You are on page 1of 9

Materials Science & Engineering A 739 (2019) 17–25

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

The influence of microstructural characteristics on yield point elongation T


phenomenon in Fe-0.2C-11Mn-2Al steel
Z.H. Caia, , S.Y. Jinga, H.Y. Lib, , K.M. Zhanga, R.D.K. Misrac, H. Dinga, Z.Y. Tanga
⁎ ⁎

a
School of Materials Science and Engineering, Northeastern University, Shenyang 110819, China
b
School of Materials Science and Engineering, Taiyuan University of Science and Technology, Taiyuan 030024, China
c
Laboratory for Excellence in Advanced Steel Research, Department of Metallurgical, Materials and Biomedical Engineering, University of Texas at El Paso, El Paso, TX
79968, USA

ARTICLE INFO ABSTRACT

Keywords: We elucidate here the underlying reasons for the yield point elongation (YPE) phenomenon in medium-Mn steel.
Medium Mn steel The cold-rolled Fe-0.2C-11Mn-2Al steel was subjected to controlled annealing temperature and soaking time
Yield point elongation such that YPE of different lengths was obtained. It was observed that the ultra-fine grained or globular micro-
Austenite stability structure is not the critical factor responsible for YPE. YPE is a consequence of close competition between the
Dislocation density
work hardening (via martensitic transformation) and softening (by stress relaxation and transfer) induced by the
TRIP effect
TRIP effect. The ferrite grains with γ fiber have the potential to have high dislocation storage ability, with
consequent increase in dislocation density and work hardening exponent, leading to decrease of YPE.

1. Introduction spreads for a certain strain, is referred as yield point elongation (YPE).
The classical explanation for YPE phenomenon is static strain aging,
Medium Mn steels continue to receive significant attention because which means that dislocations can be pinned by clusters of interstitial
of excellent mechanical properties and potential for automotive struc- atoms, or Cottrell atmospheres [15]. A different mechanism that pro-
tural components. In the past decade, numerous studies have focused motes Lüders band formation is ultra-fine grained (UFG) microstructure
on obtaining large fraction of austenite with proper stability by opti- (grain size < 1 µm) [16]. With decreasing grain size, an increasing
mizing chemical composition and adjusting heat treatment [1–4]. It number of dislocations are trapped at grain boundaries and become
was reported that ductility could be increased simultaneously with immobile [17–19], which has a similar effect as the trapping of dis-
strength by transformation-induced plasticity (TRIP) mechanism, con- locations by interstitials. Medium-Mn steels commonly exhibit large
tributing to superior tensile property > 30 GPa⋅%, when austenite was YPE, some of them easily exceeding 10% [14,20,21]. It was reported
present with ferrite or martensite in medium Mn steels. Especially, that both low IA temperature and a globular microstructure supported
ultra-high strength (> 1.5 GPa) combined with high ductility (> 20%) the formation of large YPE. Moreover, it was recommended to provide a
were achieved in (austenite + martensite) duplex microstructure [5]. martensitic microstructure prior to intercritical annealing in order to
In order to achieve good strength-ductility balance, austenite should limit YPE [20].
gradually transform over a wide range of strain [6]. It has been proven Although there are some studies on the YPE, the micromechanical
that Lüders bands directly results from the mechanical stability of origin of YPE is still unknown. The study described here concentrates
austenite [7]. Austenite stability is known to depend on chemical on the influence of microstructural characteristics on YPE. A detailed
composition [8], grain size [4,9], morphology [10], and crystal- characterization of microstructural evolution, including austenite
lographic orientation of austenite [11], etc, among which chemical (fraction, size, stability) and ferrite (size, texture) characteristics, and
composition and grain size are the primary parameters [4,11]. Auste- the resulting mechanical properties including YPE and work hardening
nite stability can be varied by controlling the intercritical annealing behavior are presented.
(IA) temperature and the annealing time [2,12]. Moreover, Lüders
bands were influenced by IA temperature in medium Mn steels [13,14]. 2. Experimental
The phenomenon of Lüders bands characterized by a stress plateau
accompanied by serrations, which initiates at the onset of yielding and The experimental steel had a nominal composition (wt%) of Fe-


Corresponding authors.
E-mail addresses: tsaizhihui@163.com (Z.H. Cai), lhyemail@163.com (H.Y. Li).

https://doi.org/10.1016/j.msea.2018.09.114
Received 13 September 2018; Received in revised form 27 September 2018; Accepted 29 September 2018
Available online 09 October 2018
0921-5093/ © 2018 Elsevier B.V. All rights reserved.
Z.H. Cai et al. Materials Science & Engineering A 739 (2019) 17–25

Table 1 3. Results and discussion


Chemical composition (wt%) of steel and critical temperatures (°C).
Mn Al C Fe Ac1 Ac3 3.1. Microstructure and mechanical properties

11.20 1.95 0.22 Bal. 630 730 SEM micrographs of cold-rolled sheets annealed at 600 °C, 650 °C,
700 °C,and 750 °C for 5 min, respectively, and then immediately quen-
ched in water, are presented in Fig. 1. As shown in Fig. 1a, the mi-
0.2C-11Mn-2Al, and the actual chemical composition is listed in crostructural constituents consisted of ferrite and austenite. The mi-
Table 1. The ingot was heated at 1200 °C for 2 h, hot forged to rods with crostructure can be described into two types: strip-like and equiaxed /
a section size of 100 mm × 30 mm, followed by air cooled to room granular. The strip structure developed during cold rolling of partially
temperature. Subsequently, the rods were soaked at 1200 °C for 2 h, hot recrystallized, evolved into equiaxed structure on annealing at 600 °C.
rolled to 4 mm in thickness and finally air cooled to room temperature. For clarity, a mixture of austenite and ferrite strips are identified by a
The as-hot rolled sheets were then cold-rolled to 1 mm thickness. rectangle in Fig. 1a, and a high-magnification micrograph of this region
In order to establish an appropriate heat treatment schedule, the is illustrated in Fig. 1e. It is distinct that austenite strips were present
critical temperatures Ac1 and Ac3 of the experimental steel were ob- between the adjacent ferrite strips. As shown in Fig. 1b and c, the
tained by dilatometry and are listed in Table 1. For obtaining different austenite and ferrite strips gradually developed into equiaxed micro-
microstructural constituents and grain size, the as-cold rolled sheets structure with increase in annealing temperature. A small number of
were annealed in the temperature range of 600–800 °C for 5 min, fol- strip structure continue to be retained on annealing at 650 °C. However,
lowed by water quenching. Additionally, to obtain similar austenite it completely disappeared after annealing at 700 °C. When the samples
fraction but different austenite stability, the as-cold rolled sheets were were annealed at 750 °C, as shown in Fig. 1d, the microstructure
annealed at 650 °C for different periods (2 min~48 h), followed by comprised of martensite and austenite.
water quenching. The measured austenite fraction as a function of annealing tem-
Tensile tests were conducted on specimens of 12.5 mm width and perature or soaking time was summarized in Fig. 2. As illustrated in
gauge length of 25 mm, using a universal testing machine (SANSCMT Fig. 2a, the austenite fraction increases from 55% to 89% in the tem-
5000) at a constant crosshead speed of 3 mm·min−1 at room tempera- perature range of 600–700 °C, followed by drastic decrease to 46%,
ture. The samples were etched with 25% sodium bisulfite aqueous so- when quenching was carried out at 750 °C because of martensitic
lution. Microstructural examination was carried out using scanning transformation, consistent with microstructure illustrated in Fig. 1d.
electron microscope (SEM) equipped with electron backscatter dif- Therefore, the austenite fraction is sensitive to annealing temperature.
fraction (EBSD) and transmission electron microscope (TEM). The vo- Furthermore, it is confirmed that the austenite fraction remains nearly
lume fraction of austenite was determined by X-ray diffraction (XRD) constant when the samples were annealed at 650 °C for different per-
with CuKα radiation using direct comparison method [22], involving iods (as evidenced in Fig. 2b).
the use of integrated intensities of (200)α and (211)α peaks and those of Fig. 3 shows the engineering stress-strain plots of the annealed
(200)γ, (220)γ and (311)γ peaks. The volume fraction of austenite VA samples. As shown in Fig. 3a, it is noticed that YPE only appears in the
was calculated using equation [23]: sample annealed at 650 °C for 5 min (henceforth nominated as
650–5 min sample). Moreover, as demonstrated in Fig. 3b, it is clear
VA = 1.4I /(I + 1.4I ) (1)
that the length of YPE decreases with extended soaking time at 650 °C.
To interpret this phenomenon, the interplay between microstructural
where Iγ is the integrated intensity of austenite and Iα is the integrated
characteristics and plastic deformation behavior including YPE and
intensity of phases with body-centered cubic structure.
TRIP effect should be elucidated.

Fig. 1. SEM micrographs of the cold-rolled samples quenched from (a) 600 °C, (b) 650 °C, (c) 700 °C and (d) 750 °C, respectively and (e) high magnification of
austenite and ferrite strips. The bright phase is austenite, and the dark one is ferrite. (A: austenite, F: ferrite, M: martensite).

18
Z.H. Cai et al. Materials Science & Engineering A 739 (2019) 17–25

Fig. 2. The measured austenite fractions of the cold-rolled samples (a) annealed at different temperatures, (b) annealed at 650 °C for different soaking time.

Fig. 3. The engineering stress-strain plots of the cold-rolled samples (a) annealed at different temperatures for 5 min, (b) annealed at 650 °C for different soaking
time.

Fig. 4. EBSD phase maps of the annealed samples. The phases colored in red and blue correspond to ferrite and austenite, respectively. The white lines are low-angle
boundaries with misorientation angles between 5° and 15°. The black lines are high-angle boundaries with misorientation angles of over 15°. (a) 650–5 min sample,
(b) 650–1 h sample, (c) 650–12 h sample, (d) 650–48 h sample, (e) 700–5 min sample.

19
Z.H. Cai et al. Materials Science & Engineering A 739 (2019) 17–25

Table 2 that, as shown in Fig. 3, YPE only appeared in the samples annealed at
The average grain size (μm) of austenite and ferrite of the annealed samples. 650 °C (referred as 650 samples). Therefore, the UFG or globular mi-
Annealed samples Austenite Ferrite crostructure is not the critical factor causing YPE in our case.
In order to decrease or eliminate the YPE, Li et al. [25] suggested
650–5 min 0.34 0.28 increasing austenite stability and dislocation density in ferrite. As
700–5 min 0.54 0.50
shown in Fig. 2, it is inferred that austenite stability decreases with
650–1 h 0.54 0.38
650–12 h 0.69 0.54
increase in annealing temperature or soaking time. Furthermore, the
650–48 h 0.85 0.66 average dislocation density in ferrite and austenite were determined by
XRD experiments [26,27]. The average dislocation density was eval-
uated by selecting the diffraction peaks of (220)γ and (211)α. The
3.2. The influences in YPE density of dislocations was estimated using Eq. (2) proposed by Dunn
[28,29].
Fig. 4 demonstrates EBSD phase maps of the annealed samples. The
phases colored in red and blue correspond to ferrite and austenite, re- = 2 /(4.35 × b 2) (2)
spectively. The white lines are low-angle boundaries with misorienta-
where ρ and β is dislocation density and full-width at half maximum
tion angles between 5° and 15°. The black lines are high-angle bound-
(FWHM) of diffraction peaks, respectively as determined by X-ray dif-
aries with misorientation angles of over 15°. The average grain size of
fraction profile, b is Burgers vector and the value is 2.48 × 10−8 cm.
austenite and ferrite determined via Channel 5 software is summarized
FWHM is primarily affected by crystallite size and microstrain. In our
in Table 2. Austenite and ferrite grain size were in the range of
present case, the increase of dislocation density of samples which were
0.1–0.9 µm, and increased with increasing annealing temperature and
subjected to interrupted tensile tests was mainly contributed from the
retention period. It is reported that UFG microstructure is likely to
increasing strain. Therefore, the variation of FWHM is primarily af-
trigger YPE, and YPE will be more distinct with decreasing grain size
fected by microstrain. In order to better understand the evolution of
[17,24]. Steineder et al. [20] reported that both low annealing tem-
dislocation density, the XRD patterns corresponding to increasing strain
perature and globular microstructure remarkably supported the for-
are shown in Fig. 5. Table 3 summarizes the estimated dislocation
mation of large YPE. As shown in Fig. 1 and Table 2, all the annealed
density of ferrite in the annealed samples. It is obvious that the dis-
samples had UFG and globular microstructure. However, it is curious
location density of ferrite was unchanged or decreased slightly with

Fig. 5. XRD patterns of the annealed samples subjected to interrupted tensile tests. (a) 650–5 min sample, (b) 650–1 h sample, (c) 650–48 h sample, (d) 700–5 min
sample.

20
Z.H. Cai et al. Materials Science & Engineering A 739 (2019) 17–25

Table 3 sample having a larger ferrite grain size than 650–12 h sample showed
Dislocation density (1014) of ferrite in the annealed samples. continuous yielding (without YPE). Therefore, the above interpretation
Annealed samples Dislocation density on the relationship between ferrite and YPE is not applicable to our
results. There are some other important factors contributed to YPE.
600–5 min 2.22 As shown in Fig. 3b, 650–2 min sample was characterized by long
650–5 min 2.49
YPE, UFG structure and high austenite stability. A smaller grain size
700–5 min 2.21
650–1 h 2.48
reduces the probability of dislocation trapping in the grain interior and
650–12 h 2.27 austenite has higher stability, thereby reducing the work hardening
650–48 h 2.23 rate. It is conceivable that when the austenite grain size is large enough,
and the hardening contributed by TRIP effect overcomes the loss of UFG
structure, YPE will disappear. Combined with XRD results (Fig. 2),
increase in annealing temperature or soaking time. Based on the above stress-strain plots (Fig. 3) and grain size (Table 2), it is assumed that
results, austenite stability, ferrite dislocation density and YPE of 650 YPE is a consequence of close competition between work hardening (via
samples decreased with increase in soaking time, which is contrary to Li martensitic transformation) and softening (by stress relaxation and
et al.′s results [25]. Thus, the question arises as to what is the under- transfer) induced by the TRIP effect. In an attempt to clarify the hy-
lying reason for YPE. pothesis, concerning the impact of grain size on YPE, 650–5 min,
650–1 h, 650–48 h and 700–5 min samples were subjected to inter-
3.3. The underlying reason for YPE rupted tensile tests.
Fig. 6 illustrates the volume fraction of austenite as a function of
It has been observed that the YPE and martensitic transformation engineering strain in the aforementioned four samples. For 650–5 min
takes places simultaneously during tensile test [30]. Nevertheless, it and 650–1 h samples (Fig. 6a and b), similarly, the austenite fraction
cannot be concluded that the YPE is induced by martensitic transfor- decreases rapidly on straining to 5%, and then remains at a certain level
mation of austenite. It has been found that the YPE also takes place until the end of YPE, and finally decreases with increase in strain.
when there is no martensitic transformation [31]. Although He et al. Martensitic transformation was accompanied by volume expansion,
[21] suggested that the large ferrite grains were responsible for the therefore, the nearby retained austenite and ferrite was inevitably
initiation of YPE, which was on the basis that the large ferrite grains squeezed by the transformed martensite, leading to additional dis-
had a lower resistance to dislocation plasticity as compared to the location propagation. As shown in Fig. 7a and b, the dislocation density
austenite phase and the small ferrite grains. However, Han et al. [32] in austenite first increases, and then remains nearly invariant, finally
proposed that the coarsening of globular ferrite grains increased the increases with increasing strain, corresponding to the change in aus-
propagation velocity of Lüders bands, resulting in the reduction in the tenite fraction. In contrast, the dislocation density in BCC structure
YPE. In our present case, as evidenced in Table 2 and Fig. 3, 650–12 h involving ferrite and martensite increases slightly throughout straining.
sample characterized by a long YPE, instead, had a larger ferrite grain Therefore, for 650–5 min and 650–1 h samples, a majority of the local
size than 700–5 min sample which had no YPE. In addition, 650–48 h

Fig. 6. The engineering stress-strain plots and the variation of austenite fraction as a function of strain. (a) 650–5 min sample, (b) 650–1 h sample, (c) 650–48 h
sample, (d) 700–5 min sample.

21
Z.H. Cai et al. Materials Science & Engineering A 739 (2019) 17–25

Fig. 7. The variation of dislocation density as a function of stain. (a) 650–5 min sample, (b) 650–1 h sample, (c) 650–48 h sample, (d) 700–5 min sample.

stress was transferred to austenite rather than ferrite. In contrast, as 3.4. Work hardening behavior and texture evolution
shown in Fig. 6c and d, for 650–48 h and 700–5 min samples, the
austenite fraction decreased quickly with increase in strain, especially The relationship between variation in dislocation density and YPE
during the initial stage of straining (< 5%), indicating a significant can be further reinforced through the study of work hardening behavior
TRIP effect, contributing to the disappearance of YPE. Correspondingly, and evolution of texture in ferrite. Fig. 8 shows the work hardening
the dislocation density both in austenite and BCC structure increase exponent of 650 samples and 700–5 min sample. It is distinct that the
quickly with increasing strain (as illustrated in Fig. 7c and d). work hardening exponent of 650–5 min and 650–1 h samples remained
Comparing the variation in dislocation density and austenite stabi- at a low value until the finish of YPE, corresponding to low dislocation
lity among the three 650 samples, it is distinct that the dislocation density in BCC structure, but high dislocation density in austenite. In
density of austenite decreases with decrease in austenite stability, contrast, the work hardening exponent of 650–48 h and 700–5 min
corresponding to a reduction in YPE. Because of higher austenite sta- samples increased rapidly during the initial stage of deformation, cor-
bility in 650–5 min and 650–1 h samples, more dislocations were re- responding to simultaneous increase in dislocations in both BCC
quired to stimulate martensitic transformation. Moreover, at similar structure and austenite. Moreover, it is noted that, at similar strain, the
strain (< 10%), the fraction of transformed martensite in 650–5 min dislocation density of austenite in 650–48 h and 700–5 min samples was
and 650–1 h samples was less. Thus, the local stress can be effectively much lower compared to 650–5 min and 650–1 h samples. Therefore, it
transferred to the nearby austenite and ferrite grains, leading to local is proven that the increase of dislocation density in austenite con-
softening and the consequent existence of YPE. Therefore, in spite of tributed little to work hardening which was primarily contributed from
higher dislocation density of austenite in the two samples, the work BCC structure. Austenite itself could not provide effective hardening,
hardening ability is still low, which was further confirmed in Fig. 8a but the accumulation of dislocations in austenite promotes more mar-
and b. In contrast, due to the lower austenite stability, more martensite tensitic transformation with consequent enhancement in hardening. For
nucleated during the initial stage of straining in 650–48 h and another, as martensite was surrounded by the soft austenite and ferrite,
700–5 min samples, which was resulted in the local stress concentration the increase in dislocation density in the BCC structure was primarily
and rapid work hardening (as evidenced in Fig. 8c and d), leading to a from ferrite, especially during the initial stage of deformation. Con-
continuous yielding (without YPE). sidering Figs. 6–8, it is deduced that when dislocations in austenite

22
Z.H. Cai et al. Materials Science & Engineering A 739 (2019) 17–25

Fig. 8. Work hardening exponent of (a) 650–5 min sample, (b) 650–1 h sample, (c) 650–48 h sample, (d) 700–5 min sample.

accumulate to a critical extent, the hardening induced by TRIP effect will deform preferentially and play a dominant role during the initial
overcomes local softening, which results in the finish of YPE. On the stage of deformation, resulting in low dislocation density and work
basis of aforementioned discussion, it is confirmed that YPE is a con- hardening exponent. Therefore, the ferrite grains with γ fiber have high
sequence of equally close competition between work hardening (via dislocation storage ability, which is also responsible for the increase in
martensitic transformation) and softening (by stress relaxation and dislocation density and work hardening exponent, leading to decrease
transfer) induced by the TRIP effect. of YPE.
The difference in the evolution of ferrite dislocation density illu-
strated in Fig. 7 could be clarified by studying the texture of ferrite. 4. Conclusions
Fig. 9 shows contour plots of the orientation distribution function
(ODF) for ϕ2 = 45° and texture intensity of γ fiber. For 650–1 h sample, (1) The YPE only appeared in samples annealed at 650 °C, and the
the ferrite grains are characterized by a strong α fiber (∥ < 110 >), and length of YPE decreased with increase in soaking time. The UFG or
γ fiber (∥ < 111 >) with decreasing intensity. Comparing Fig. 9a with globular microstructure was not the critical factor responsible for
c, it is distinct that 650–48 h sample had a similar intensity of α fiber to YPE in 650 samples. YPE is a consequence of close competition
650–1 h sample. However, the intensity of γ fiber was stronger in between the work hardening (via martensitic transformation) and
650–48 h sample. In contrast, in 700–5 min sample, the ferrite grains softening (by stress relaxation and transfer) induced by the TRIP
had a strong γ fiber, and α fiber trends to be weaker or nearly absent. It effect.
was reported that the stored energy of α fiber was lower, thus the (2) The length of YPE decreased with decrease in austenite stability.
crystal defects such as dislocations, dislocation cell walls were difficult With increase in austenite stability, more dislocations were re-
to be stored, which implied that the work hardening ability will be low quired to stimulate martensitic transformation. Austenite itself
[33,34]. In comparison, the stored energy of γ fiber was higher, dis- could not provide effective hardening, but the accumulation of
locations pile-up, tangle and cross-slip easily, resulting in high dis- dislocations in austenite promotes more martensitic transformation
location density and work hardening ability. Moreover, α fiber has a with consequent enhancement in hardening.
lower Taylor factor than γ fiber [35], indicating that ferrite with α fiber (3) The formation of YPE was not affected by ferrite grain size but

23
Z.H. Cai et al. Materials Science & Engineering A 739 (2019) 17–25

Fig. 9. The contour plots of the orientation distribution function (ODF) for ϕ2 = 45° and texture intensity of γ fiber. (a, b) 650–1 h sample, (c, d) 650–48 h sample, (e,
f) 700–5 min sample.

texture. The ferrite grains with γ fiber have high dislocation storage (2015081011). R.D.K. Misra gratefully acknowledges support from the
ability, which was also responsible for the increase in dislocation National Science Foundation, USA through grant # DMR 1602080.
density and work hardening exponent, leading to decrease of YPE. R.D.K. Misra also acknowledges collaboration with Northeastern
University as an Honorary Professor.
Acknowledgements
References
The authors acknowledge support from the National Natural Science
[1] H.W. Luo, J. Shi, C. Wang, et al., Acta Mater. 59 (2011) 4002–4014.
Foundation of China (51501035 and 51574077), the Fundamental
[2] M.J. Merwin, Mater. Sci. Forum 539–543 (2007) 4327–4332.
Research Funds for the Central Universities (N170204017) and [3] J. Shi, X.J. Sun, M.Q. Wang, et al., Scr. Mater. 63 (2010) 815–818.
Shanxi International Science and Technology Cooperation Program [4] S. Lee, S.J. Lee, B.C. De Cooman, Scr. Mater. 65 (2011) 225–228.

24
Z.H. Cai et al. Materials Science & Engineering A 739 (2019) 17–25

[5] H. Lee, M.C. Jo, S.S. Sohn, et al., Acta Mater. 147 (2018) 247–260. [20] K. Steineder, D. Krizan, R. Schneider, et al., Acta Mater. 139 (2017) 39–50.
[6] M.M. Wang, C.C. Tasan, D. Ponge, et al., Acta Mater. 111 (2016) 262–272. [21] B.B. He, Z.Y. Liang, M.X. Huang, Scr. Mater. 150 (2018) 134–138.
[7] J.H. Ryu, J.I. Kim, H.S. Kim, et al., Scr. Mater. 68 (2013) 933–936. [22] A.K. Srivastava, D. Bhattacharjee, G. Jha, (et a), Mater. Sci. Eng. A 445–446 (2007)
[8] J.G. Speer, E. De Moor, K.O. Findley, et al., Metall. Mater. Trans. A. 42A (2011) 549–557.
3591–3601. [23] A. Grajcar, H. Krzton, J. Achiev. Mater. Manuf. Eng. 35 (2009) 169–176.
[9] Z.H. Cai, H. Ding, R.D.K. Misra, et al., Acta Mater. 84 (2015) 229–236. [24] A. Vinogradov, Adv. Eng. Mater. 17 (2015) 1710–1722.
[10] X.C. Xiong, B. Chen, M.X. Huang, et al., Scr. Mater. 68 (2013) 321–324. [25] Z.C. Li, R.D.K. Misra, H. Ding, et al., Mater. Sci. Eng. A 712 (2018) 206–213.
[11] G.K. Tirumalasetty, M.A. Van Huis, C. Kwakernaak, et al., Acta Mater. 60 (2012) [26] Z.Y. Liang, Y.Z. Li, M.X. Huang, Scr. Mater. 112 (2016) 28–31.
1311–1321. [27] W. Li, W.Z. Xu, X.D. Wang, J. Alloy Compd. 474 (2009) 546–550.
[12] H.F. Xu, J. Zhao, W.Q. Cao, et al., Mater. Sci. Eng. A 532 (2012) 435–442. [28] C.G. Dunn, E.F. Koch, Acta Metall. 5 (1957) 548–554.
[13] M. Callahan, O. Hubert, F. Hild, et al., Mater. Sci. Eng. A 704 (2017) 391–400. [29] P. Gay, P.B. Hirsch, A. Kelly, Acta Metall. 1 (1953) 315–319.
[14] F. Yang, H.W. Luo, E.X. Pu, et al., Int. J. Plast. 103 (2018) 188–202. [30] X.G. Wang, L. Wang, M.X. Huang, Acta Mater. 124 (2017) 17–29.
[15] A.H. Cottrell, B.A. Bilby, Dislocation theory of yielding and strain ageing of iron, [31] X.G. Wang, M.X. Huang, J. Iron Steel Res. Int. 24 (2017) 1073–1077.
Proc. Phys. Soc. A 62 (1949) 49e62. [32] J. Han, S.H. Kang, S.J. Lee, et al., J. Alloy Compd. 681 (2016) 580–588.
[16] Y. Tomota, A. Narui, N. Tsuchida, ISIJ Int. 48 (2008) 1107–1113. [33] R.L. Every, M. Hatherly, Texture 1 (1974) 183–194.
[17] C.E. Carlton, P.J. Ferreira, Acta Mater. 55 (2007) 3749–3756. [34] I.L. Dillamore, C.J.E. Smith, T.W. Watson, Mater. Sci. J. 1 (1967) 49–54.
[18] A.A. Nazarov, A.E. Romanov, R.Z. Valiev, Nanostruct. Mater. 4 (1994) 93–101. [35] N. Rajmohan, Y. Hayakawa, J.A. Szpunar, et al., Acta Mater. 45 (1997) 2485–2494.
[19] K.T. Park, D.H. Shin, Metall. Mater. Trans. A 33 (2002) 705–707.

25

You might also like