You are on page 1of 11

Materials and Design 100 (2016) 4757

Contents lists available at ScienceDirect

Materials and Design


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

Effect of critical temperatures on microstructures and mechanical


properties of NbTi stabilized IF steel processed by multiaxial forging
Sumit Ghosh ,1, Ajay Kumar Singh, Suhrit Mula
Department of Metallurgical and Materials Engineering, Indian Institute of Technology Rookee, Roorkee 247667, Uttarakhand, India

a r t i c l e

i n f o

Article history:
Received 10 December 2015
Received in revised form 10 January 2016
Accepted 19 March 2016
Available online 24 March 2016
Keywords:
Interstitial free steel
Thermomechanical simulation
Critical temperatures
Multiaxial forging
Mechanical properties
Transmission electron microscopy (TEM)

a b s t r a c t
In the present study, evolution of microstructure during multiaxial forging (MAF) of a NbTi stabilized IF steel
and its mechanical properties have been investigated. The forging schedule was designed on the basis of critical
temperatures Ar3, Ar1 (evaluated from dilatometric curve through thermomechanical simulator) and recrystallization stop temperature, Tnr (determined from Boratto equation). MAF was performed for 5 cycles in 3 different
phase regimes; in pure -region (1050 C), transformation zone (800 C) and pure -region (650 C). The
deformed samples were cooled by normal air cooling. EBSD and optical microscopy investigation conrmed the
formation of ne ferrite grains (~5 m) due to strain induced transformation of unstable at 800 C and ultrane
ferrites (~1 m) through subgrains formation at pure -ferritic region at 650 C. The specimen forged in pure region showed a 4-fold improvement of yield strength (YS) compared to that of the starting material (141 MPa)
without much interfering its ductility (25%). This is ascertained to the development of bimodal grain structures
and formation of ultrane carbide precipitates which were conrmed by EBSD and TEM analysis. The theoretical
YS was estimated through analysis of different strengthening mechanisms and found to be highly corroborated
with the experimentally obtained result.
2016 Elsevier Ltd. All rights reserved.

1. Introduction
Interstitial free (IF) steels are widely used in applications starting
from automotive body parts to electronic components as well as enamel
wares to house hold appliances and structural industries due to their
wide range of mechanical properties and high formability [1,2]. The
major drawback of these steels is the low level of yield strength.
Capdevila et al. [3] explained that IF steels have interstitial-free bcc ferrite matrix which results low yield strength and high strain rate sensitivity. Various research groups throughout the world are trying to
improve the mechanical strength without much affecting their formability and/or ductility. Various mechanisms are proposed for the occurrence of grain renement during thermomechanical deformation
depending on the material and processing methods and temperatures
employed. Lim et al. [4] performed multiaxial compression studies on
-FeC alloy at room temperature. The grain renement was explained
on the basis of intersection of micro-bands formed due to strain localization leading to formation of subgrains. These subgrains subsequently
rotated to form ultra ne crystallites. The formation of substructures
due to multiaxial forging (MAF) of a HSLA steel was reported to be generally of equiaxed bounded by high angle boundaries [5]. Deep

Corresponding author.
E-mail address: sumit.rkvm@gmail.com (S. Ghosh).
1
1st and 2nd authors have equal contribution.

http://dx.doi.org/10.1016/j.matdes.2016.03.107
0264-1275/ 2016 Elsevier Ltd. All rights reserved.

drawability of P-added ferritic IF steel was reported to improve as the


nish rolling temperature decreased from 620 to 560 C [6]. Bhowmik
et al. [1] stated that generation of submicron grains during MAF of IF
steel lead to tremendous increase in the tensile strength with some
loss of ductility. The formation of micro-voids in cryogenically rolled
IF steel found to reduce YS in comparison to that obtained for the sample rolled at ambient temperature [7]. The microstructural renement
of IF steel occurred monotonously with increasing the number of
ECAP passes [8]. Both yield and ultimate strength increased with the
number of passes, while the ductility lost after 1 pass was partially
regained at higher number of passes. From the above studies, it can be
concluded that the grain size strengthening is one of the vital methods
where one can expect an improvement in the mechanical strength
without much affecting its ductility.
Nowadays, MAF is known to be a unique processing technique to
obtain ultrane grained (UFG) microstructure [9]. Alexander [10]
proposed that MAF exhibits certain advantages over other processes
such as ECAP or rolling, e.g. it is a simple process based on the free
forging operation. But, Kundu et al. [11] stated that the extent of
work hardening is very less in the MAF-processed materials. This is
attributed to the increase in strain rate sensitivity with strain in
high strength materials. So, enhancement of the mechanical strength
without much affecting the formability is a challenging job. It can be
noticed from the literature that, in most cases, the improvement of
the YS is also limited to b500 MPa for the thermomechanically processed IF steels. Rarely, the YS is reported to be over 600 MPa; and in

48

S. Ghosh et al. / Materials and Design 100 (2016) 4757

that case the total elongation is reported to be very poor, e.g., only
7.5% [1]. To the best of our knowledge, hardly any literature is available on the effect of different critical temperatures on the thermomechanical behavior (i.e. on MAF) of IF steel. Therefore, aim of the
present work is to investigate the inuence of MAF processing parameters on the microstructures and mechanical properties of a
NbTi stabilized IF steel deformed in pure -region, transition
region and pure -phase regime. And also examine the possibilities
of achieving high YS with signicance ductility, moderate YS with
good ductility etc. by simple MAF which is industrially reliable. In
the current investigation, the microstructural evolution due to MAF
at different critical regions and quantitative results were studied in
detail by light microscopy, EBSD and TEM. The accumulated dislocation densities were calculated using X-ray diffraction line prole
analysis (XRDLPA) to explain the strengthening mechanisms. The effect of temperature and strain hardening exponent on deformation
behavior were also investigated to correlate the YS and uniform
elongation. The YS obtained through analysis of strengthening
mechanisms is discussed in detail and the same was correlated
with that of the experimentally obtained results. A quantitative
fractography analysis was also carried out to correlate % elongation
of the respective samples.

Fig. 1. Schematic design of multi-axial forging for 1 cycle.

for the calculation of Tnr (recrystallization stop temperature). Using


the Boratto equation (as shown below), the Tnr was estimated to be
~970 C for the present steel.

2. Material and experimental procedure


The material, i.e. interstitial free (IF) steel used for the present
study was supplied by TATA Steel, Jamshedpur, India. The chemical
composition obtained by optical emission spectroscopy (Spectrolab,
Germany) analysis is given in Table 1. The dimensions of the specimen used for the multiaxial forging were in the ratio of 1.5:1.22:1.0
[12]. The samples were machined to obtain the required size of
30 mm 24.5 mm 20 mm. In the multiaxial forging (MAF) process,
the specimen was rotated by 90 after each pass and a uniaxial compressive true strain of ~ 0.4 was applied to the longest side at each
MAF pass. The schematic diagram shown in Fig. 1 represents to one
complete cycle of MAF process.
Assuming the conservation of volume and material isotropy, this procedure enables to maintain the initial dimensional ratio at the end of
each pass [12]. In this work, the MAF performed in the three different
phase regimes, i.e. in pure -region (at 1050 C), transformation
zone (at 800 C) and pure -region (at 650 C). The dilatometry test
was performed in the Gleeble-3800 thermomechanical simulator to obtain the useful critical temperatures such as austenite to ferrite start
transformation temperature (Ar3) and austenite to ferrite nish transformation temperature (Ar1).
A cylindrical shaped sample was heated up to the temperature of
1200 C at a heating rate of 5 C/s and then the sample was cooled
down to room temperature maintaining the same cooling rate (i.e.
5 C/s). The changes in the slope of the dilatometry curve can be observed during the phase transformation due to expansion/contraction
of the specic volume of different phases [13]. At the atmospheric pressure, the pure iron exists in BCC ferrite () crystal structure up to
b912 C and beyond this temperature, it transforms to FCC austenitic
() structure. During the transformation of ferrite to austenite, the
atomic volume changes by 1% [13]. The critical temperatures, Ar3 and
Ar1 obtained from the cooling curve (as shown in Fig. 2), respectively,
are 860 and 701 C. Boratto et al. [14] postulated a standard equation


p 
p
Tnr 887 464C 6445Nb644 Nb 732V230 V 890Ti
363Al357Si
The samples were rst homogenized at temperature of 1200 C for
1 h and then cooled down to the forging temperature within the furnace. The MAF were performed in a Birson friction screw press at
~ 1050, ~ 800 and ~ 650 C. Graphite powder mixed with acetone was
used for the lubrication during the forging to cause relatively better homogeneous deformation. The samples were held in the furnace at the
particular temperature at least for 10 min to achieve the forging temperature after the each pass. The MAF was performed at a strain rate
of about 10 s1 and an equivalent total true strain of 6 (0.4 15; 15
passes for 5 cycles) was maintained for each sample. After completion
of the forging operation (5 cycles = 15 passes) each forged sample
was air cooled to room temperature. The microstructural analysis was
performed using optical microscopy (Leica DMI 5000M), FE-SEM,
EBSD and transmission electron microscopy (TEM). EBSD analysis was
performed using FEI-Quanta 200FE-SEM equipped with the TSL data acquisition system. For EBSD analysis, specimens were mechanically
polished up to ne cloth using colloidal silica followed by electro
polished in an electrolyte of 20% perchloric acid in methanol at
40 C using 21 V for 50 s. The EBSD scans were performed on electrochemically polished samples with a step size of 0.5 m and subsequent
analysis was performed using TSL-OIM software. To study the dislocation density in MAFed specimens, X-ray diffraction (XRD) study was
carried out at scan rate of 0.5/min using Cu K radiation (1.5409 A)
in a Bruker AXS D8 Advance instrument. The transmission electron microscope (FEI Technai 20 G2S-Twin) was operated at 200 kV. The samples for TEM analysis were prepared by thinning down the thickness to
~ 100 m through grinding on silicon carbide coated emery papers of

Table 1
Chemical composition (wt.%) of the IF steel obtained by optical emission spectroscopy analysis.
Element

Mn

Si

Al

Cu

Ni

Nb

Ti

Fe

wt.%

0.005

0.52

0.005

0.02

0.009

0.048

0.013

0.01

0.016

0.05

0.006

99.3

S. Ghosh et al. / Materials and Design 100 (2016) 4757

Fig. 2. Dilatometry curve for TiNb microalloyed steel obtained through Gleeble-3800.

800, 1200, 1500 grit size. The samples of 3 mm diameter disc were
punched out from this thin foil. The disc samples were electropolished using a FEI twin jet electro polisher in a solution of 90% methanol + 10% perchloric acid kept at 20 C temperature. The electropolishing was carried out at 40 V.
The hardness measurements were performed using a Vickers hardness tester (FIEVM50 PC) with an applied load of 10 kg for a dwell
time of 30 s. The tensile tests were conducted on ASTM E8 sub-size specimens (10 mm gauge length, 3 mm width and 2 mm thickness). The
tests were carried out at a constant strain rate of 5 104 s1 using a
H25 K-S Tinius Olsen machine. In each case, at least 3 samples were tested to verify the reproducibility of the results. The fractured surface of
the tensile specimens was analyzed using a scanning electron microscope (ZEISS, 51-ADD0048) at an operating voltage of 15 kV to determine the mode of failure under uniaxial tensile loading.
3. Results and discussion
3.1. Microstructural investigation
The optical microstructure of as-cast specimen is shown in Fig. 3a.
The cast inhomogeneous microstructure consists of different size ferrite
grains; proeutectoid smaller size ferrites as the main micro-constituents
along with larger size ferrites. No pearlite is present in the microstructure as the C content is very less (0.005%). The annealing treatment
was carried out at 1200 C for 1 h to obtain uniform initial grains. The
optical microstructure of the homogenizing annealed steel (hence onwards called as starting material) is shown in Fig. 3b. The annealing
treatment led to the formation of coarse grained ferrites as compared

49

to the cast structure. The average grain size is measured by line intercept method and the average grain size of the as-cast and starting material samples are 210 and 260 m, respectively. During the homogenizing
annealing, all alloying elements are dissolved in austenite grains and
new austenite grains are developed. During cooling, pro-eutectoid ferrite nucleated at the austenite grain boundaries at Ar3 temperature
and grew further to reach room temperature. Alloying elements are
likely to precipitate out in the form of carbides during cooling.
Earlier in the Experimental section, we have already mentioned that
the recrystallization stop temperature of the present steel was estimated to be ~ 1000 C, and critical temperatures Ar1 (701 C) and Ar3
(860 C) were obtained from dilatometric analysis. On the basis of the
above critical temperatures, the starting material samples were deformed by multi-axial forging for 5 cycles at 3 different temperatures:
at 1050 C (recrystallization region, ), 800 C ( transition region)
and 650 C (pure region). In each case the deformed specimen was
cooled down to room temperature by normal air cooling. Initially, the
samples were homogenized for 1 h at 1200 C, and then, multi-axial
forging (MAF) were carried out maintaining a constant temperature of
1050 C. After rst cycle of forging, the sample was kept in the furnace
to attain a temperature of 1050 C before proceeding to next cycle of
forging. Thus, grain renement is promoted in each pass by repeated deformation of recrystallized austenite grains. Finally, a recrystallized
small size austenite grains structure is produced by MAF at 1050 C
( region) from the deformed austenite grains. After deformation
when normal air cooling was employed, the pro-eutectoid ferrites
were nucleated along recrystallized austenite grain boundaries at
around Ar3, and complete ferritic transformation took place at room
temperature (C = 0.005%). Thus, the room temperature microstructure
of the specimen deformed in -region consists of equi-axed ferrites
(Fig. 4a) with an average grain size of ~ 35 m (as shown in Fig. 4b).
Cuddy et al. [15] also found uniform equi-axed ferritepearlite microstructure in low carbon Nb microalloyed steel after hot rolling at
above recrystallization temperature followed by air cooling. The grain
growth of the pro-eutectoid ferrites are expected to take place during
the cooling process after forging. As stated by Yang et al. [16], coarsening
of the austenite grains is expected to be hindered during the period of
intermittent heating because of the presence of ne carbide and/or nitrides precipitates. Arribas et al. [17] showed that the presence of ne
TiN particles inhibits grain growth of austenites during inter-pass
delay times and maintained a ne recrystallized austenite grains structure. Hu et al. [18] demonstrated that the inhibiting the grain growth
may also be attributed to the solute drag effect promoted by Nb in
solid solution. Gong et al. [19] recently investigated the dissolution kinetics of NbC and (Ti, Nb)C in HSLA steels during holding period at
1200 C of the hot rolled steel (850 C). They reported that rate of austenite grain coarsening was less for the NbTi steel than that of the Nb
steel because of the high temperature stability of (Ti, Nb)C. By
thermomechanical control processing, Zhang et al. [20] reported to develop a high strength re resistant steel having high Nb and low Mo. The

Fig. 3. Optical microstructure of IF steel: (a) as cast sample (b) annealed at 1200 C.

50

S. Ghosh et al. / Materials and Design 100 (2016) 4757

high YS at 600 C was ascertained to the formation of cementites and


precipitation of complex carbides (Nb,Mo,V)C of b 10 nm size in ferritic
matrix.
During deformation in transition region (just below Ar3) at
800 C, complete recrystallization does not occur between the deformations passes, and the deformation strain is retained from one pass to the
next pass. The retained strain thus increases number of nucleation sites
for transformation. Therefore, a ne ferritic grains (FF) structure
is likely to form, which is also observed in the present study as shown in
Fig. 4c. The microstructure is consisted of FF (av. grain size of ~ 5 m)
embedded with larger size ferrites (av. size ~ 32 m, estimated from
300 grains by line intercept method). At the beginning of the deformation, the microstructure is composed of mainly smaller size nucleated
ferrites (just started to nucleate from ) and retained austenites.
Taking into account that austenite is a low stacking fault energy material, a typical dislocations substructure is formed inside the remaining
austenitic grains during deformation in this region. The dislocations

cell walls are regions of high free energy and become sites for ferrite nucleation, and thus lead to strain-induced transformation of the unstable
austenite. Analogous to the microstructural evolution observed by Junior et al. [21], it is suggested that the transformation occurs
with the ferrite nucleation sites in the remaining austenite grains.
Thus, the application of large deformation tends to: (i) enhance
transformation; and (ii) activate dynamic phenomena such as dynamic
recrystallization and strain-induced dynamic phase transformation. As a
result, MAF of IF steel in this region followed by air cooling formed a ne
equiaxed ferrite grains (~5 m).
The maximum grain renement was obtained when the thermomechanical treatment was carried out in the pure ferritic region at 650 C
(below Ar1). Room temperature microstructure is composed of two
kinds of ferrite grains: approximately 70% of the grains having an average size ~2527 m and 30% of the grains (subgrains) with an average
size of ~ 13 m (as shown in Fig. 4g). When the coarse ferrite grains
are repeatedly deformed, they get strain hardened and nally recovered

Fig. 4. Optical microstructure after (a) ve cycle MAF at 1050 C. (b) Corresponding average grin size distribution. (c) Optical microstructure after ve cycle MAF at 800 C.
(d) Corresponding EBSD image with average grain size distribution. (e) and (f) Grain boundary misorientation prole. (g) Optical microstructure after ve cycle MAF at 650 C. (hi)
Corresponding EBSD image with average grain size distribution. (j) and (k) Grain boundary misorientation prole. (UFF ultrane ferrites; FF ne ferrites).

S. Ghosh et al. / Materials and Design 100 (2016) 4757

51

Fig. 4 (continued).

and recrystallized to generate new equiaxed grains. Multidirectional


forging for a large number of cycles leads to the accumulation of a
large strain, which results in the formation of high density dislocation
substructures both at the ferrite boundaries and inside the grains.
These substructures nally yield ne grained ferrites [18]. It is generally
agreed that lowering the deformation temperature can promote deformation induced ferrite transformation (DIFT) and decrease the ferrite
grain size [22]. The strain accumulation and then change in the strain
path in all the 3 directions during MAF can introduce a large amount of
micro-shear bands in various directions. This promotes dislocations activation, accumulation and rearrangement between them, increase in the
misorientation of grain boundaries, conversion of the low angles grain
boundaries to high angles boundaries, and nally leads to the formation
of ultra ne ferrites [23]. Development of a trimodal grains structure
(large pancaked austenite, ultra-ne austenite and submicron ferrite
grains) was achieved during deformation of a duplex low-density Fe
18Mn8Al0.8C steel at 1000 C [24]. Formation of micro-shear bands
followed by dynamic recrystallization of austenites and the DIFT reported to be the dominant mechanisms for the formation of the trimodal

microstructure. In addition, due to the presence of micro-alloying elements, such as Nb and Ti in the present steel, high density dislocations
produced during the MAF are favorable for nucleation sites of Nb/Ticarbonitride precipitates [25]. These precipitates also play an important
role in the microstructural renement during thermomechanical [26].
At high temperature, the dislocations produced during MAF undergo a
recovery process associated with dislocation movement and absorption.
This leads to the formation of low density dislocations in the microstructure and form larger size recrystallized grains.
EBSD analysis has been carried out for the two selected samples;
MAFed at transition zone and pure -region, which showed signicantly rened microstructures after 5 cycles of forging. The EBSD inverse pole gure map and superimposed grain boundary map of both
the specimens revealed a severely deformed equiaxed ne grain structure, as shown in Fig. 4d and h, respectively. Fine ferrite grains embedded within the larger size ferrites regions can be visible from the EBSD
image (Fig. 4d and h). The grain colors are determined by the orientation of each grain as shown in the inset unit triangle of the gures.
The orientation image microstructure in the specimen deformed in

52

S. Ghosh et al. / Materials and Design 100 (2016) 4757

transition region resulted in partially ner grains (~5 m) along


with larger sized grains (~32 m) structure (Fig. 4d). The distribution
of average grain size is shown in Fig. 4e. It exhibited that the grains
are mostly with low angle grain boundaries (Fig. 4f). Large amount of
low angle boundaries can be attributed to the presence of comparatively
large sized grains. Fig. 4h reveals the deformed structure of the sample
forged at -region and the average grain size distribution is presented
in Fig. 4j. This shows the formation of ultrane grains with a size of 1
3 m (~ 30%) along with larger sized grains of ~ 2527 m (70%). A
large number of ultrane grains can clearly be observed in the magnifying EBSD image shown in Fig. 4i. It can be identied that the new grains
were developed along the grain boundaries of the pre-existing ferrite
grains. From the analysis of the microstructures, it can also be noticed
that the misorientation angle of the grain boundaries (Fig. 4k) is comparatively higher than that of the transition region's forged sample.
This is due to the fact of the dynamic recrystallization of ferrite grains
during MAF at pure -ferritic region [27].
TEM images of the specimens MAFed at 650 C followed by air
cooling are shown in Fig. 5ad. Fig. 5a shows high density dislocations
produced during the MAF. The presence of fcc Niobium carbide (NbC)
along with its SADP is presented in Fig. 5b. Apart from the spot pattern
(appearing as regular hexagon) of the NbC from [111] zone axis, the
SADP also reveals ring pattern as well originated from the background
consisting of ne ferrite grains. Distribution and size of NbC precipitates
can be ascertained from the TEM image shown in Fig. 5c. The image was
recorded from the diffracted beam from the spot as encircled in Fig. 5b.
It can be noticed (from Fig. 5c) that the ultrane precipitates of ~ 9
10 nm size are distributed randomly in the matrix of ne ferrites. A
large number of micro-shear bands in various directions can clearly be

identied from the TEM micrographs, as one is shown in Fig. 5d. The formation of micro-shear bands in various directions is due to the change
in the strain paths during MAF in -ferritic region.
3.2. Mechanical properties
The mechanical properties of the MAFed samples were evaluated by
Vickers hardness measurements and tensile tests, and these are compared with those of the as-cast and starting material (annealed) specimens. Vickers hardness measurements of the samples processed at
different conditions (Fig. 6) were carried out to correlate YS of the
MAFed samples, and the comparative results are shown in Table 2.
The average hardness value of the annealed sample is only 0.68 GPa.
And the hardness values increased rapidly due to straining during
MAF at different critical temperature zones. There is 2 fold increases in
the hardness value (1.39 GPa) when MAF was carried at 1050 C. The
hardness value further enhanced when MAF was done at lower temperatures i.e. below Ar3 (at 800 C) and pure ferritic region (650 C). The
maximum hardness was measured to be 1.91 GPa for the samples
forged at 650 C. This is an ~ 2.8 times higher when compared to that
of the starting material sample. Increase in the hardness is mainly due
to the renement of the ferrite grains continuously with decreasing
the processing temperature. It is well-known that the hardness is proportional to the YS of an engineering structural material [28]. Analysis
of the improvement of the YS is discussed in the relevant section (in
the stress-strain plot).
The engineering stress-strain curves of the as-cast, annealed starting
material and MAFed specimens are shown in Fig. 7. Compared to the mechanical strength of the as-cast sample (YS = 165 MPa, UTS = 280 MPa),

Fig. 5. (a) TEM bright eld image of dislocation structure of MAFed specimens at 650 C, (b) SADP spot pattern of NbC from [111] zone axis and ring pattern from the background ne ferrite
matrix, (c) TEM dark eld image, recorded from the diffracted beam from the spot as encircled (sky colour) in Fig. 4b, (d) micro-shear bands formation in various directions.

S. Ghosh et al. / Materials and Design 100 (2016) 4757

53

Fig. 6. Vickers hardness of as received annealed and MAFed specimens.

the annealed sample showed slight less YS and UTS (YS = 141 MPa,
UTS = 256 MPa) with a corresponding increase (~12%) in the ductility.
The tensile ductility of the annealed sample was estimated to be ~46%,
while the ductility of the as-cast specimen was 34%. The reduced YS
and improved elongation % of the starting material are attributed to the
formation of comparatively larger size grains (~260 m; Fig. 3b) than
that of the as-cast structure (210 m; Fig. 3a).
The YS and UTS (Fig. 7) of the specimen forged at -recrystallized region found to increase to 435 and 473 MPa, respectively, as compared to
that of the homogenized annealed specimen (YS = 141 MPa, UTS =
256 MPa). The corresponding YS and UTS further improved to 466 and
496 MPa, respectively, after MAFed at austenite-ferrite ( ) transition region (just below Ar3 region) followed by cooling under same conditions. The specimen MAFed in pure ferritic region showed the
maximum enhancement of the YS and UTS, and the corresponding
values are 601 and 628 MPa, respectively. It can be observed that the enhanced YS, 601 MPa, is N4 times than that of the starting annealed material (141 MPa). It can also be noticed (from Fig. 7) that the YS and UTS
of the MAFed specimens in 3 specic phase regions increased with
slight expense of the ductility. The ductility of the recrystallized controlled MAFed specimen decreased to ~ 31%, which corresponds to an
average grain size of ~35 m (Fig. 4a). The specimen forged in transition
region showed a ductility ~28%, which corresponds to ferrite grain size
of ~5 m embedded with larger size ferrites of ~32 (Fig. 4c-d). The specimen rolled in pure ferritic region showed a total elongation of ~ 25%.
We have already described earlier that the microstructure of the specimens deformed at the pure ferritic region comprised of 2 types of grains
(Fig. 4g-h); ultrane ferrite grains (30%, ~13 m size), large size ferrites (70%, ~ 2527 m). Therefore, it can be observed that the YS of
the air cooled specimens is signicantly improved without much affecting the ductility, especially when forged in pure ferritic phase regime.
This is mainly due to 2 reasons: (i) extensive grain renement and (ii)
dual size grain distribution obtained by MAF in this region. It is known
that the grain renement is a fundamental strategy to enhance YS of
metallic materials without much sacricing their ductility [26]. Therefore, improvement of the YS is due to the formation of ner ferrite grains
and the large size grains are responsible for retaining the ductility. Bodin
et al. [29] also observed the formation bimodal distribution of ferrite
grains in CMn steel (0.1C, 0.5Mn, 0.0045N) after intercritical rolling
within (825775) C. They also found an improvement of the YS while
retaining the ductility almost same. Gao et al. [30] reported an improvement of YS and formability of 17% Cr ferritic stainless steel due to the

Fig. 7. Engineering stress vs. engineering strain curve.

formation of in-grain shear bands and {111} recrystallized textures


when hot rolling was carried out at low nishing temperature at
700 C. Cizek et al. [31] studied the strengthening behavior of an IF
steel after the HPT at room temperature and demonstrated that the improvement of the YS occurred mainly due to the dislocation strengthening and grain size renement. The dislocation strengthening reported to
be effective up to an equivalent strain of 3; whereas grain size renement is saturated at an equivalent strain of 15. Further deformation
lead to an additional moderate strengthening by the formation of high
angles grain boundaries (HAGBs) leading to a higher HallPetch coefcient. Enhancement of the YS and UTS of the samples forged at recrystallized -region is mainly due to the formation of smaller size ferrites
(transformed from the ne recrystallized austenite grains). Thus, the
grain boundary area of austenites increases which provides a large
number of nucleation sites for ferrites during cooling. Hence, the room
temperature microstructure consists of mainly small size equiaxed ferrite grains (~35 m). On the other hand, the YS of the MAFed samples
in transition region found to further increase than that of the recrystallized controlled MAFed samples due to the additional renement
of ferrite grains. The microstructure found to consist of equiaxed ne
ferrites of ~ 5 m size embedded with larger size ferrites of ~ 32 m.
The formations of ne ferritic grains are due to the (also discussed earlier in the section of microstructural investigation): (i) enhancement of
transformation; and (ii) dynamic recrystallization and straininduced dynamic phase transformation. The strain induced dynamic
ferrite transformation was also reported to enhance mechanical properties in a TRIP steel deformed just below Ac3 temperature [32].
Intercritical deformation followed by quenching and partitioning of a
HSLA steel resulted an ultrahigh YS with a signicant ductility due to
the formation of ne grained ferritic structure by deformation induced
ferrite transformation [33]. Overall, it can be concluded that the improvement of the YS of the thermo-mechanically treated samples is
well-corroborated with the hardness values of the corresponding
samples.
The dislocation density of the MAFed samples was calculated from
the X-ray diffraction (XRD) data to analyze the strengthening effect

Table 2
The mechanical properties of the homogenized annealed and multi-axially forged IF steel specimens in different critical zones.
Forging temperature (C)

YS (MPa)

UTS (MPa)

Elongation (%)

Strain hardening exponent (n)

1050
800
650
Annealed sample
As cast sample

435 5
466 4
601 6
141 3
165 4

473 3
496 7
627 8
256 5
281 3

30.44 1.5
27.55 1
25 1.5
46 1.7
34 1.5

0.126
0.11
0.086
0.254
0.207

54

S. Ghosh et al. / Materials and Design 100 (2016) 4757

due to strain hardening. The XRD patterns of the controlled multi-axial


forged specimens are shown in Fig. 8a. All these patterns have revealed
the presence of only bcc -ferrite phase. However, width of the peaks is
found to be broadened (a magnifying view of peak broadening for 110
reection is shown separately in Fig. 8b) as the deformation temperature decreased. The broadening of the peaks occurs mainly due to the
renement of crystallite size and increase in the dislocation lattice strain
[34]. The instrumental peak broadening has been eliminated from the
total broadening using the standard broadening data of polycrystalline
Al2O3 as per the Gaussian prole t [35]:
r

r

2obs 2i

where, obs and i, respectively, are the integral breadth of the corresponding hkl reection at the full width at half intensity maxima
(FWHM) of the forged specimen and standard Al2O3.
Therefore, r is the total broadening due to the crystallite size and
lattice microstrain. The crystallite size and lattice microstrain have
been estimated from the analysis of 3 peaks of each specimen by
using the plot between Brcos vs. sin as per WilliamsonHall technique
[35]. The dislocation density (d) can be estimated from the following
equation [36] using the average crystallite size (D) and lattice
microstrain () as follows:

1
p
2 2
d 2 3

the reported values, respectively, were 3.6 1014 m2 (i.e. at a strain


of 1.15) and 6.88 1014 m2 (at a strain of 4.6). In the present study
also, almost the same level of dislocation density has been obtained
for the multiaxially forged samples in -region (3.97 1014 m2 and
-region (8.15 1014 m2), although both deformations were carried
out at higher temperatures. In case of the specimen deformed at transition region, the dislocation density found to be 4.72 1014. Therefore, it
can be noticed that the accumulated dislocation density reached the
maximum value for the specimen forged in -region compared to that
of the other 2 forged samples. This is accomplished to the accumulation
of high amount of lattice microstrain in -region compared to other 2
forged samples. The accumulated lattice strain results in the formation
of high density dislocation substructures at the ferrite boundaries as
well as inside the grains [12]. These substructures nally developed ultrane ferrite grains by recovery during cooling.
It should be remembered that usually the mean crystallite size estimated by X-ray diffraction line prole analysis (XRDLPA) is lower than
the actual grain size of the material [37,38]. Because of severe plastic deformation, the grains are usually divided into subgrains or dislocation
substructures, and these dislocation cells are separated from each
other by low-angle grain boundaries [39]. The estimated crystallite
size is generally comparable to the average domains size which diffracts
X-rays coherently to produce broadened peaks [37,38]. To correlate the
YS obtained by experimental measurement, the theoretical YS due to
dislocation strengthening has been estimated using Taylor's equation
[28] and tabulated in Table 3;

Db
p .
3

where b is the Burgers vector of the -iron. For BCC -ferrite, b a

where a = 0.28664 nm [28]. Table 3 shows the detailed values of crystallite size, lattice microstrain and dislocation density, theoretical and
experimentally obtained YS of the forged samples.
It can be noticed that the average lattice microstrain increased from
2.71 103 for the specimen forged -region (1050 C) to 4.33 103
corresponding to the specimen forged at 650 C in single phase region. Correspondingly, the dislocation density, d, found to increase
from 3.97 1014 m2 to 8.15 1014 m2. Using the same technique,
Sarkar et al. [37] also calculated the dislocation density of a single and
4 passes ECAPed specimens (at room temperature) of an IF steel and

Taylor 0 MGbhi1=2

where, 0 is the friction stress (=30 MPa), a constant ( is taken as


0.33), G is the shear modulus (= 82 GPa), b is the Burgers vector
(=0.25 nm) and M is the Taylor factor = 3 for untextured polycrystalline materials [37].
It can be noticed that for all the specimens (Table 3), the experimentally obtained YS is very close to the theoretical YS as estimated using
Taylor's equation (Eq. (3)). It can be remembered that the theoretical
value of the YS was estimated only due to dislocation density, which includes the grain size effect. It should be remembered that the dislocation density increases with decrease in the crystallite size (Eq. (2)),

Fig. 8. (a) X-ray diffraction pattern of the MAFed IF steel specimen in different processing conditions showing peak broadening. (b) Magnifying view of peak broadening for 110 reection.

S. Ghosh et al. / Materials and Design 100 (2016) 4757

55

Table 3
Crystallite size, microstrain and dislocation density values, theoretical and experimentally obtained yield strength for IF steel MAFed at different conditions.
Processing conditions

Crystallite size (nm)

Lattice microstrain (103)

Dislocation (m2)

Taylor (MPa)

y (MPa)

MAF in -region at 1050 C


MAF in - transition zone at 800 C
MAF at 650 C

82.03
69.05
46.22

2.71
3.67
4.33

3.97 1014
4.72 1014
8.15 1014

439
471
609

435
466
601

which is equivalent to grain size for ultrane structure. As we have earlier mentioned that the improvement of the YS is due to the grain size
renement, dislocation strengthening and precipitation hardening.
While, the Taylor's equation does not consider the contribution of the
YS from the precipitation strengthening. Therefore, the theoretical YS
practically would have been much higher if all the strengthening contributions were considered. Also for the theoretical calculations, the material is always assumed to be isotropic with respect to all strengthening
mechanisms. But in the practical material, always some defects are
present and it is anisotropic in nature in that sense. Therefore, experimental YS always should be lower than that of the theoretical YS if all
the strengthening components are included in the strength estimation.
So, it can be concluded that the theoretical YS of the present material
would have been much higher than that of the experimentally obtained
values after considering the contribution from the precipitation
strengthening.

3.3. Strain hardening behavior


It is well-known that the strain-hardening exponent (n) is the basic
deformation performance parameter of metallic materials and it determines the ow behavior of a material when it is being deformed [28].
Material having higher n has better formability than that of the material
having less value of n. The slope of the logarithmic plot of the equation
= Kn (log = log K + n log) is equal to the strain hardening exponent (n) [28]. In the present study, n was calculated from the true
stresstrue strain curves to analyze the effect of the deformation temperatures on the tensile behavior of the IF steel. Fig. 9 shows the variation of the YS and uniform elongation as a function of the strain
hardening exponent (n) for the thermomechanically treated samples
compared to that of the as-cast and annealed specimens. The values of
n are also summarized in Table 2 along with the other mechanical
properties.
In the cast specimen, the value of n was estimated to be 0.207. After
homogenizing annealing treatment, it increases to 0.254 due to softening of the material. The value of n is gradually decreased with increase in

the YS. The minimum value of n reaches to 0.086 corresponding to the


YS of 601 MPa, which is correspond to the specimen deformed in pure
ferritic region. The uniform elongation is found to be more for the sample having high value of n. This can be clearly evident from Fig. 9 that the
annealed specimen, which shows a large uniform elongation of 30% has
the highest value of n (0.254). Generally, the material exhibiting a low
YS and large uniform elongation demonstrates a large amount of formability [40]. Because, the materials having low YS and large uniform
elongation are usually associated with large work-hardening capacity.
The value of uniform elongation can also been related with the inverse
of the YS [40]. It can clearly be observed from Fig. 9 that the maximum
YS of 601 MPa, which corresponds to the specimen deformed in pure
ferritic region, shows a uniform elongation of only ~6.4% (n = 0.086);
whereas, the maximum value of uniform elongation reaches to ~ 30%
(n = 0.254) corresponding to the YS of 141 MPa for the annealed sample. It can be noticed that the MAF among the 3 regions, the value of n is
higher (0.126) for the specimen MAFed in pure austenitic phase region
compared that of the specimens deformed at other 2 regions, i.e.
transition region (0.11) and at pure region (0.086). The presence of
ferrites during deformation leads to high rate of strain hardening
followed by formation of very ne ferrite grains. Therefore, it can be
concluded that the specimen having high % elongation and low YS
shows high value of strain hardening exponent (n), which is essential
for better formability. It can be noticed that total elongation lies within
a narrow range between 30 and 25% after MAF at 3 different phase regimes, while YS and UTS improve signicantly. The MAFed specimen
deformed in the ferritic region showed a remarkable improvement of
YS (601 MPa as compared to 141 MPa of starting materials) with a
total elongation of ~25%; but its uniform elongation is limited to ~6.4%
only corresponding to a poor value of n (0.086). This indicates that the
formability is poor when deformed in the ferritic region. The extent of
work hardening generally is very less in the MAF-processed materials
[11]. This can be attributed to the increase in strain rate sensitivity with
strain especially for the materials of high strain rate sensitivity like IF
steel [3]. Bhowmik et al. [1] reported almost same level improvement
in the YS (600 MPa) after MAF of the IF steel; but the total elongation
was reported to be much inferior (only 7.5%) compared to that of the
present study (25%). Therefore, as compared to the previously reported
works, better combination of mechanical strength and ductility has
been achieved in the present study. The multi axial forging is associated
with heterogeneities in microstructure. We are working out to overcome
the limitations associated with the process by optimizing temperatures,
number of cycles, strain rate per pass etc. in order to obtain desired microstructure and formability. The proper optimization of the parameters
may led to the increase in the uniform elongation and formability without losing much of the strength. After achieving the goal, we will communicate our ndings to the same reputed Journal as early as possible.
3.4. Fractography analysis

Fig. 9. Variation of YS and Uniform elongation (UE) with strain hardening exponent.

The fractured surface of the tensile specimens was examined under


scanning electron microscope to analyze the mode of failure and the
fractographs are shown in Fig. 10ae. Presence of dimple marks in the
as-cast and annealed specimens indicates a ductile mode of failure.
Larger size and deeper dimple ruptures (av. size ~ 13 m) can clearly
be observed (in Fig. 10b) on the fractured surface of the annealed sample as compared to that of the as-cast sample (~9 m) (Fig. 10a). This
observation correlates well with the high amount of tensile elongation

56

S. Ghosh et al. / Materials and Design 100 (2016) 4757

Fig. 10. SEM fractographs of tensile test specimen: (a) as-cast, (b) annealed, (ce) ve cycles MAF at 1050, 800 and 650 C, respectively.

obtained in the annealed specimen. The average dimple size has been
compared to the % elongation of the respective sample processed
under different conditions (Fig. 11). Initially, both the dimple size and

ductility increased compare to that of the as-cast sample because of


the increase in the grain size due to annealing. Thereafter, (from
Fig. 11) the average dimple size found to gradually decrease with decrease in the % elongation corresponding to decrease in the processing
temperature (Fig. 11).
This is because of the increase in the strength (equivalent to brittleness) of the material (Fig. 10ce) due to the grain size renement
and strain hardening associated with the plastic deformation [18,28].
It can be remembered that the % elongation of the samples forged in
the and critical phase regions varies marginally from 31 to
28% only. This can be also be correlated with the fractographs of
the corresponding samples (Fig. 10c, d). The specimen MAFed at
pure ferritic region showed relatively smaller size shallow dimple
marks (Fig. 10e) indicating comparatively brittle appearance ductile
failure.

4. Conclusions

Fig. 11. The average dimple size vs. % elongation of the respective sample processed under
different conditions.

In the present study, we have investigated the microstructure and


mechanical properties of NbTi stabilized IF steel multiaxially forged
for 5 cycles in 3 critical temperature regions followed by normal air

S. Ghosh et al. / Materials and Design 100 (2016) 4757

cooling. On the basis of the results and their analysis, the following conclusions can be ascertained:
(1) The grain renement mechanisms are found to be inuenced by
the deformation temperatures in the 3 different phase regimes.
Grain renement during MAF at -region (1050 C) is promoted
by continuous dynamic recrystallization of deformed austenite
grains and formed ne equiaxed ferrite grains (~35 m) on normal cooling. On the other hand, formation of ne ferrite grains
(~5 m) is attributed to the strain induced transformation of unstable during MAF in transformation zone (800 C). The
ultrane ferrite grains formation (~1 m) in pure ferritic region
(650 C) occurs through subgrains formation by deformation induced ferrite transformation (DIFT).
(2) The maximum renement of grains was achieved when the specimen forged in pure ferritic region. This is attributed to the development of micro-shear bands in multiple directions, repeated
change in the accumulated strain paths to promote dislocation
activity, increase in the misorientation angles of grain boundaries
and DIFT. This has been conrmed by detailed EBSD study and
TEM investigation.
(3) Multiaxially forged sample at pure ferritic state (650 C) shows
highest value of mechanical properties (YS-601 MPa and UTS628 MPa) without much interfering the total elongation (25%)
compared to the other two cases. The improvement of the YS simultaneously with retaining high ductility is accomplished to the
fact of bimodal grain structure obtained by MAF in this region.
The ne size ferrites (average size ~ 13 m) is accountable for
the high YS, while the comparatively larger size ferrite grains
(average size ~2527 m) liable for the retained ductility.
(4) The improvement in the mechanical properties along with high
ductility can be enlightened to the presence of special features
in the microstructure such as ultrane precipitates of TiC and/
or NbC in ferrite matrix, which has been conrmed by TEM study.

Acknowledgement
The authors are highly acknowledged the TATA Steel, Jamshedpur
for providing the steel for the research purposes and Indian Institute
of Technology Roorkee for providing the research facilities to carry out
the work.
References
[1] A. Bhowmik, S. Biswas, S.S. Dhinwal, A. Sarkar, R.K. Ray, D. Bhattacharjee, S. Suwas,
Microstructure and texture evolution in interstitial-free (IF) steel processed by
multi-axial forging, Mater. Sci. Forum 702703 (2012) 774777.
[2] R. Rana, W. Bleck, S.B. Singh, O.N. Mohanty, Development of high strength interstitial free steel by copper precipitation hardening, Mater. Lett. 61 (2007) 29192922.
[3] C. Capdevila, V. Amigo, F.G. Caballero, C.G. Andres, M.D. Salvador, Inuence of microalloying elements on recrystallization texture of warm-rolled interstitial free steels,
Mater. Trans. 51 (2010) 625634.
[4] S.M. Lim, M.E. Wahabi, C. Desrayaud, F. Montheillet, Microstructural renement of
an FeC alloy within the ferritic range via two different strain paths, Mater. Sci.
Eng. A 532 (2007) 460461.
[5] A.K. Padap, G.P. Chaudhari, V. Pancholi, S.K. Nath, Microstructural evolution and mechanical behavior of warm multi-axially forged HSLA steel, J. Mater. Sci. 47 (2012)
78947900.
[6] W.M. Guo, Z.C. Wang, L. Sheng, X.B. Wang, Effect of nish rolling temperature on microstructure and mechanical properties of ferritic rolled P-added high strength interstitial free steel sheet, J. Iron Steel Res. Int. 18 (2011) 4246.
[7] G. Anand, A. Sinha, P.P. Chattopadhyay, Variation of tensile behaviour of interstitial
free steel rolled at cryogenic and room temperature, J. Inst. Eng. India Ser. D 93
(2013) 97103.
[8] K. Mathis, T. Krajnak, R. Kuzel, J. Gubicza, Structure and mechanical behaviour of
interstitial-free steel processed by equal-channel angular pressing, J. Alloys
Compd. 509 (2011) 35223525.
[9] H. Halfa, Recent trends in producing ultrane grained steels, J. Miner. Mater.
Charact. Eng. 2 (2014) 428469.

57

[10] D.J. Alexander, New methods for severe plastic deformation processing, J. Mater.
Eng. Perform. 16 (2007) 360374.
[11] A. Kundu, R. Kapoor, R. Tewari, J.K. Chakravartty, Severe plastic deformation of copper using multiple compression in a channel die, Scr. Mater. 58 (2008) 235238.
[12] G.A. Salishchev, O.R. Valiakhmetov, R.M. Galeyev, Formation of submicrocrystalline
structure in the titanium alloy VT8 and its inuence on mechanical properties, J.
Mater. Sci. 28 (1993) 28982902.
[13] C.G. Andres, F.G. Caballero, C. Capdevila, L.F. Alvarez, Application of dilatometric
analysis to the study of solidsolid phase transformations in steels, Mater. Charact.
48 (2002) 101111.
[14] F. Boratto, R. Barbosa, S. Yue, J.J. Jonas, Effect of chemical composition on the critical
temperature of microalloyed steels, Proceeding on Int. Conf. THERMEC1988, vol. 1
1988, pp. 383390.
[15] L.J. Cuddy, Grain renement of Nb steels by control of recrystallization during hot
rolling, Metall. Mater. Trans. A 15 (1984) 8798.
[16] J.H. Yang, Q.Y. Liu, D.B. Sun, X.Y. Li, Recrystallization behavior of deformed austenite
in high strength microalloyed pipeline steel, J. Iron Steel Res. Int. 16 (2009) 7580.
[17] M. Arribas, B. Lopez, J.M. Rodriguez-Ibabe, Additional grain renement in recrystallization controlled rolling of Ti-microalloyed steels processed by near-net-shape
casting technology, Mater. Sci. Eng. A 485 (2008) 383394.
[18] J. Hu, L.X. Du, H. Xie, X.H. Gao, R.D.K. Misra, Microstructure and mechanical properties of TMCP heavy plate microalloyed steel, Mater. Sci. Eng. A 607 (2014) 122131.
[19] P. Gong, E.J. Palmiere, W.M. Rainforth, Dissolution and precipitation behaviour in
steels micro-alloyed with niobium during thermo-mechanical processing, Acta
Mater. 97 (2015) 392403.
[20] Z.Y. Zhang, Q.L. Yong, X.J. Sun, Z.D. Li, J.Y. Kang, G.D. Wang, Microstructure and mechanical properties of precipitation strengthened re resistant steel containing high
Nb and low Mo, J. Iron Steel Res. Int. 22 (2015) 337343.
[21] A.M.J. Junior, L.H. Guedes, B. Oscar, Ultrane grain renement during the simulated
thermomechanical processing of low carbon steel, J. Mater. Res. Technol. 1 (2012)
141147.
[22] Z.Q. Sun, W.Y. Yang, J.J. Qi, A.M. Hu, Deformation enhanced transformation and dynamic recrystallization of ferrite in a low carbon steel during multipass deformation,
Mater. Sci. Eng. A 334 (2002) 201206.
[23] B. Zhao, F. Liu, G. Li, R. Xu, J. Yang, Effect of temperature and strain on microstructure
renement in low carbon microalloyed steel during multi-axis deformation process,
Met. Mater. Int. 19 (2013) 549553.
[24] A. Mohamadizadeh, A.Z. Hanzaki, A. Kisko, D. Porter, Ultra-ne grained structure
formation through deformation-induced ferrite formation in duplex low-density
steel, Mater. Des. 92 (2016) 322329.
[25] R. Priestler, P.H. Li, C. Zhou, A.K. Ibraheem, Microally precipitation in HSLA steel austenite, Microstruct. Sci. 26 (1998) 447454.
[26] T. Gladman, The Physical Metallurgy of Microalloyed Steels, The Institute of Material, London, 1997 280288.
[27] S. Ghosh, S. Mula, Thermomechanical processing of low carbon NbTi stabilized
microalloyed steel: microstructure and mechanical properties, Mater. Sci. Eng. A
646 (2015) 218233.
[28] G.E. Dieter, Mechanical Metallurgy, second ed. Mc Graw-Hill Book Co., New York,
1976 325337.
[29] A. Bodin, J. Sietsma, S. Van der Zwaag, On the nature of the bimodal grain size distribution after intercritical deformation of a carbonmanganese steel, Mater. Charact.
47 (2001) 187193.
[30] F. Gao, F.X. Yu, F.T. Liu, Z.Y. Liu, Hot deformation behavior and ow stress prediction
of ultra puried 17% Cr ferritic stainless steel stabilized with Nb and Ti, J. Iron Steel
Res. Int. 22 (2015) 827836.
[31] J. Czek, M. Janecek, T. Krajnak, J. Straska, P. Hruska, J. Gubicza, H.S. Kim, Structural
characterization of ultrane-grained interstitial-free steel prepared by severe plastic
deformation, Acta Mater. 105 (2016) 258272.
[32] A. Mohamadizadeh, A.Z. Hanzaki, S.H. Manesh, A. Imandoust, The effect of strain induced ferrite transformation on the microstructural evolutions and mechanical
properties of a TRIP-assisted steel, Mater. Sci. Eng. A 607 (2014) 621629.
[33] H. Liu, H. Sun, B. Liu, D. Li, F. Sun, X. Jin, An ultrahigh strength steel with ultranegrained microstructure produced through intercritical deformation and partitioning
process, Mater. Des. 83 (2015) 760767.
[34] T. Ungar, S. Ott, P.G. Sanders, A. Borbely, J.R. Weertman, Dislocations, grain size and
planar faults in nanostructured copper determined by high resolution X-ray diffraction and a new procedure of peak prole analysis, Acta Mater. 46 (1998)
36933699.
[35] V.D. Mote, Y. Purushotham, B.N. Dole, WilliamsonHall analysis in estimation of lattice strain in nanometer-sized ZnO particles, J. Theor. Appl. Phys. 6 (2012) 18.
[36] Y.H. Zhao, Z. Horita, T.G. Langdon, Y.T. Zhu, Evolution of defect structures during cold
rolling of ultrane-grained Cu and CuZn alloys: Inuence of stacking fault energy,
Mater. Sci. Eng. A 474 (2008) 342347.
[37] A. Sarkar, A. Bhowmik, S. Suwas, Microstructural characterization of ultrane-grain
interstitial-free steel by X-ray diffraction line prole analysis, Appl. Phys. A Mater.
Sci. Process. 94 (2009) 943948.
[38] B.D. Cullity, Elements of X-ray Diffraction, Edison Wesley, London, 1978 259270.
[39] S.M. Dasharath, C.C. Koch, S. Mula, Effect of stacking fault energy on mechanical
properties and strengthening mechanisms of brasses processed by cryorolling,
Mater. Charact. 110 (2015) 1424.
[40] D.H. Kang, D.W. Kim, S. Kim, G.T. Bae, K.H. Kim, N.J. Kim, Relationship between
stretch formability and work-hardening capacity of twin-roll cast Mg alloys at
room temperature, Scr. Mater. 61 (2009) 768771.

You might also like