You are on page 1of 10

Materials Research Express

PAPER You may also like


- Numerical investigation and experimental
Effect of cold rolling reduction on microstructure, validation of a plasticity model for sheet
steel forming
mechanical properties, and texture of deep Tales Carvalho-Resende, Tudor Balan,
Salima Bouvier et al.

drawing dual-phase (DP) steel - Strain hardening and micro-deformation


behavior in advanced DP and TRIP steels:
EBSD examinations and crystal plasticity
To cite this article: Yinghui Zhang et al 2019 Mater. Res. Express 6 096530 simulations
N Saeidi, M Jafari, F Ashrafizadeh et al.

- Effect of material inhomogeneity on the


cyclic plastic deformation behavior at the
microstructural level: micromechanics-
View the article online for updates and enhancements. based modeling of dual-phase steel
Surajit Kumar Paul

This content was downloaded from IP address 200.192.219.160 on 11/05/2022 at 19:58


Mater. Res. Express 6 (2019) 096530 https://doi.org/10.1088/2053-1591/ab2d71

PAPER

Effect of cold rolling reduction on microstructure, mechanical


RECEIVED
11 April 2019
properties, and texture of deep drawing dual-phase (DP) steel
REVISED
4 June 2019
ACCEPTED FOR PUBLICATION
Yinghui Zhang, Qiangqiang Yuan, Jieyun Ye, Xinger Weng and Zhigang Wang
27 June 2019 School of Material Science and Engineering, Jiangxi University of Science and Technology, Ganzhou 341000, People’s Republic of China
PUBLISHED
E-mail: wzgang2008cn@163.com
5 July 2019
Keywords: deep drawing dual-phase (DP) steel, cold rolling reduction, texture, r value, deep drawability

Abstract
In this study, the microstructure, mechanical properties, and texture of deep drawing dual-phase (DP)
steel subjected to different cold rolling reductions were characterized via optical microscopy (OM),
scanning electron microscopy (SEM), and x-ray diffraction (XRD) techniques. The results revealed
that the microstructure of the steel is composed of ferrite and a small amount of martensite after cold
rolling annealing. The ferrite grain size varied only slightly with the reduction. In addition, the
orientation density along the γ orientation line increased continuously with increasing cold rolling
reduction ratio. Similarly, the {112}〈110〉 and {223}〈110〉 components along the α orientation line
and the adjacent {554}〈225〉 and {332}〈113〉 components were strongly developed. The ratio of the
{111} plane texture volume fraction to that of the {100} plane texture, i.e., V{111}/V{100}, increased
initially and then decreased, reaching the optimum value at 75% reduction. The maximum strength of
the steel (i.e., 430 MPa) was realized at 55% reduction. However, the best elongation and r values
(29.8% and 1.46, respectively) were obtained at 75% reduction.

1. Introduction

Compared with the traditional high strength dual-phase (DP) steel, cold-rolled deep drawing DP steel has been
extensively researched in recent years due to its high initial work hardening rate, relatively low yield ratio,
optimal combination of strength and plasticity, and excellent deep drawability [1]. Cold deformation and
continuous annealing are indispensable processes in the preparation of cold-rolled deep drawing DP steel. The
effect of cold rolling reduction ratio on the microstructure and texture of single ferritic interstitial free (IF) extra-
deep drawing steel has been investigated in various studies [2–4]. The results revealed that a high cold rolling
reduction ratio increases the density of defects and amount of distortion in the material, thereby promoting the
nucleation of a favorable texture during recrystallization. When the reduction ratio is excessively large, the
deformation texture will be enhanced, but the stable orientation will shift after deformation, thereby affecting
the annealing texture [5]. The precipitation becomes more dispersed when the reduction ratio exceeds a certain
level, and exerts a strong pinning effect on growth of the recrystallized grains during the subsequent annealing.
This results in weakening of the γ fiber and a reduction in the index of plane anisotropy (r value) [6].
Unlike the case of IF steel, a small amount of hard pearlite structure is formed during the hot rolling of DP
steel, owing to the high carbon content of the steel. However, the pearlite colony can be broken and elongated
during cold rolling, resulting in uneven deformation of the matrix and, in turn, possible formation of other
texture features. Kang [7] and Seong [8] discussed the effects of chemical components (Al and Mo) and hot
rolling parameters on the microstructure and texture of deep drawing DP steel. Wang et al [9] investigated the
microstructure and texture evolution during annealing of deep drawing DP steel consisting of chromium and
molybdenum. The results revealed that the γ fiber remains the main component of the steel after cold rolling,
but {001}〈110〉, {112}〈110〉, and {223}〈110〉 textures, which are unfavorable for the deep drawability, were also
formed. Nevertheless, published studies on the influence of cold rolling reduction on the texture of deep
drawing DP steel are rare, and the effect of the cold rolling process on the content of unfavorable components

© 2019 IOP Publishing Ltd


Mater. Res. Express 6 (2019) 096530 Y Zhang et al

Table 1. Chemical composition of the tested steel (wt.%).

C Si Mn P S Sol.Al Mo Cr

0.016 0.022 1.56 0.016 0.0038 0.34 0.40 0.10

Figure 1. The schematic diagram of tensile sample.

remains unknown. Therefore, based on previous studies, deep drawing DP steel with chromium and
molybdenum was designed in the present study [9]. The microstructure, properties, and texture of DP steels
subjected to different cold rolling reduction have been compared. The effect of reduction on the annealing
microstructure and texture of the cold-rolled plates has been systematically investigated.

2. Materials and methods

A 25 kg DP steel was designed and smelted in the laboratory (the chemical composition of the steel is shown in
table 1). The casting billet was forged into a 130 mm×100 mm×35 mm billet, which was then heated to
1200 °C, and held for 1.5 h at temperature prior to hot rolling. The initial and finishing rolling temperatures
were 1150 °C and 920 °C during hot rolling, respectively. The billet was rolled in five passes to a thickness of
4 mm, cooled to 700 °C, and then held at this temperature for 2 h for simulated coiling. After pickling, the hot-
rolled plate was cold-rolled at reduction ratios of 55%, 65%, 75%, and 85%. The cold-rolled sheet was then wire
cut along its rolling direction into a 220 mm×70 mm sample, to simulate continuous annealing on a ULVAC-
CCT-AY-II continuous annealing simulator. During the annealing process, the plate was heated to 860 °C
(heating rate: 10 °C s−1), held at this temperature for 80 s, and then rapidly cooled to room temperature (cooling
rate: 40 °C s−1).
The tensile sample (gauge length: 50 mm) and metallographic sample taken along the rolling direction were
prepared. The mechanical properties of the materials were determined using an MTS810 tensional machine.
The Lankford parameter (also known as plastic strain ratio and denoted by the letter r) is a measure of the plastic
anisotropy and is used extensively as an indicator of the drawability of sheet metals [10]. The calculation
equation is shown in formula (1) (L0 and b0 are the initial length and width while L and b are the final length and
width, respectively). In this study, the r value was calculated by measuring the width of 15% engineering strain.
The other important factor for sheet metal formability is its ability to strain harden during forming operation. In
this paper, the strain hardening exponent (n) which is defined by the slope of the true stress-true strain curve in a
tensile test and expressed as formula (2) (σ is the true stress, ε is the true strain, k is the strength coefficient, and n
is the strain hardening exponent.). As figure 1 shows the schematic diagram of tensile sample.

e
r= b =
In ()b
b0
=-
ln ()
b
b0
(1)
ea ln ( )
L 0 b0
Lb
ln
L
L0
+ ln
b
b0

D log s
s = ken , and n = (2)
D log e

After conventional metallographic preparation, polished surfaces were etched with 4% natal acid alcohol
solution (4 ml HNO3, 96 ml C2H5OH). The microstructure of the cross-section was observed via optical
microscopy (Carl Zeiss AxioScope A1) and scanning electron microscopy (SEM; ZEISS Sigma field emission
scanning electron microscope). Furthermore, incomplete pole figures of the {110}, {200}, and {211} crystal
faces occurring at a distance of 1/4 thickness from the sample surface were measured by means of x-ray

2
Mater. Res. Express 6 (2019) 096530 Y Zhang et al

Figure 2. Scanning electron micrographs of steel plates subjected to different cold rolling reduction ratios and hot-rolled plate (a)hot-
rolled plate; (b) hot-rolled plate; (c) 55%; (d) 65%; (e) 75%; (f) 85%.

diffraction (XRD). These figures were used to determine the orientation distribution function (ODF), which was
used to evaluate the macro-texture.

3. Results and discussion

3.1. Influence of cold rolling reduction ratio on microstructure and texture of cold-rolled steel plate
Figure 2(a) shows the SEM images of hot-rolled plate. Because of the high coiling temperature (700 °C, held for
2 h), the microstructure is mainly of polygonal ferrite and some pearlite, and the average ferrite grain size is
about 5–20 μm. There are two morphologies of pearlite in figure 2(a), lamellar pearlite and degenerated pearlite
due to insufficient carbon diffusion. High magnification morphology of degenerated pearlite is shown in
figure 2(b). After cold rolling of hot rolled plate, the ferrite is prolonged and show fiber morphology along rolling
direction, while most of pearlite, especially for degenerated pearlite, is also prolongated, and a small amount of
lamellar pearlite cluster is retained and becomes smaller than hot-rolled state. Figures 2(c)–(f) shows the SEM
images of sheets subjected to different cold rolling reduction ratios. The cold-rolled deep drawing DP steel is
composed of ferrite and minor pearlite, which inherited from hot rolled plate (enclosed in the red square in

3
Mater. Res. Express 6 (2019) 096530 Y Zhang et al

Figure 3. Variation in the orientation density along the α and γ orientation lines of steel plates subjected to different cold rolling
reduction ratios. (a) α orientation line; (b) γ orientation line.

figure 2). As shown in figure 2(c), the fibrous morphology is vague and only a few deformation bands are formed
at a reduction ratio of 55%. When the reduction ratio increases, the ferrite grains are gradually elongated along
the rolling direction and exhibit fibrous morphology. The deformation bands formed at reduction ratios of 65%
and 75% (see figures 2(d) and (e)) are wider than those formed at other ratios (55% and 85%). When the
reduction ratio increases, pearlite-colony breakage begins and the diameter decreases gradually. As shown in
figure 2(f), at a reduction of 85%, the width of the fibrous morphology decreases, deformation bands are almost
eliminated, number of pearlite colonies is reduced, and diameter of the colonies decreases.
The XRD-measured macro-texture (see figure 3) shows the orientation lines of steel plates subjected to
different cold rolling reductions. On the γ orientation line, the orientation density of each component increases
with increasing cold rolling reduction ratio. The maximum orientation density along this line, especially for
{111}〈110〉 and {111}〈112〉 components, occurs at a reduction ratio of 85%. In contrast, the orientation density
varies only modestly at reduction ratios ranging from 55% to 75%, indicating that the texture of the γ fiber
improved at large cold rolling reductions. Furthermore, {112}〈110〉, {223}〈110〉, and the adjacent components
on the α orientation line are strongly developed. The peak on this line shifts, however, from {113}〈110〉 to {223}
〈110〉, when the cold rolling reduction ratio increases. This is consistent with the findings of Inagaki [11], who
proposed two formation paths, i.e., {001}〈100〉→{001}〈110〉→{112}〈110〉→{223}〈110〉 and {110}
〈001〉→{554}〈225〉→{111}〈112〉→{111}〈110〉→{223}〈110〉, for the cold-rolled texture. In addition,
Rajib reported that the stability of the α fiber in IF steel increases with increasing cold rolling reduction [12]. The
results obtained in the present study (see figure 3(a)) concur with this finding, i.e., the {223}〈110〉 becomes the
most stable orientation when the reduction ratio is increased. In addition, the largest orientation density of the α
orientation line is greater than that of the γ orientation line. This indicates that the cold-rolling-induced α fiber
deformation texture will probably be stronger than the corresponding γ fiber deformation texture.

3.2. Influence of cold rolling reduction ratio on microstructure and texture of annealed steel plate
The engineering stress-strain curves of DP steel with different cold-rolling reduction show in figure 4(a).
Figure 4(b) shows the mechanical properties of annealed steel plates subjected to different cold rolling reduction
ratios. The highest strength, i.e., 447 MPa, of deep drawing DP steel is obtained at a cold rolling reduction ratio
of 55%. The maximum elongation and r value (29.8% and 1.46, respectively) are realized at a reduction of 75%.
Figure 4(c) shows that the cold rolling reduction ratio has a small effect on the work hardening index (n value) of
DP steel. and the optimal r value occurs at a reduction ratio of 75%. This indicates that the deep drawability of
the DP steel is significantly higher than that of the traditional deep drawing DP steel consisting of chromium and
molybdenum [13].
Figure 5 shows the SEM microstructures of annealed steel plates subjected to different cold rolling reduction
ratios. The annealed structures are composed of ferrite and fine martensite as well as a small number of carbides
distributed along the grain boundaries. Compared with cold rolling reduction ratios of 75% and 85%, reduction
ratios of 55% and 65% lead to a slightly larger grain size of the annealed steel plate. The reason is that, as the

4
Mater. Res. Express 6 (2019) 096530 Y Zhang et al

Figure 4. Mechanical properties of annealed sheets with different cold rolling reduction ratios of DP steels. (a) engineering stress-
strain curves; (b) strength and elongation of DP steels; (c) n value and r value.

reduction ratio increases, the deformation stored energy will gradually rise, leading to an increase in the
nucleation rate of the recrystallized nuclei during annealing [14]. According to Park and Lee [15], the high Gibbs
free energy G associated with heavy deformation of the material results in the rapid selection of nucleation
orientation. The stored energy is then released based on the anisotropy of elastic strain. The recrystallization
driving force increases with increasing strain, and is expressed as follows:
1
P = - dg / dV = s 2d /Ed - er Er2
2
Where, P, σ, E, s d2 /Ed , and ε are the respective recrystallization driving force, residual stress, Young’s modulus,
energy density, and strain. The subscript characters d and r denote deformation and recrystallization,
respectively.
Although the ferrite grain size decreases (in general) with increasing reduction ratio, the strength at a ratio of
55% is significantly higher than at the other ratios (see figure 4(b)). The main reasons for this are as follows: at
small deformation levels, coarse pearlite colonies subjected to heavy deformation during cold rolling are more
easily broken than those in the hot-rolled plate. Therefore, the colonies in the hot-rolled plate will undergo a
phase transition to austenite, thereby resulting in a coarse-grained martensitic phase during the annealing
process. This hard phase (indicated by the red arrow in figure 4(e)) increases the strength of the material. In
addition, since the stored energy of deformation is insufficient at low pressure, the ferrite grains might be
unevenly sized (despite the completion of recrystallization) and, hence, the strength will increase due to
uncoordinated deformation. As the reduction ratio increases to 75%, the ferrite grain size becomes relatively
uniform and fine martensitic phases are distributed on the grain boundaries, as shown in figure 5(e). The
amount of martensite in the grain boundaries increases with further increases in the reduction ratio, and the
grain size is remarkably refined, which is beneficial for improving the properties of the steel.
The difference between r values is closely correlated with the variation pattern of the texture. Figure 6 shows
the texture evolution at the sample surface during cold rolling to different reductions. As shown in the figure,
each annealed sheet is characterized by a typical strong γ fiber recrystallization texture. However, the maximum

5
Mater. Res. Express 6 (2019) 096530 Y Zhang et al

Figure 5. Optical micrographs showing the microstructures of annealed steel plates that were subjected to different cold rolling
reduction ratios. (a) 55%; (b) 65%; (c) 75%; (d) 85%; (e) and (f) Scanning electron micrographs of plates subjected to reductions of
55% and 75%, respectively.

density of the fiber varies with the cold rolling reduction ratio. The maximum density is 7.6 at a reduction ratio
of 55%. The maximum strength increases with increasing reduction ratio, and in the case of the {111}〈112〉
texture (maximum strength: ∼11.3) occurs at a ratio of 85%. The γ fiber texture is significantly weakened after
annealing. Moreover, Goss components ({110}〈001〉), {554}〈225〉, and {332}〈113〉) are generated in the
annealed sheet.
As shown in figure 4(c), the r value increases initially with increasing reduction ratio and then decreases. This
is attributed to the deep drawability of DP steel, which is mainly determined by the type and intensity of the
recrystallized texture. That is, a large r value and excellent deep drawability are realized when the steel plate has a
strong {111} texture and a weak {100} texture. As the reduction ratio increases, the strength of the unfavorable
texture (α fiber) decreases gradually, whereas the strength of the favorable texture (γ fiber) increases, thereby
leading to an increase in the r value. Daniel [16] performed calculations based on a relaxed constraints (RC)
model and concluded that the {111} texture is an ideal favorable texture, owing to its relatively large average r
value. However, the average r value of the {112}〈110〉 component is slightly smaller than that of the {111}
texture. The smallest average r value is obtained for the {001}〈110〉 component.
Figure 7 shows the variation in the orientation density of the ε orientation line (〈110〉//TD), α orientation
line (〈110〉//RD), and γ orientation line (〈111〉//ND) of the annealed steel plates. The volume fraction of

6
Mater. Res. Express 6 (2019) 096530 Y Zhang et al

Figure 6. Constant j2 =45° ODF sections obtained at the sample surface for reductions of (a) 55%; (b) 65%; (c) 75%; (d) 85%.

various orientations is also shown. The α orientation line in figure 6(a) indicates that the strength is mainly
concentrated on the {112}〈110〉 and {111}〈110〉 components. The texture strength varies modestly with the
reduction ratio, but the orientation density of {112}〈110〉 decreases significantly at a reduction ratio of 85%. At
reduction ratios of 55% and 65%, the orientation strength of {111}〈112〉 along the γ orientation line (see
figure 6(b)) is lower than that of {111}〈110〉. The reverse is true for reduction ratios of 75% and 85%, i.e., the
orientation strength of {111}〈112〉 is greater than that of {111}(110), with the maximum strength occurring at a
reduction of 85%. The corresponding orientation densities (6.5 and 10, respectively) indicate that the {111}
〈112〉 orientation is the final stable orientation after annealing. Likewise, as indicated by the ε orientation line in
figure 7(c), {111}〈112〉, {554}〈225〉, and {332}〈113〉 are the strongest textures after annealing, and the strength
of the {554}〈225〉 texture is slightly higher than that of {332}〈113〉. The strength of both textures increases
gradually with increasing cold rolling reduction ratio. For example, when the cold rolling reduction ratio is 85%,
the strength of {554}〈225〉 is ∼12, which is far greater than that (5.5) associated with a reduction ratio of 55%.
The growth of these two textures can be attributed to the following two factors: on the one hand, the brass
texture ({110}〈112〉) of the hot-rolled steel plate was formed via variant selection, and steadily increased during
cold rolling; on the other hand, the {554}〈225〉 texture can form ‘in situ’ from the 20° shear bands of high-
density {111}〈112〉 grains [15]. The relevant literature indicates that {554}〈225〉 and {332}〈113〉 occur
commonly in low carbon steel and yield improved deep drawability of the material [17–19].
Owing to an increase in the orientation density of the unfavorable {100} texture, the r value decreases with
further increases in the reduction ratio (85%). Similarly, the grain size is further refined when the stored energy
increases. This refinement probably triggered nucleation of martensite, resulting in finer martensite along the
grain boundaries (than in other regions). Due to the variant of phase transition (α→γ→ a¢), generation of
the {100} texture is more likely than the formation of other textures [20]. Figure 6(d) shows the volume fraction
(calculated by the TexTools software) of {111}, {211}, and {100} plane textures associated with different
reductions and the volume ratio of {111} and {100}. As shown in the figure, the volume fraction of the {111}
texture increases with increasing cold rolling reduction ratio, but begins to decrease at a reduction ratio of 75%.
In contrast, the volume fraction of the {100} plane texture begins to decrease and, hence, the maximum
V{111}/V{100} occurs at this reduction ratio. The strength of the {111} texture increases with increasing
reduction, while the strength of the {100} texture rises, owing to the excessive reduction, and accompanying
decrease in the r value. Furthermore, the volume fraction of the {211} plane texture is larger than that of the

7
Mater. Res. Express 6 (2019) 096530 Y Zhang et al

Figure 7 Variation in the orientation density along the α and γ orientation lines of annealed steel plates and the volume-ratio influence
of various orientations along the α orientation line. (b) γ orientation line, (c) ε orientation line, and (d) volume fractions and ratio of
the volume fractions.

{111} texture and, hence, {211} is the most stable orientation. This is different from the variation pattern of
single ferritic deep-draw steels, e.g., IF steel and baking hardened (BH) steel. For the single ferrite matrix, the
{111} texture becomes the main texture after annealing. However, in the case of DP steel consisting of ferrite
and martensite, the {211} texture becomes the main texture after annealing.

4. Conclusion

(1) Sharp deformation bands occur at reduction ratios of 65% and 75%, and the diameter of the pearlite
colonies decreases gradually. The strength of the γ fiber is enhanced with increasing reduction ratio.
Moreover, at a reduction of 85%, {223}〈110〉 becomes the most stable component along the α fiber.
(2) The orientation density along the γ orientation line increases continuously with increasing cold rolling
reduction ratio; {112}〈110〉 and {223}〈110〉 on the α orientation line and the adjacent {554}〈225〉 and
{332}〈113〉 are strongly developed. At a reduction of 85%, the orientation density of the {100} texture
increases slightly, resulting in a significant decrease in the V{111}/V{100} ratio and, consequently, a slight
decrease in the r value.
(3) The cold rolling reduction ratio has only a modest influence on the strength and elongation. At a reduction
of 55%, the sample strength is relatively high. At 75% reduction, an elongation and a relatively high r value
of 29.8% and 1.46, respectively, are obtained.

8
Mater. Res. Express 6 (2019) 096530 Y Zhang et al

Acknowledgments

This is project was supported by the Key Research and Development Program of Jiangxi Province (general
project) (Grant No. 20161BBE50065); National Natural Science Foundation of china (Grant No. 51704132).

ORCID iDs

Zhigang Wang https://orcid.org/0000-0002-7123-335X

References
[1] Kwon O, Leeb K and Kim G 2010 New trends in advanced high strength steel developments for automotive application Mater. Sci.
Forum 638–642 136
[2] Wang Z-G, Zhao A-M and Zhao Z-Z 2012 Microstructures and mechanical properties of C–Mn–Cr–Nb and C–Mn–Si–Nb ultra-high
strength dual-phase steels International Journal of Minerals Metallurgy & Materials 19 915–22
[3] Yoshida H et al 2007 Effect of niobium addition on the texture formation of high strength cold-rolled low carbon steel sheets Mater. Sci.
Forum 558–559 425–30
[4] Naderi M et al 2011 Analysis of microstructure and mechanical properties of different high strength carbon steels after hot stamping
Journal of Materials Processing Tech 211 1117–25
[5] Yoshinaga N, Kestens L and De Cooman B C 2005 α→γ→α transformation texture formation at cold-rolled ultra low carbon steel
sufaces Mater. Sci. Forum 495–497 1267–72
[6] Symbols–Recrystallization and Related Annealing Phenomena (Second Edition). Recrystallization & Related Annealing Phenomena:
xxi–xxii
[7] Kang J Y, Lee H C and Han S H 2011 Effect of Al and Mo on the textures and microstructures of dual phase steels Materials Science &
Engineering A 530 (none) 183–90
[8] Bae S H and Lee H W 2013 Effect of Mo contents on corrosion behaviors of welded duplex stainless steel Metals & Materials
International 19 563–9
[9] Wang Z G, Zhao A M, Zhao Z Z et al 2013 Precipitation behavior and textural evolution of cold-rolled high strength deep drawing dual-
phase steels Journal of Iron and Steel Research, International 20 61–8
[10] Yazawa Y et al 2003 Development of ferritic stainless steel sheets with excellent deep drawability by {111} recrystallization texture
control Jsae Review 24 483–8
[11] Inagaki H 1994 Fundamental aspect of texture formation in low carbon steel ISIJ Int. 34 313–21
[12] Saha R and Ray R K 2007 Texture and grain boundary character of a Ti J. Mater. Sci. 42 9548–52
[13] Meng Q et al 2009 Effect of water quenching process on microstructure and tensile properties of low alloy cold rolled dual-phase steel
Mater. Des. 30 2379–85
[14] Hutchinson B et al 1998 Texture in hot rolled austenite and resulting transformation products Materials Science and Engineering A 257
9–17
[15] Park Y B, Lee D N and Gottstein G 1997 Evolution of recrystallization textures from cold rolling textures in interstitial free steel Mater.
Sci. Technol. 13 289–98
[16] Daniel D and Jonas J J 1990 Measurement and prediction of plastic anisotropy in deep-drawing steels Metall. Trans. A 21 331–43
[17] Sánchez-Araiza M et al 2006 Texture evolution during the recrystallization of a warm-rolled low-carbon steel Acta Mater. 54 3085–93
[18] Daniel D, Savoie J, Jonas J J et al 1993 Textures induced by tension and deep drawing in low carbon and extra low carbon steel sheets
Acta Metallurgica Et Materialia 41 (6) 1905-1920
[19] Zhuang D D et al 2017 Microstructure and texture evolution during recrystallization of low-carbon steel sheets J. Iron Steel Res. 24
84-90
[20] Zhang L et al 2016 Transformation of {100 }texture induced by surface effect in ultra-low carbon electrical steel J. Mater. Sci. 51
8087–97

You might also like