA) or ultra-rapid annealing (URA) treatment [17–20]. RTA is
structures have been subjected to cold-rolling and sub-critical characterized by rapid-heating (RH  200–300 1C/s) of cold-rolled
steels above Ac1, followed by short isothermal holding ( 5–60 s)
and rapid cooling (RC Z100 1C/s) [15–19]. Higher heating rate in
n
Corresponding author. Tel.: þ91 3222 283282; fax: þ 91 3222 282280. RTA, compared to the conventional inter-critical annealing
E-mail address: debalay@metal.iitkgp.ernet.in (D. Chakrabarti). (RH 20–50 1C/s) prevents the growth of recrystallized ferrite

0921-5093/$ - see front matter & 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.msea.2012.11.109
390 A. Karmakar et al. / Materials Science & Engineering A 564 (2013) 389–399

grains and helps in obtaining fine-grain size [17–20]. Heating-rate


also plays an important role in inter-critical annealing as it can
vary over a wide range starting from slow-heating (say, 0.5 1C/s)
to RTA ( 300 1C/s). Compared to the studies on sub-critical
annealing lesser attention has been paid to understand the effect
of initial microstructures on inter-critical annealing.
Therefore, in the present study ferrite–pearlite and ferrite–
martensite structures have been cold-rolled and inter-critically
annealed after slow- and rapid-heating

2. Experimental details

Samples from a 6 mm thick hot-rolled strip containing 0.1 C,


0.33 Si, 1.42 Mn, 0.01 P, 0.003 S, 0.035 Al, 0.05 Nb, 0.05 V and
0.007 N (all wt%) has been reheated to 1150 1C, soaked for 30 min
and subjected to three different heat-treatments, namely furnace-
cooling (FC), step-quenching (SQ) and intermediate-quenching
(IQ), as mentioned below:

(i) Furnace-cooling (FC): Slow rate of cooling ( o1 1C/s) from the


reheating temperature down to the ambient temperature,
(ii) Step-quenching (SQ): Rapidly cooling the samples from the
reheating temperature to an inter-critical annealing tem-
perature of 750 1C and holding for 1 h before ice-water
quenching,
(iii) Intermediate quenching (IQ): Ice-water quenching of the
samples from the reheating temperature followed by inter-
critical annealing at 750 1C for 1 h and ice-water quenching
once again.

Heat-treated samples were cold-rolled to 80% reduction and


were annealed either inside a programmable muffle furnace at
600 1C for 1–10 h (sub-critical annealing) or immersed inside a
salt-bath at 850 1C for 5–60 s (inter-critical annealing by RTA
treatment), before water-quenching. Selection of annealing tem-
perature was based on the dilatometric study carried out using
BAHRs-Thermoanalyse (DIL 805 A/D model) dilatometer. K-type
thermocouple, which continuously monitored the sample tem-
perature during annealing, showed a heating rate (RH) of 30 1C/s
for sub-critical treatment and  300 1C/s for RTA treatment.
Besides RTA, inter-critical annealing has also been carried out
after slow-rate of heating (RH  0.5 1C/s). Cross-section of the
annealed strips were prepared following standard metallographic
techniques and investigated by optical microscope, scanning
electron microscope (SEM) and image analysis. Determination of
average ferrite grain size was based on the measurement of at
least 500 grains in terms of equivalent circle diameter (ECD) size.
Size of at least 200 martensitic islands and pearlite colonies were
also measured from each sample in terms of ECD and the average
values are determined. Sub-size tensile specimens have been
prepared
pffiffiffiffiffiffi following ASTM E-8 M standard (maintaining
L0 = A0 ¼ 4:5, where L0 and A0 are the gage length and cross-
sectional area of the samples, respectively) and tested in an
Instron-8862 Universal Testing Machine (10 t capacity) at room
temperature (25 1C) using a cross-head velocity of 1 mm/min.

3. Results and discussion Fig. 1. (a) Ferrite–pearlite, (b) ferrite-blocky martensite and (c) ferrite-fibrous
martensite microstructures after 80% cold-rolling.
3.1. Microstructures of cold-rolled samples
initial microstructures are reported earlier [21]. Microstructures
FC, SQ and IQ heat-treatments on the as-received strip resulted of the cold-rolled samples are presented in Fig. 1. Besides bending
in three different starting-microstructures namely, ferrite– and thinning of cementite (y) lamellae, dislocation pile-up against
pearlite, ferrite-blocky martensite and ferrite-fibrous martensite, the lamellae in ferrite–pearlite structure can disintegrate them
respectively. Detailed discussion and characterization of the upon cold-rolling, Fig. 1(a) [13]. Cold-rolling of ferrite-blocky
A. Karmakar et al. / Materials Science & Engineering A 564 (2013) 389–399 391

martensite structure leads to the heterogeneous plastic-flow with uniform strain-distribution and finer y-particle size (i.e. stronger
higher strain-accumulation inside the softer ferrite phase, com- grain boundary pinning). Finer y-particle size in ferrite-fibrous
pared to harder martensite phase, Fig. 1(b). On the other hand, martensite structure can be related to higher stability of y-particles
more homogeneous plastic deformation is expected in ferrite- in that sample, with respect to the other starting structures. This
fibrous martensite structure, due to the fibrous morphology with aspect has been studied earlier and attributed to the partitioning of
alternate layers of ferrite and martensite phases, Fig. 1(c). Mn from ferrite to austenite during the inter-critical annealing of
martensite, as a part of intermediate quenching (IQ) heat-
treatment [21]. High Mn level of the austenite was inherited by
3.2. Microstructures of the sub-critically annealed samples
martensite after water-quenching and ultimately lead to Mn
enrichment within the y-particles, which formed by the decom-
With respect to the starting microstructures, the average ferrite
position of martensite by virtue of cold-rolling and annealing
grain size after cold-rolling and sub-critical annealing followed the
treatment [21]. Mn enrichment inside the y-particles converts
sequence: ferrite–pearlite (14–17 mm)4ferrite-blocky martensite
cementite (Fe3C) to iron-manganese carbide, which is also known
(7–9 mm)4ferrite-fibrous martensite (5–6.3 mm). Average y-parti-
as Cohenite (Fe3 xMx)C [16]. This slowed down the coarsening-rate
cle size measured after annealing also followed the same trend:
of those particles at the time of annealing due to the involvement
ferrite–pearlite (0.35–0.70 mm)4ferrite-blocky martensite
of Mn diffusion, which is a much slower process than the
(0.20–0.30 mm)4ferrite-fibrous martensite (0.16–0.25 mm). Lower
diffusion of C.
strain-accumulation in ferrite and the presence of coarser
y-particle size can explain the largest ferrite grain size obtained
3.3. Microstructures of the inter-critically annealed samples
after the sub-critical annealing of cold-rolled, ferrite–pearlite
structure, Fig. 2(a). Large y-particles are ineffective in pinning
3.3.1. Rapid transformation annealing (RTA)
down the propagating ferrite grain boundaries and hence, retard-
In case of cold-rolled ferrite–pearlite structure, increase in RH
ing the grain growth during the annealing. Finer ferrite grain sizes
from 30 1C/s to 300 1C/s resulted in 30–401C rise in Ac1 from
in ferrite-fibrous martensite structure, Fig. 2(b), compared to the
 750 1C to  780 1C and Ac3 from  905 1C to 930 1C. Those
other starting structures could have resulted from the more
values were much higher than the equilibrium transformation
temperatures (Ae1 680 1C and Ae3  860 1C), predicted from the
Thermo-Calcs software, which is expected [8,20,22]. Annealing
temperature (850 1C) selected for this study, therefore, lies in the
inter-critical temperature range. Microstructures obtained after
rapid-transformation annealing (RTA) are shown in Fig. 3. Due to
small thickness of the samples ( 1 mm) microstructure did not
vary through the thickness. Microstructure after RTA treatment
was comprised of untransformed ferrite grains and a mixture of
fine-ferrite, bainite and martensite, which were the transforma-
tion product of austenite after water-quenching. With the
increase in annealing time from 5 s to 60 s, fraction of trans-
formed phases, fT (i.e. bainite and martensite) increased continu-
ously and the fraction of recrystallized ferrite, fR, also followed the
similar trend, Table 1.
Presence of partially recrystallized ferrite after 5 s annealing
along with the harder (dark) constituents, Fig. 4(a), indicated to
the interaction between ferrite-recrystallization and austenite
formation, which is a characteristic feature of the RTA treatment
[16–20]. Recrystallized ferrite fraction (fR) has been obtained in
the present study following three different techniques:
(i) hardness measurement, (ii) electron backscatter diffraction
(EBSD) analysis and (ii) microstructural observation. Vickers
micro-hardness readings (20 g load) were taken entirely from
the ferrite regions far from the harder constituents. Assuming a
linear relationship between the hardness variation and the soft-
ening due to recrystallization, fR has been calculated for different
annealing times, using the equation given in [4,22–24]. The
hardness measured in the cold-rolled samples (240–255 HV),
completely recrystallized samples (150–160 HV), and in the
samples annealed to any intermediate stage, have been used for
the calculation. EBSD analysis can also distinguish between the
strain-free recrystallized grains, surrounded by high-angle
boundaries (misorientation 4151), and the deformed matrix,
containing low-angle sub-boundaries (2–151 misorientation). In
the EBSD scan, recrystallized and deformed ferrite regions can be
separated by color codes as shown in Fig. 4(b). Thick- and thin-
lines in Fig. 4(b) represent the high-angle and low-angle bound-
aries, respectively. The recrystallized fraction measured by hard-
ness testing, EBSD study and microstructural observation
Fig. 2. Microstructures of: (a) ferrite–pearlite and (b) ferrite-fibrous martensite matched closely with each other (within75% accuracy) and
starting structures after cold-rolling and sub-critical annealing at 600 1C for 6 h. the average of those measurements are reported in Table 1.
392 A. Karmakar et al. / Materials Science & Engineering A 564 (2013) 389–399

Fig. 3. Microstructures of: (a) ferrite–pearlite, (b) ferrite-blocky martensite, and (c,d) ferrite-fibrous martensite after cold-rolling and RTA at 850 1C for (a–c) 5 s and
(d) 30 s. Coarse- and fine-grain regions in the bimodal grain structure in (d) and corresponding ferrite grain size distribution (area frequency vs range of grains) in (e) are
indicated by arrows.

Table 1
Comparison between the different starting microstructures in terms of average ferrite grain size (Da) and the tensile properties after 80% cold-rolling and rapid-
transformation annealing (RTA) treatment at 850 1C for different durations (5–60 s). Also given are the transformed austenite fractions, fT, 5 s and fT, 60 s, after 5 s and 60 s
of annealing times, respectively, and recrystallized ferrite fractions, fR, 5 s and fR, 60 s, after the same annelaing times.

Da (mm) fT, 5 s (%) fT, 60 s (%) fR, 5 s (%) fR, 60 s (%) YS (MPa) UTS (MPa) eu (%) et (%)

Ferrite–pearlite 9.0–11.5 37 65 65 100 654–826 835–911 9.7–10.9 12.6–15.0


Ferrite-blocky martensite 6.0–10.0 28 78 90 100 850–1080 1040–1269 8.0–10.6 15.7–19.7
Ferrite-fibrous martensite 3.6–5.8 28 65 75 100 941–1210 1125–1293 8.3–15.0 12.0–24.0

nn
YS: Yield strength, UTS: Ultimate tensile strength, eu: Uniform elongation and et: total elongation.

Ferrite recrystallization kinetics, austenite formation kinetics and As both ferrite-recrystallization and austenite formation are
the interaction between both the phenomena depend on several the diffusion controlled processes and high heating rate reduces
factors, as discussed below. the time for the carbon diffusion, rapid heating can slow-down
A. Karmakar et al. / Materials Science & Engineering A 564 (2013) 389–399 393

the non-isothermal recrystallization and transformation kinetics In case of ferrite–pearlite and ferrite-blocky martensite starting
compared to the slow-heating rate [20,22–25]. Presence of structures, due to the higher heating rate and small dwell time
deformed ferrite grains with internal strain (i.e. strain energy), used in RTA, y-particles did not get time to distribute uniformly
on the other hand, may accelerate the ferrite recrystallization throughout the microstructure, resulting in the formation of
process [12,22–24]. Hence, higher strain-accumulation in ferrite coarse-blocks of austenite ( 410 mm) in the matrix of recrystal-
might have resulted in highest recrystallized fraction in ferrite- lized ferrite grains by the short-range C-diffusion, covering prior-
blocky martensite starting structure, compared to the other pearlite or prior-martensite regions respectively, Figs. 3(a,b) and
starting structures, Table 1. Kinetics of austenite formation will 5(a–c) [4,22]. Ferrite–martensite structure developed after RTA
primarily depend on the available ferrite-carbide interface area, therefore, followed the same spatial distribution as the micro-
which is the nucleation site for austenite [17,18,22]. Rapid structural constituents in the cold-rolled sample (Fig. 5(a–c)),
annealing with short dwell time prevents the spheroidization of especially, at lower annealing time as expected from earlier
fragmented cementite lamellae in cold-rolled ferrite–pearlite studies [23–26]. Fibrous martensite morphology, on the other
structure, and unspheroidized carbides provide larger interfacial hand, resulted in more uniform distribution of y-particles and
area for austenite formation compared to the spheroidized hence, finer austenite islands after RTA, compared to the other
carbides [22]. This can explain the higher austenite fraction starting structures, Figs. 3(c,d) and 5(d–f).
observed in ferrite–pearlite structure after the initial stage (5 s) The austenite fraction formed even after 60 s annealing (65–
of annealing, Table 1. 78%) remained below the equilibrium austenite fraction (  90%).
Schematic diagram in Fig. 5 is used to represent the micro- This difference can arise as the final stage of ferrite-austenite
structural changes taking place during the inter-critical annealing transformation is controlled by the Mn diffusion, which requires
treatment for different starting structures and heating rates. prolonged isothermal holding [21]. In case of ferrite-fibrous
martensite starting structure, y-particles are enriched with Mn
compared to ferrite-blocky martensite. Average Mn level obtained
by EDS (energy dispersive spectroscopy) analysis from the y
particles of ferrite-fibrous martensite structure is  1.9770.2 wt%
whilst the same for ferrite-blocky martensite structure is  0.847
0.3 wt%. As Mn diffusion from y to austenite is a slow, austenitiza-
tion by the dissolution of y becomes sluggish. Presence of deformed
ferrite grains on the other hand, accelerates the austenitization
kinetics by strain-induced transformation and also by increasing
the diffusivity of C and Mn through the deformed matrix
[10,17,18,26,27].
Clear evidence of fine-microalloy precipitates in Fig. 6, such as,
Nb(C,N) and fine, y-particles, which formed during annealing of
deformed martensite, can retard the ferrite-recrystallization and
therefore, promote the recrystallization–transformation interac-
tion to take place [10,14,17,18,22]. The effect of y-particles on the
recrystalliztaion–transformation interaction is expected to be
most severe in ferrite-fibrous martensite starting structure, as
Mn enrichment within y increases the stability of those particles
and thus widens the three phase-field comprised of ferrite,
austenite and y [27]. Presence of un-dissolved y-particles, as
observed after 5 s annealing, retards the ferrite recrystallization.
The small austenite islands formed during the transformation can
also impose a strong pinning effect on the ferrite sub-grain
boundaries and grain boundaries slowing down the recrystalliza-
tion- and growth-rate of ferrite, respectively, Fig. 5(e,f), as
discussed by Ogawa et al. [14]. Therefore, microalloy precipitates,
y-particles, and uniform distribution of small austenite islands
help in generating fine grained dual-phase steels in ferrite-fibrous
martensite starting structure by retarding the ferrite recrystalli-
zation, increasing the recrystallization–transformation interac-
tion, and inhibiting the grain-growth of recrystallized ferrite
grains. In case of ferrite–pearlite or ferrite-blocky martensite
starting structures as the austenite forms preferentially at prior-
pearlitic or prior-martensitic regions lack of y-particles and
austenitic islands allow the recrystalliztaion and grain-growth
of ferrite, Fig. 5(a–c). During the inter-critical annealing of cold-
rolled steels less than 10% austenite has been reported to be
ineffective in retarding the ferrite recrystallization [14,28]. As a
result ferrite grain sizes in those samples become larger than that
Fig. 4. (a) Partially recrystallized ferrite matrix (bright region) with harder in the samples having ferrite-fibrous martensite starting structure
constituents, such as, bainite and martensite (dark region) in a RTA treated after the RTA treatment, Table 1.
sample (850 1C, 5 s) of ferrite–pearlite structure. Arrows indicating to deformed Dual phase steels with uniform distribution of less than 30%
ferrite grain, recrystallized ferrite grain and harder constituent; (b) EBSD map
separating the deformed ferrite grains (in yellow) and recrystallized regions (in
martensite in the ferrite matrix are known to provide excellent
blue). (For interpretation of the references to color in this figure legend, the reader combination of strength, ductility and formability [10,16,27,29–31].
is referred to the web version of this article.) Ferrite–pearlite and ferrite-blocky martensite starting structures are
394 A. Karmakar et al. / Materials Science & Engineering A 564 (2013) 389–399

Fig. 5. Schematics of microstructural evolution during inter-critical annealing by (a–f) RTA and (g,h) slow heating. Ferrite–pearlite starting structure: (a) cold-rolled
sample; (b) just before the start of austenite formation during RTA showing partially recrystallized ferrite matrix with carbide particles, (c) subsequent austenite formation
in large blocks. Ferrite-fibrous martensite starting structure: (d) cold-rolled sample, (e) just before the start of austenite formation during RTA showing uniform
distribution of carbide particles, (f) subsequent development of bimodal ferrite grain structure with fine austenite islands. Ferrite-blocky martensite starting structure:
(g) Slow heating of the cold-rolled sample (refer Fig. 5a) to just before the austenite formation during annealing showing uniform distribution of carbides and
(h) subsequent austenite formation around the carbides and along the ferrite grain boundaries.
A. Karmakar et al. / Materials Science & Engineering A 564 (2013) 389–399 395

those austenitic islands due to Mn enrichment, led to predomi-


nantly martensitic transformation upon water-quenching, with
less than 1% bainite. Besides Mn level C level inside the initial
martensite may also influence the precipitation of carbide parti-
cles and the dissolution of those particles into austenite. Similar
martensite fraction present in ferrite-blocky martensite and
ferrite-fibrous martensite starting structures [21] is expected to
generate similar C level in martensite and the effect of C can
therefore, be ignored.
Bimodal ferrite grain size has been observed in case of ferrite-
fibrous martensite starting structure after cold-rolling and RTA
with coarse-ferrite grains (6–10 mm) being present along with the
ultra-fine ferrite grains (1–2 mm), Fig. 3(d). According to recent
studies, grain size less than 3 mm are considered as ultrafine
grains [34–36]. Ferrite grain size distribution in Fig. 3(e) shows
two separate peaks (indicated by arrows) which clearly depicted
the bimodal nature of ferrite grains in Fig. 3(d). Ogawa et al. [14]
observed similar bimodality after rapid inter-critical annealing of
cold-rolled bainitic structure and attributed that to the interac-
tion between austenite formation and ferrite recrystallization.
Pinning of ferrite sub-boundaries by fine austenite islands
(  1 mm) in the un-recrystallized regions stabilize the sub-grain
structure, which gradually converts into ultra-fine ferrite grains
by extended recovery, Fig. 5(e,f). Ultra-fine grains along with
coarse, recrystallized ferrite grains constitute the bimodal grain
structure, Figs. 3(d) and 5(f).

3.3.2. Slow inter-critical annealing


Slow-rate of heating (RH  0.5 1C/s) of the cold-rolled samples
to the inter-critical temperature produced more uniform distri-
bution of ferrite and martensite, with respect to RTA, as austenite
formed not only around the carbide particles (which were uni-
formly distributed) but also along the ferrite grain boundaries,
Fig. 7(a,b). Ferrite-fibrous martensite starting structure showed
the finest ferrite grain size (  5.9 71.0 mm) even after the slow
Fig. 6. (a) Dark field TEM micrographs showing finely distributed precipitates and
corresponding (b) SAD (selected area diffraction) pattern of (a) confirmed the heating, compared to ferrite–pearlite ( 14.4 73.2 mm) and
presence of Nb (C,N). ferrite-blocky martensite (  12.4 72.7 mm) starting structures,
Fig. 7(a,b). Ferrite grain sizes of the cold-rolled and slowly heated
samples were similar to the sub-critically annealed samples,
not suitable in providing such uniform distribution of martensite
Fig. 2, which indicated that the deformed ferrite grains recrys-
after rapid annealing. Due to the occurrence of ferrite recrystalliza-
tallized fully before the austenite formation, Fig. 5(g). Stable grain
tion, migrating ferrite grain boundaries does not allow the austenite
boundaries surrounding the recrystallized ferrite grains allowed
nuclei to form at those locations and austenite nucleates only at the
the austenite to form on them, Fig. 5(h) [22–25]. Desired micro-
ferrite–cementite interface of prior-pearlitic (or martensitic) regions
structure can also be obtained by two-step annealing of a cold-
[22–24]. High heating-rate also prevents the growth of any grain
rolled structure comprised of sub-critical annealing followed by
boundary nucleated austenite by reducing the time available for
the RTA treatment. Sub-critical annealing below Ac1 for sufficient
carbon diffusion.
time (4 h at 600 1C) distributes the y-particles uniformly through-
In general, austenite formation under the rapid-heating is
out the microstructure so that small austenitic islands (1–4 mm)
expected to be governed by non-partitioning local equilibrium
can form around each y-particles during RTA at 850 1C for 30 s,
condition, which mainly requires the diffusion of C with highly
Fig. 7(c) [10,23]. However, the problem associated to slow-
localized diffusion of Mn and Si [32,33]. Segregation of Mn and Si
heating or two-step annealing treatment of cold-rolled structure
at the ferrite/austenite interface may however, increase the
is the increase in average-ferrite grain size to more than 10 mm,
hardenability of austenite [17,18,22]. On the other hand, fine
Fig. 7. Use of the ferrite-fibrous martensite starting structure may
austenite grain size can reduce the hardenability. Fast-heating
however, solve that problem and provide martensitic islands of
and short holding time could not allow the sufficient carbon
less than 2 mm in size distributed in the matrix of fine-ferrite
diffusion, which is necessary for the homogenization of coarse
grains (  6 mm) after slow inter-critical annealing or ultra-fine (or
austenite-blocks with respect to carbon [20–24]. Lack of homo-
bimodal) ferrite grains (  3 mm) after RTA.
genization resulted in the formation of high fraction of bainite
(3–13%) after the RTA treatment of cold-rolled ferrite–pearlite
and ferrite-blocky martensite starting structures. The distribution 3.4. Evaluation of the mechanical properties
of ferrite and martensite was far more uniform in ferrite-fibrous
martensite starting structure, with two consecutive martensitic Typical engineering stress–strain curves of the inter-critically
films only 3–4 mm apart and separated by ferrite film, Figs. 1(c) annealed samples are presented in Fig. 8. The variation in
and 5(d). After cold-rolling and subsequent annealing, such a mechanical properties, such as, strength and ductility of the RTA
structure produced fine, and more uniform distribution y-parti- samples are plotted in Fig. 9(a,b) with respect to the annealing
cles and hence, austenitic islands, Fig. 5(e,f). High hardenability of time(s) at 850 1C. Higher strength and continuous yielding of RTA
396 A. Karmakar et al. / Materials Science & Engineering A 564 (2013) 389–399

Fig. 8. Engineering stress–strain curves of the RTA samples (850 1C, 30 s) for
different starting microstructures. Also given is the tensile curve of the slowly
heated sample for the same annealing condition.

ferrite-fibrous martensite structures after RTA treatment can be


due to the finest ferrite grain size and uniform distribution of the
small martensitic islands, Figs. 8 and 9 and Table 1. In case of
ferrite–pearlite and ferrite-blocky martensite starting structures,
relatively coarser ferrite grains and coarse patches of harder
constituents with high bainite fraction, offered lower strength
compared to the ferrite-fibrous martensite starting structure.
Variation in mechanical properties of RTA samples as the
function of annealing parameters was controlled by a more
complex interaction between ferrite-recrystallization and auste-
nite formation. Initial drop in YS (and also UTS) with the increase
in the annealing time from 5 s to 15 s, Fig. 9(a), can be attributed
to the softening of the ferrite grains by recrystallization. Subse-
quent rise in strength with further increase in annealing time till
60 s could be dictated by the increase in austenite fraction, which
contributed to the increased fraction of the harder constituents
i.e. martensite and bainite.
Fine ferrite grain size ( r5 mm) not only contributed signifi-
cant grain boundary strengthening (  350 MPa) but also
improved the ductility by activating more number of slip-
systems during plastic deformation, compared to the coarser
grains, Fig. 9(b). Besides high strength, RTA treated samples from
the ferrite-fibrous martensite starting structure showed the best
strain-hardening ability possibly due to the bimodal ferrite grain
structure with more uniform distribution of fine-martensitic
islands, Fig. 9(c) [16,27,29–31,36]. Fine martensitic islands are
expected to improve the strain-hardening ability by generating
geometrically necessary dislocations adjacent to them in the
ferrite matrix to accommodate the transformation strain
[16,27,29–31]. Samples heated at a slow-rate to the inter-
Fig. 7. Microstructures developed after slow heating and inter-critical annealing
(850 1C, 30 s) of (a) ferrite–pearlite and (b) ferrite-fibrous martensite starting critical temperature showed much lower strength values but
structures; (c) microstructure obtained after sub-critical annealing (600 1C, 4 h), better uniform elongation with respect to the RTA treated
followed by RTA (850 1C, 30 s) and water-quenching. Ferrite and martensite are samples, Fig. 8. This can be related to the coarser ferrite grain
indicated by arrows. size with uniform distribution of martensite in a network fashion
in slowly heated samples.
samples with respect to the sub-critically annealed samples,
which showed discontinuous yielding with yield strength,
YS  320–487 MPa, is very much expected as RTA developed 3.4.1. Fractographic studies
dual-phase or multi-phase microstructures containing martensite Fractographic studies on the tensile tested specimens of RTA
and bainite. Taking account of the high-strength of RTA samples, samples showed the presence of voids, Fig. 10, suggesting that the
uniform elongation, eu  10% and total elongation, et  20% can be failure occurred predominantly by the ductile fracture mechanism
considered as satisfactory, Table 1. Comparing between different irrespective of the starting microstructures and the annealing
starting microstructures, the best tensile properties obtained in conditions. In terms of starting microstructures, the average void
A. Karmakar et al. / Materials Science & Engineering A 564 (2013) 389–399 397

Fig. 9. Variation in: (a) yield strength and (ii) total elongation of RTA treated
samples for different starting microstructures as the function of annealing time at
850 1C; (c) strain hardening rate of the RTA treated samples.

size decreased in the following sequence: ferrite–pearlite Fig. 10. SEM fractographs of the RTA treated samples for different starting
(1.5370.15 mm)4ferrite-blocky martensite (1.0470.12 mm)4 microstructures: (a) ferrite–pearlite, (b) ferrite-blocky martensite and (c) ferrite-
ferrite-fibrous martensite (0.8870.10 mm), Fig. 10. Occasionally fibrous martensite. Quasi-cleavage region is indicated by arrow in (b).
in the RTA samples, evidence of quasi-cleavage type of fracture has
been observed, Fig. 10(b). Those features were found only in case of
ferrite–pearlite and ferrite-blocky martensite starting structures RTA treated samples, besides high strength and strain-hardening
and possibly resulted from the cracking or shearing of coarse ability.
blocks of martensite. Quasi-cleavage fracture, however, is not For further understanding of the fracture mechanism, cross-
expected to affect the ductility as severely as the cleavage fracture. sections of the broken tensile specimens perpendicular to the
Predominance of ductile fracture by void nucleation, growth and fracture surfaces have been polished and studied under SEM. Fine-
coalescence and the complete absence of cleavage facets (repre- scale voids formed by the interface decohesion around the small
sentative of brittle fracture) ensured satisfactory ductility in the martensitic islands in the RTA treated sample of ferrite-fibrous
398 A. Karmakar et al. / Materials Science & Engineering A 564 (2013) 389–399

4. Conclusions

Inter-critical annealing has been carried out using two differ-


ent heating rates on 80% cold-rolled samples having different
starting microstructures, namely ferrite–pearlite, ferrite-blocky
martensite and ferrite-fibrous martensite. Sub-critical annealing
has also been carried out to understand the microstructural
development in case of slow inter-critical annealing. Major con-
clusions derived from the study are summarized below.

 In terms of the starting microstructures, ferrite-fibrous mar-


tensite structure showed finest ferrite grain sizes (3–6 mm)
irrespective of the annealing treatments, compared to ferrite–
pearlite (9–17 mm) and ferrite-blocky martensite (6–12 mm)
starting structures.
 Fine scale arrangement of alternate layers of ferrite and
martensite resulted in more uniform distribution of fine
cementite (y) particles enriched with Mn and finally, small
martensite islands after annealing microalloy precipitates,
y-particles and even small austenite islands retard the ferrite
recrystallization and grain-growth kinetics, which helps in
generating fine ferrite grain size after inter-critical annealing.
 In terms of heating rate, slow heating ensures more uniform
distribution of martensite in the ferrite matrix as a result on
austenite formation not only by the dissolution of y, but also
along the recrystallized ferrite grain boundaries. Although it
cannot restrict ferrite grain growth.
 Due to the ultra-fine ferrite grain size (or even bimodal grain
structure) and more uniform distribution of martensite
islands, ferrite-fibrous martensite microstructure provided
the best combination of strength, ductility and strain hard-
ening ability, compared to the other starting structures. Also
uniform distribution avoided the risk of martensite cracking
and improved the ductility.

Acknowledgments
Fig. 11. SEM micrographs showing the voids along the transverse-section just
beneath the fracture surface. (a) Bimodal ferrite grain structure in RTA treated
ferrite-fibrous martensite sample, where (b) the voids nucleated preferentially in The authors acknowledge financial support from the Council of
the regions containing fine-ferrite grains and martensite islands (indicated by Scientific and Industrial Research (CSIR), New Delhi and the
arrow in (a)). provision of research facilities from the Department of Metallur-
gical and Materials Eng., Steel Tech. Centre and Central Research
Facility in Indian Institute of Technology (I.I.T.) Kharagpur.
martensite starting structure, Fig. 11(a,b). Void-density was much
lower in the coarse (recrystallized) ferrite grains, Fig. 11(a), which References
might have allowed the plastic deformation (by void growth) to
happen, and thereby, improved the strength-ductility combination [1] P. Ghosh, B. Bhattacharya, R.K. Ray, Scr. Mater. 56 (2007) 657–660.
of the bimodal grained samples. Beneficial effect of bimodal grain [2] S.G. Chowdhury, E.V. Pereloma, D.B. Santos, Mater. Sci. Eng. A 480 (2008)
540–548.
structures on the tensile properties of low-carbon steel has been [3] R.O. Rocha, T.M.F. Melo, E.V. Pereloma, D.B. Santos, Mater. Sci. Eng. A 391
discussed by the present authors in a recent article [36]. On the (2005) 296–304.
other hand, refinement in ferrite grain size and martensitic island [4] V. Massardiera, A. Ngansopa, D. Fabre guea, J. Merlin, Mater. Sci. Eng. A 527
(2010) 5654–5663.
size can improve the tensile ductility of dual-phase steel [29–31]. [5] R. Song, D. Ponge, D. Raabe, ISIJ Int. 45 (2005) 1721–1726.
With respect to the coarse-grained dual-phase structures, fine [6] R. Ueji, N. Tsuji, Y. Minamino, Y. Koizumi, Acta Mater. 50 (2002) 4177–4189.
structures ensure more uniform strain-distribution and higher [7] N. Tsuji, T. Maki, Scr. Mater 60 (2009) 1044–1049.
[8] K. Hase, N. Tsuji, Scr. Mater. 65 (2011) 404–407.
martensitic plasticity, which avoids the brittle fracture by marten- [9] R.A. Grange, Metall. Trans. 2 (1971) 65–78.
site cracking and promotes the ductile fracture by interface decohe- [10] H. Azizi-Alizamini, M. Militzer, W.J. Poole, ISIJ Int. 51 (2011) 958–964.
sion [29–31]. Coarse band of martensite (and bainite) as observed [11] N. Nakada, Y. Arakawa, K.S. Park, T. Tsuchiyama, S. Takaki, Mater. Sci. Eng. A
553 (2012) 128–133.
after the RTA treatment of cold-rolled ferrite–pearlite and ferrite- [12] H. Zakerinia, A. Kermanpur, A. Najafizadeh, Mater. Sci. Eng. A 528 (2011)
blocky martensite samples can be detrimental for the tensile 3562–3567.
ductility due to increased risk of martensite cracking [29–31]. [13] K. Hono, M. Ohnuma, M. Murayama, S. Nishida, A. Yoshie, T. Takahashi, Scr.
Mater. 44 (2001) 977–983.
Banded ferrite-martensite morphology has also been reported to
[14] T. Ogawa, N. Muruyama, N. Sugiura, Y. Yoshinaga, ISIJ Int. 50 (2010) 469–475.
reduce the work-hardening rate and hence, the uniform elongation [15] Y. Okitsu, N. Tanaka, N. Tsuji, Scr. Mater. 60 (2009) 76–79.
[30,31]. [16] M. Calcagnotto, D. Ponge, D. Raabe, Mater. Sci. Eng. A 527 (2010) 7832–7840.
Considering the requirements in automotive grade steels [17] C. Lesch, P. Alvarez, W. Bleck, J.Gil Sevillano, Metall. Mater. Trans. 38A (2007)
1882–1890.
texture and formability of the inter-critically annealed samples [18] P. Alvarez, C. Lesch, W. Bleck, H. Petitgand, J. Schöttler, J.Gil Sevillano, Mater.
will be carried out in future. Sci. Forum 500–501 (2005) 771–778.
A. Karmakar et al. / Materials Science & Engineering A 564 (2013) 389–399 399

[19] V. Andrade-Carozzo, P.J. Jacques, Mater. Sci. Forum 539–543 (2007) [28] D.Z. Yang, E.L. Brown, D.K. Matlock, G. Krauss, Metall. Mater. Trans. A 16
4649–4654. (1985) 1385–1392.
[20] T. Lolla, G. Cola, B. Narayanan, B. Alexandrov, S.S. Babu, Mater. Sci. Technol. [29] N.J. Kim, G. Thomas, Metall. Mater. Trans. A 12 (1981) 7832–7840.
27 (2011) 863–875. [30] S. Sun, M. Pugh, Mater. Sci. Eng. A 335 (2002) 298–308.
[21] S.M. Hasan, A. Haldar, D. Chakrabarti, Mater. Sci. Technol. 28 (2012) 823–828. [31] M. Majinani, W.J. Poole, Metall. Mater. Trans. A 38 (2007) 328–339.
[22] J. Huang, W.J. Poole, M. Militzer, Metall. Mater. Trans. A 35 (2004) [32] G.R. Purdy, D.H. Weichert, J.S. Kirkaldy, Trans. TMS-AIME 230 (1964)
3363–3375. 1025–1034.
[23] H. Azizi-Alizamini, M. Militzer, W.J. Poole, Metall. Mater. Trans. A 42 (2011) [33] Xue-Ling Cai, A.J. Garratt-Reed, W.S. Owen, Metall. Mater. Trans. A 16 (1985)
1544–1557. 543–557.
[24] R.R. Mohanty, O.A. Girina, N.M. Fonstein, Metall. Mater. Trans. A 42 (2011) [34] H. Beladi, G.L. Kelly, P.D. Hodgson, Metall. Mater. Trans. A 38A (2007)
3680–3690. 450–463.
[25] F.G. Caballero, C. Capdevilla, C. Garcia De Andres, ISIJ Int. 43 (2003) 726–735. [35] S. Patra, S. Roy, Vinod Kumar, A. Haldar, D. Chakrabarti, Metall. Mater. Trans.
[26] N. Peranio, Y.J. Li, F. Roberts, D. Raabe, Mater. Sci. Eng. A 527 (2010) A 42A (2011) 2575–2590.
4161–4168. [36] S. Patra, Sk. Md. Hasan, N. Narasaiah, D. Chakrabarti, Mater. Sci. Eng. A 538
[27] M. Calcagnotto, D. Ponge, D. Raabe, ISIJ Int. 52 (2012) 874–883. (2012) 145–155.

Das könnte Ihnen auch gefallen