Author’s Accepted Manuscript
Effects of intercritical annealing process on
microstructures and tensile properties of cold-rolled
7Mn steel
Feng Yang, Haiwen Luo, Chundong Hu, Enxiang
Pu, Han Dong
www.elsevier.com/locate/msea
PII: S0921-5093(16)31624-0
DOI: http://dx.doi.org/10.1016/j.msea.2016.12.119
Reference: MSA34550
To appear in: Materials Science & Engineering A
Received date: 29 June 2016
Revised date: 26 November 2016
Accepted date: 30 December 2016
Cite this article as: Feng Yang, Haiwen Luo, Chundong Hu, Enxiang Pu and Han
Dong, Effects of intercritical annealing process on microstructures and tensile
properties of cold-rolled 7Mn steel, Materials Science & Engineering A,
http://dx.doi.org/10.1016/j.msea.2016.12.119
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Effects of intercritical annealing process on microstructures and
tensile properties of cold-rolled 7Mn steel
Feng Yanga, Haiwen Luob*, Chundong Hua, Enxiang Pua, Han Donga
a
Central Iron & Steel Research Institute, Xue Yuan Nan Lu 76, Beijing 100081, China
b
School of Metallurgical and Ecological Engineering, University of Science and Technology
Beijing, Xue Yuan Lu 30, Beijing 100083, China
*
Correspondence author: Prof. Haiwen Luo, luohaiwen@ustb.edu.cn
Abstract:
Influences of intercritical annealing temperature and duration on the microstructures
and tensile properties of a newly designed medium Mn steel, Fe-7 wt.% Mn-0.3 wt.%
C-2 wt.% Al, have been studied and discussed in this paper. Two types of cold-rolled
ferritic microstructures, i.e. cell-like and lath-like, could transform to granular and
lamellar austenite grains respectively during the subsequent intercritical annealing
(IA). Both the IA temperature and duration strongly affect the fraction of retained
austenite that have transformed during tensile deformation and the resultant tensile
properties. In comparison with a Al-free 7Mn steel, it is found that the addition of Al
has led to the reduced fraction but enhanced stability of RA grains; thus, they may
transform gradually during deformation, which can make the maximum contribution
to the sustainable work hardening. The developed steel exhibits the product of
strength and plasticity up to 66GPa·%, which is much better than the Al-free 7Mn
steel and almost one of the best tensile properties among the existing medium-Mn
steels but with relatively lean alloying.
Key Words: medium Mn steel; TRIP effect; microstructure evolution; mechanical
properties
1. Introduction
Due to the increasing pressure on the lower emission of vehicles, the automotive
industries are strongly motivated to design automobiles using higher and higher
strength steels but still maintaining good formability. To meet this requirement, the
3rd generation of automobile steel, in which the product of ultimate tensile strength
(UTS) and total elongation (TE) is more than 30GPa·%, has been extensively studied
in the past ten years [1-8]. Although many advanced steels have been developed for
this goal, the medium-Mn steels, whose Mn content is usually in the range of 5-12
wt.%, are considered as the most promising candidate for the 3rd generation
automotive steels due to their excellent combination of strength and ductility.
In 1972, Miller [9] firstly developed a 0.11C–5.7Mn (all alloy compositions are
in wt%, unless otherwise mentioned) steel with 29% fraction of retained austenite
(RA), which had a tensile strength of 878MPa and a total elongation of 34%.
Motivated by this success with medium-Mn alloy, Mervin [3, 4], Cao [5, 6] and Park
[7, 8] have all developed different medium-Mn steels that have the product of UTS
and TE higher than 30GPa% and RA fractions more than 20%. Since the addition of
Al into medium Mn steel can enhance Ae3 temperature and suppress excessive
transformation of austenite, the intercritical annealing has to be performed at higher
temperature and more solute atoms like Mn and C can be partitioned into austenite [7];
therefore, some medium Mn steels alloyed by Al have been developed for improved
tensile properties. For examples, Lee et al. [10-12] annealed the cold-rolled steel with
the composition of 10Mn-0.3C-3Al-2Si at 800℃, leading to the product of UTS and
TE to nearly 72GPa·%; Cai et al. [13] designed a cold rolled 11Mn-0.18C-3.8Al steel,
which exhibited the product of UTS and TE up to 65 GPa·% after annealing at 770℃.
Although the mechanical properties of these recently developed TRIP/TWIP steels are
comparable to the traditional TWIP steels [14, 15], they are not cost effective enough
in comparison with the TWIP steels because they still contain relatively high contents
of alloying elements like Mn and Al, which brings a lot of difficulties in industrial
production besides the high alloying cost. For examples, the higher Mn content of
medium-Mn steel not only requires more heat input to melt ferromanganese alloys but
also results in decreased thermal conductivity of continuous casting slabs, the latter
may cause internal cracks between columnar grains in slabs due to the larger thermal
stress gradient during cooling. Therefore, it is important to greatly improve the tensile
properties of medium Mn steel using the least alloying, i.e. the contents of Mn and Al
should be as low as possible. In this paper, we have presented a new medium Mn steel
that has a relatively lean composition, whose nominal composition is 7Mn-0.3C-2Al,
but exhibits excellent tensile properties.
2. Experimental procedures
The studied 7Mn steel, whose chemical composition is 7.2Mn-0.22C-2.4Al, was
melted in a vacuum induction furnace and then cast into a ingot. The ingot was
homogenized at 1200℃ for 2 hours and forged between 1200℃ and 850℃ into slabs
with the dimension of 40mm×120mm×200mm. The forged slabs were reheated to
1180℃ for 1h and hot-rolled to 3.8mm-thick plates between 1150 and 850℃. The
hot-rolled plates were annealed at 700℃ for 30 minutes and then cold rolled to about
1.1mm thickness with the reduction of 70%. Next, the cold-rolled sheets were
intercritically annealed at the temperatures of 680℃, 700℃ and 720℃ for various
durations of 0.5-5 hours respectively, finally cooled to room temperature in air.
The microstructures in both cold rolled and intercritically annealed steels were
examined by scanning electron microscope (SEM, HITACHI S-4300) and
transmission electron microscope (TEM, HITACHI H-800). The samples for SEM
were etched in 4% nital after a standard polishing. Thin foils for TEM were
mechanically ground to thickness of about 40μm and then electro-polished in a
twin-jet machine with a solution of 6% perchloric acid and 94% alcohol at about -20℃.
X-ray diffraction (XRD) measurements using Cu-Kα radiation were performed after
mechanical and electro polishing. The calculations of the volume fraction of retained
austenite (RA) were based on the integrated intensities of the (200)α, (211)α, (200)γ,
(220)γ, and (311)γ diffraction peaks using the following equation 1 [16]:
Vγ=1.4Iγ/( Iα+1.4Iγ)………..………………….……………………………………...(1)
where Iγ is the mean integrated intensity of the three austenite diffraction peaks,
(200)γ, (220)γ, and (311)γ; Iα is the mean integrated intensity of two ferrite diffraction
peaks, (200)α and (211)α. The tensile samples were machined from annealed strips
along the rolling direction with gauge length and width of 50 mm and 12.5 mm,
respectively. Tensile tests were carried out at ambient temperature using a universal
testing machine (WE300) at a constant cross-head speed of 2mm/min. The Mn and Al
contents in austenite were measured by Scanning Transmission Electronic Microscope
(STEM, Tecnai G2F20) equipped with Energy Disperse Spectrum. The average carbon
content of retained austenite grains was estimated using the following equation [17]:
aγ(Å)=3.578+0.033XC+0.00095XMn+0.0056XAl ………...… ……………………....(2)
where aγ is the lattice parameter of austenite, which can be calculated from XRD
results, XC, XMn and XAl are the mass concentrations of C, Mn and Al in austenite,
respectively.
3. Results
3.1 Microstructural evolution
The microstructures of both cold-rolled and annealed samples are given in Fig. 1.
Fig. 1a shows the severely deformed microstructure after cold rolling, in which grains
were elongated significantly along the rolling direction, leading to a band-like
morphology. After the intercritical annealing (IA) at 680℃ for 1hour, many elongated
grains transformed to polygonal grains while some still remained band-like (Fig. 1b).
After the IA at 700℃ for 1hour, more elongated grains transformed to polygonal ones
while only a small amount of grains were still band-like (Fig.1c). A further increase of
IA temperature to 720℃ removed almost all the cold-rolled microstructures, leading to
all the grains formed granularly with a larger size (Fig.1d). The prolonged duration of
IA has the similar influence on the microstructures to the increase of IA temperature.
It can be seen from Fig. 2 that both the fraction and size of granular grains increase
significantly when the IA period at 700℃ increases from 30 minutes to 5 hours.
TEM photos in Fig. 3 reveal two types of cold-rolled dislocation microstructures:
dislocation cells and lath-like, as shown in Fig. 3a and 3b respectively. They are very
similar to the microstructures in the cold-deformed stainless steels [18, 19]. During
the subsequent IA, the equiaxed and lath-shaped austenite and ferrite grains, which
could be seen in Fig. 3c and 3d, should transform from the cell-like and lath-like
microstructures during the process of IA. With the increase of temperature and
duration of IA, the lath-like ferrite and austenite grains would gradually transform to
equiaxed type and the grain size would enlarge at the same time.
The XRD spectra of specimens after both various IA process and the tensile
fracture are given in Figs. 4a and 4b respectively. The RA fractions, Vγ, were
calculated using Equation 1 from Figs. 4a and 4b and shown in Fig. 4c. It is noted that
the measured maximum RA fraction is about 35%; in contrast, the maximum RA
fraction in the Al-free 7Mn-0.2C hot-forged steel is about 50% after annealing, as
reported in Ref. [20, 21]. Such a large difference clearly results from the addition of
Al, leading to the increase of Ae3 temperature and the suppression of the excessive
reverse transformation to austenite during IA, as suggested by D.-W. Suh et al.[7]. In
addition, the starting hot-forged microstructure results in lath-shaped austenite grains
formed after IA; whilst the cold rolled one results in granular austenite grains with a
little coarser size, whose mechanical stability should be lower than lath-shaped
austenite grains. Therefore, the addition of Al leads to a smaller fraction of austenite
retained.
When the duration of IA is either 0.5 h or 1 h, the RA fractions of 7Mn steel
remain nearly unchanged with the increase of IA temperature; when it is 5 hours,
however, the RA fraction decreases rapidly with the increase of IA temperature. Kang
et al. [22] have calculated the dependence of RA fraction of medium Mn steels on the
IA temperature via a thermodynamic model, their results indicate that the RA
fractions should firstly increase and then decrease with increasing temperature, which
is different from our data in Fig. 4c. One of the possible reasons is that their
thermodynamic calculations take into account only the partition of solute elements
during IA, rather than the size and morphologies of austenite grains that can both
affect the mechanic stability of retained austenite grains as well. The RA fraction of
the cold-rolled 7Mn steel is 5.5% as measured. It increases to about 30% after all the
studied IA treatments at 680-720℃ except for that at 720℃ for 5 h, which leads to
only 15% austenite retained. The measured RA fractions after the tensile fracture are
also given in Fig. 4c, in which a great gap between the RA fractions before and after
tensile deformation can be clearly seen. This indicates that a large fraction of austenite
grains have transformed to martensite during deformation. Particularly for the
specimens after the IA at 700℃ and 720℃, almost all the retained austenite grains
have transformed during deformation; in contrast, much larger fraction of austenite
grains, about 11~22%, are still retained in the specimens annealed at 680℃ after
fracture. This suggests that the austenite grains formed after IA at 680℃ have higher
mechanic stability than those formed at 700℃ and 720℃.
3.2 Engineering stress-strain curves and mechanical properties
Fig. 5a shows the engineering stress-strain curves of 7Mn steel after the IA at 680℃,
700℃ and 720℃ for 1 hour respectively. It can be seen that the increasing IA
temperature leads to lower yield strength (YS), defined as the lower yielding point
here, and higher UTS; whilst total elongation (TE) reaches the maximum at 700℃.
The influences of IA duration are shown in Fig. 5b. It can be seen that the longer
duration of IA makes a similar influence like higher IA temperature, i.e. it results in
lower YS and higher UTS, but TE decreases with the increase of IA duration when
annealed at 720℃.
All the data on the tensile properties of cold-rolled 7Mn steel after various IA
processes are summarized in Fig. 6. It can be seen again in Fig. 6a that the higher
temperature and the longer duration of IA both lead to the decrease of YS and increase
of UTS. It is noted that the values of YS and UTS of 7Mn steel are in the ranges of
600-950MPa and 900-1300MPa respectively. Fig. 6b shows that TE reaches the
maximum after the IA at 700℃ no matter how long the duration of IA is. The products
of UTS and TE after different IA processes are given in Fig. 6c. It is worth noting that
all the achieved products are larger than 30GPa·%. Particular, the product of UTS
(995MPa) and TE (66.1%) even reaches 66GPa·% when the 7Mn steel is annealed at
700℃ for 1 hour. Such tensile properties are much better than the Al-free 7Mn steel
with similar compositions [3~6, 20, 21], which shall be discussed latter.
4. Discussion
4.1 Effect of annealing process on austenite stability of 7Mn steel
The stability of austenite grains formed during IA varies with both the temperature
and duration of IA, as indicated by the fractions of RA that have transformed during
deformation in Fig. 4c. It is well known that the stability of austenite is governed by
chemical composition [23], grain size [24] and austenite morphology [25]. Here we
try to quantify the stability of austenite grains formed during different IA processes
using the Ms temperature, which generally depends on compositions. The average Mn
and Al contents of austenite grains have been measured by STEM/EDS and are listed
in Table 1. It is noted that the higher IA temperature leads to decreasing Mn content
but the duration of IA has little influence. The latter is different from our previous
research on the 5Mn-0.2C steel [26], in which the Mn content of austenite increases
with the prolonged annealing duration because a lot of cementite particles that have
formed during heating dissolve gradually during IA. In this case, however, a large
addition of Al into 7Mn steel could suppress the precipitation of cementite during
heating, leading to almost the constant contents of Mn and C of austenite grains
during IA.
The measured Mn and Al concentration in Table 1 are quite close to the calculated
equilibrium Mn and Al concentration, which are 11.73% Mn-1.99% Al at 680℃, 10.78%
Mn- 2.06% Al at 700℃ and 10.02% Mn- 2.13% Al at 720℃. This indicates that the
reverse austenite transformation is almost close to equilibrium during IA. The
calculated C contents by Equation 2, however, are a little lower than the equilibrium
C contents, which is related to the fitting accuracy of Equation 2. The Ms temperature
can be calculated by the following equation [27]:
Ms(℃)=539-423C-30.4Mn+30Al….……………………...………………………....(3)
where C, Mn and Al are the contents of C, Mn and Al in mass percentage. It can be
seen from Table 1 that Ms temperature increases with the increment of IA temperature
because both Mn and C contents of austenite decrease at higher temperatures. The
duration of IA has a smaller influence on Ms than the IA temperature. Since the higher
Ms temperature represents the lower stability of austenite, the RA after the IA at 680℃
should be the most stable while those after the IA at 720℃ the most unstable. In
addition, both grain size and morphology of austenite can affect the mechanic stability.
It has been found that small and lath-typed austenite grains are more stable than the
large and granular ones in general [24, 25]. Moreover, both the size and volume
fraction of granular grains, either austenite or ferrite, increase with the increases of
both IA temperature and duration, as discussed in section 3.1. Therefore, the higher
temperature and longer duration of IA lead to not only the lower C and Mn contents of
austenite grains but also more granular austenite grains, both deteriorating the
mechanic stability of austenite grains.
The IA at 680℃ for 0.5 hour produced the most stable austenite, only about 30%
of which transformed to martensite during the deformation to fracture, as shown in
Fig.4c. In contrast, the IA at 720℃ for various holdings produced the least stable
austenite grains because they have the largest austenite grain size, the lowest Mn and
C contents and austenite grains are almost all granular, they have transformed to
martensite almost completely during deformation. Particularly for the IA at 720℃ for
5 hours, some austenite grains have even transformed to martensite during air cooling
after the IA. While the specimens annealed at 700℃ have the moderate grain size of
austenite, the Mn and C concentration are at middle levels; in addition, there are a
small amount of lath-typed grains, so the stability of austenite is moderate.
4.2 Effect of austenite stability on tensile properties of 7Mn steel
With the increase of IA temperature, TE firstly increases and then decreases, but
the slopes of tensile flow curves, namely the work hardening rate (WHR), keeps
increasing after yielding, as shown in Fig. 5a. These phenomena could be interpreted
through Fig. 7, in which both WHR and true stress-strain (σ-ɛ) curves for the samples
annealed at different temperatures for 1 hour are given. The WHR vs. ε curves could
be divided into four stages, as shown in Fig. 7a-c. At the stage I, WHR decreases
rapidly most likely due to the fast dynamic recovery [28]. At the stage II, namely at
the yield point elongation, WHR fluctuates around nearly 1GPa during the entire
period of the yield point elongation. Ryu et al. [29] pointed out that TRIP effect in
medium Mn steel was mainly strain-induced, so that the passage of Lüders band can
cause heterogeneous work hardening. At the stage of III and IV, WHR first increases
to a maximum and then decreases gradually, which is similar to the observation by
Han et al. [30]. They found the unstable austenite grains with coarse size or low Mn
and C contents to transform to martensite at relatively low strain, leading to the first
peak on the WHR curve; next, the stable austenite with fine grains or high Mn and C
content transformed to martensite later. Due to the presence of serrated flow, the
WHRs in the stage IV for specimens annealed at 700℃ and 720℃ fluctuate, which is
similar to the serrated flow stress of TRIP/TWIP steels [31-34] and Al alloys [35, 36].
By comparison of Fig. 7a, 7b and 7c, we found a big difference in the mean value
of WHR at the stages of IV. The mean WHR value of specimen annealed at 680℃ is
between 1-1.8GPa, which is the lowest among the three specimens, and the specimen
begin to neck at the strain of 0.19; for the specimen annealed at 700℃ the mean WHR
value is between 1.5-2GPa and the value of WHR equals flow stress at the strain of
0.47; the specimen annealed at 720℃ has the highest mean WHR value which is
between 1.5-3.5GPa and the specimen is fractured at the strain of 0.34. In the case of
annealing duration for 1 hour, the RA grains after the IA at 680℃ have the smallest
size, the highest Mn and C contents and the largest fraction of lath-like grains, all of
them lead to the most stable austenite grains so that the smallest fraction of RA, about
40%, has transformed during deformation (Fig. 4c), resulting in the low WHR and the
limited ability to delay the onset of necking. In contrast, the IA at 720℃ for 1 hour
leads to the least stable austenite grains so that more than 90% of them transformed to
marteniste quickly during the tensile deformation, leading to the high value of WHR,
as displayed in Fig. 7c; in addition, since a large fractions of martensite is formed at
the relatively low strain and hard martensite are the potential sites for crack initiation
and propagation [37, 38], it also leads to the limited elongation. The austenite grains
in the specimen annealed at 700℃ for 1 hour have the moderate stability since their
mean grain size and Mn and C contents are at middle levels and some of the austenite
grains are even lath-shaped. The different grain sizes and morphologies of RA grains
result in different stabilities so that they may transform to martensite at different
strains of deformation, leading to the most significant TRIP effect and then the highest
elongation, as presented in Fig. 7b.
Influence of IA duration at 720℃ on the stress-strain curve can also be interpreted
by the stability of RA grains. The stability of RA grains after the IA at 720℃ for 0.5
hour is the highest, while the stability is the lowest after the IA for 5 hours, as has
been discussed in section 4.1. During deformation, the relatively stable austenite
grains could transforms to martensite with relatively low rate, resulting in higher
ductility and a lower tensile strength, as presented in Fig. 5b. In the case of the IA at
680℃, the RA grains are stable enough for all the studied IA durations, although the
5-hour IA duration should lead to more unstable RA grains than the shorter durations;
therefore, the IA at 680℃ generally leads to relatively superior TE, as shown in Fig.
6b. Although the specimen annealed at 680℃ for 5 hours has the highest WHR value,
its UTS is still lower than the specimen annealed for 0.5 hour because its YS is much
lower, as shown in Fig. 6a.
4.3 Influence of Al alloying on the tensile properties
The tensile properties of our developed steel are much superior than the Al-free
7Mn steel having the similar compositions in Ref. [20, 21, 26], which is firstly due to
austenite grains formed with different solute contents (Mn, C and Al contents) during
IA , leading austenite grains retained with different fraction and stability.
The previous researches on the medium Mn steels indicate that the Mn content of
RA is often close to equilibrium values [26]. Therefore, the equilibrium contents of
solute elements can be used to calculate Ms temperature by Eq.3. The Ms
temperatures of austenite grains in the 5Mn-0.2C, 7Mn-0.2C and 7Mn-0.3C-2Al
steels after the IA at various temperatures have been calculated to represent their
stability, as shown in Fig.8, without taking the influences of microstructure
morphologies and grain sizes into account. It is clearly seen that the austenite grains
formed after the IA of 7Mn-0.3C-2Al steel at 700℃ contain higher Mn and C content
than those in Al-free steels, resulting in relatively lower Ms temperature of RA (Fig.
8b), i.e. enhanced stability.
In the Fig. 1b of Ref. [19], the work hardening rate (WHR) curve of Al-free 7Mn
steel can be divided to three stages. The WHR first decreases, and then increases,
finally decrease again during deformation. Although the hot-forged Al-free 7Mn steel
can contain the RA fraction about 50% after the IA at 650℃ [20], the stability of
austenite grains is not high enough so that most of retained austenite grains may
transform to martensite at the early stage of deformation, as indicated by the sudden
increase of WHR at the true strain of 0.1 and then continuous decrease afterwards.
This results in high UTS but inferior plasticity. In contrast, although the measured
fraction of retained austenite grains in our Al-alloyed 7Mn steel is no more than 35%,
their austenite stability is in a proper range due to different morphology and size so
that they transforms step by step during deformation, as indicated by the multi-peaks
of WHR curve in Fig.7b. This leads to much larger plasticity at the similar level of
strength, i.e. the better combination of strength and plasticity.
5. Conclusion
The effects of IA temperature and duration on the microstructures and mechanical
properties of 7Mn-0.3C-2Al steel have been investigated, the following conclusions
can be drawn.
(1) Two types of microstructure, i.e. cell-like and lath-like, are formed during cold
rolling. They transform to granular and lamellar austenite grains respectively
during IA. The size and the volume fraction of granular grains, either austenite or
ferrite, both increase with the increase of IA temperature and duration. When the
duration is short, e.g. 0.5-1 hour, the annealing temperature has little influence on
the fractions of austenite retained after IA; whilst it is as long as 5 hours, the
increase of IA temperature leads to the significant decrease of retained austenite
fraction.
(2) Both the IA temperature and duration can affect the composition, size and
morphology of austenite grains, and all of them have significant influence on the
stability of austenite grains formed during IA. The IA at 680℃ for 0.5 hour
produces the most stable austenite grains due to the highest contents of C and Mn,
the smallest grain size and the largest fraction of lath-type grains whilst that at 720℃
for 5 hours the austenite grains with the least stability due to the lowest C and Mn
contents, the biggest grain size and almost no lath-like grains. During tensile tests,
only a small fraction of the former austenite grains have transformed to marteniste
because most of them are too stable; while almost all of the later austenite grains
have transformed very quickly, both lead to reduced ductility.
(3) In comparison with the reported Al-free 7Mn steel that has similar composition,
the RA grains in the present Al-alloyed 7Mn steel have higher mechanic stability
after the IA at 700℃ for 1 hour, although their fraction is smaller. They could
transform gradually during deformation, making the maximum contribution to the
continuous work hardening. Finally, the product of UTS and TE up to 66GPa·% is
achieved in the medium Mn steel firstly by such a lean alloying, to our
knowledge.
Acknowledgements
The authors gratefully acknowledge the joint financial support from National Natural
Science Foundation of China and Bao steel Group Co. Ltd (No. U1460203). The
authors would also like to thank Li Meng and Cunyu Wang for stimulating discussion.
References
[1] H. Y. Li, X. W. Lu, X. C. Wu, Y. A. Min, X. J. Jin, Bainitic transformation during the two-step
quenching and partitioning process in a medium carbon steel containing silicon, Mater. Sci.
Eng. A 527 (2010) 6255-6259.
[2] C. Y. Wang, J. Shi, W. Q. Cao, H. Dong, Characterization of microstructure obtained by
quenching and partitioning process in low alloy martensitic steel, Mater. Sci. Eng. A 527
(2010) 3442-3449.
[3] M. J. Merwin, Low-Carbon Manganese TRIP Steels, Mater. Sci. Forum 539-543 (2007)
4327-4332.
[4] M. J. Merwin, Microstructure and properties of cold rolled and annealed low-carbon
manganese TRIP steels, Iron & Steel Tech. 5 (2008) 66-84.
[5] W. Q. Cao, C. Wang, J. Shi, M. Q. Wang, W. J. Hui, H. Dong, Microstructure and mechanical
properties of Fe-0.2C-5Mn steel processed by ART-annealing, Mater. Sci. Eng. A 528 (2011)
6661-6666.
[6] W. Q. Cao, C. Wang, C. Y. Wang, J. Shi, M. Q. Wang, H. Dong & Y. Q. Weng, Microstructure
and mechanical property of the 3rd generation steels, Sci. China Tech. Sci. 55 (2012)
1814-1822.
[7] D. -W. Suh, S. -J. Park, T. -H. Lee, C. -S. Oh, S. -J. Kim, Influence of Al on the
Microstructural Evolution and Mechanical Behavior of Low Carbon Manganese
Transformation-Induced-Plasticity Steel, Metall. Mater. Trans. A 41 (2010) 397-408.
[8] S. -J. Park, B. Hwang, K. H. Lee, T.-H. Lee, D.-W. Suh, H. N. Han, Microstructure and tensile
behavior of duplex low density steel containing 5 mass aluminum, Scripta Mater. 68 (2013)
365-369.
[9] R. L. Miller, Ultrafine-grained microstructures and mechanical properties of alloy steels.
Metall. Trans. 3 (1972) 905-912.
[10] S. Lee, B. C. De Cooman, Tensile Behavior of Intercritically Annealed 10pct Mn Multi-phase
Steel, Metall. Mater. Trans. A 45 (2014) 709-716.
[11] S. Lee, B. C. De Cooman, Effect of the Intercritical Annealing Temperature on the
Mechanical Properties of 10Pct Mn Multi-phase Steel, Metall. Mater. Trans. A 45 (2014)
5009-5016.
[12] S. Lee, W. Woo, B. C. De Cooman, Analysis of the Tensile Behavior of 12 pct Mn
Multi-phase (a+γ) TWIP+TRIP Steel by Neutron Diffraction, Mater. Trans. A 47 (2016)
2125-2140.
[13] Z. H. Cai, H. Ding, R. D. K. Misra, Z. Y. Ying, Austenite stability and deformation behavior
in a cold-rolled transformation-induced plasticity steel with medium manganese content, Acta
Mater. 84 (2015) 229-236.
[14] O. Grassel, L. Kruger, G. Frommeyer, L. W. Meyer, High strength Fe-Mn-(Al, Si)
TRIP/TWIP steels development-properties-application, Inter. J. Plasticity 16 (2000)
1391-1409.
[15] O. Bouaziz, S. Allian, C. P. Scott, P. Cugy, D. Barbier, High manganese austenitic twinning
induced plasticity steels: A review of the microstructure properties relationships, Curr. Opin.
Solid State Mater. Sci. 15 (2011) 141-168.
[16] B. K. Jha, R. Avtar, V. S. Dwivedi, Structure-property correlation in carbon low alloy high
strength wire rod/wires containing retained austenite, Trans. Indian Inst. Met. 49(1996)
133-142.
[17] D. J. Dyson, B. Holmes, Effect of alloying additions on the lattice parameter of austenite, J.
Iron Steel Inst. 208 (1970) 469-474.
[18] R. D. K. Misra, S. Nayak, P. K.C. Venkatasurya, V. Ramuni, M. C. Somani, L. P. Karjalainen,
Nanograined/Ultrafine-Grained Structure and Tensile Deformation Behavior of Shear Phase
Reversion Induced 301 Austenitic Stainless Steel, Metall. Mater. Trans. A 41 (2010)
2162-2174.
[19] R. D. K. Misra, Z. Zhang, P. K. C. Venkatasurya, M. C. Somani, L. P. Karjalainen, Martensite
shear phase reversion induced nanograined ultrafine-grained Fe-16Cr-10Ni alloy: The effect
of interstitial alloying elements and degree of austenite stability on phase reversion, Mater.
Sci. Eng. A 527 (2010) 7779-7792.
[20] J. Shi, X. J. Sun, M. Q. Wang, W. J. H, H. Dong, W. Q. Cao, Enhanced work-hardening
behavior and mechanical properties in ultrafine-grained steels with large-fractioned
metastable austenite, Scripta Mater. 63 (2010) 815-818.
[21] C. Zhao, C. Zhang, W. Q. Cao, Z. G. Yang, Variation in retained austenite content and
mechanical properties of 0.2 C–7Mn steel after intercritical annealing, International Journal
of Minerals, Metallurgy, and Materials, 23(2016) 161-167.
[22] S. Kang, J. G. Speer, D. Krizan, D. K. Matlock, E. De Moor, Prediction of tensile properties
of intercritically annealed Al-containing 0.19C-4.5Mn (wt%) TRIP steels, Mater. Design 97
(2016) 138-146.
[23] S. -J. Lee, S. Lee, B. C. De Cooman, Martensite transformation of sub-micron retained
austenite in ultra-fine grained manganese transformation-induced plasticity steel, Inter. J.
Mater. Res. 104 (2013) 423-429.
[24] Y. -S. Jung, Y. -K. Lee, D. K. Matlock, M. C. Mataya, Effect of grain size on strain-induced
martensitic transformation start temperature in an ultrafine grained metastable austenitic steel,
Met. Mater. Int. 4 (2011) 553-556.
[25] J. Chiang, B. Lawrence, J. D. Boyd, A. K. Pilkey, Effect of microstructure on retained
austenite stability and work hardening of TRIP steels, Mater. Sci. Eng. A 528 (2011)
4516-4521.
[26] C. Wang, J. Shi, C. Y. Wang, W. J. Hui, M. Q. Wang, H. Dong, W. Q. Cao, Development of
Ultrafine Lamell ar Ferrite and Austenite Duplex Structure in 0.2C5Mn Steel during
ART-annealing, ISIJ International, 51(2011) 651-656.
[27] K. W. Andrews, Empirical formulae for the calculation of some transformation temperatures,
J. Iron Steel Inst. 203 (1965) 721-727.
[28] D. K. Wilsdorf, Theory of Plastic Deformation properties of low energy dislocation structures,
Mater. Sci. Eng. A 113(1989) 1-41.
[29] J. H. Ryu, J. I. Kim, H. S. Kim, C. -S. Oh, H. K. D. H. Bhadeshia, D.-W. Suh, Austenite
stability and heterogeneous deformation in fine-grained transformation-induced
plasticity-assisted steel, Scripta Mater. 68 (2013) 933-936.
[30] J. Han, S. -J. Lee, J. -G. Jung, Y. -K. Lee, The effects of the initial martensite microstructure
on the microstructure and tensile properties of intercritically annealed Fe–9Mn–0.05C steel,
Acta Mater. 78 (2014) 369-377.
[31] S. Hong, S. Y. Shin, J. Lee, D. H. Ahn, H. S. Kim, S. K. Kim, K. G. Chin, S. Lee, Serration
phenomena occurring during tensile tests of three high-manganese twinning-induced
plasticity (TWIP) steels, Metall. Mater. Trans. A 45 (2014) 633-646.
[32] P. D. Zavattieri, V. Savic, L. G. Hector, J. R. Fekete, W. Tong, Y. Xuan, Spatio-temporal
characteristics of the Portevin-Le Châtelier effect in austenitic steel with twinning induced
plasticity, Inter. J. Plasticity, 25 (2009) 2298-2330.
[33] S. Lee, Y. Estrin, B. C. De Cooman, Effect of the strain rate on the TRIP-TWIP transition in
austenitic Fe-12 pct Mn-0.6 pct C TWIP Steel, Metall. Mater. Trans. A, 45 (2014) 717-730.
[34] C. Jung, B. C. De Cooman, Temperature dependence of the flow stress of Fe-18Mn-0.6C-xAl
twinning induced plasticity steel, Acta Mater. 61 (2013) 6724-6735.
[35] S. H. Fu, T. Cheng, Q. C. Zhang, Q. Hu, P. T. Cao, Two mechanisms for the normal and
inverse behaviors of the critical strain for the Portevin–Le Chatelier effect, Acta Mater. 60
(2012) 6650-6656.
[36] H. F. Jiang, Q. C. Zhang, X. D. Chen, Z. J. Chen, Z. Y. Jiang, X. P. Wu, J. H. Fan, Three types
of Portevin-Le Chatelier effects Experiment and modeling, Acta Mater. 55 (2007) 2219-2228.
[37 Q. Lai, O. Bouaziz, M. Gane, L. Brassart, M. Verdier, G. Parry, A. Perlade, Y. Brechet, T.
Porden, Damage and fracture of dual-phase steels: Influence of martensite volume fraction,
Mater. Sci. Eng. A 646 (2015) 322-331.
[38] C. Lardron, E. Maire, O. Bouaziz, J. Adrien, L. Lecarme, A. Bareggi, Validation of void
growth models using X-ray microtomography characterization of damage in dual phase steels,
Acta Mater. 59 (2011) 7564-7573.
Figure captions:
Fig. 1 SEM micrographs of cold-rolled sample (a) and samples intercritically
annealed at 680℃ (b), 700℃ (c) and 720℃ (d) for 1hour.
Fig. 2 SEM micrographs of samples annealed at 700℃ for 30minutes (a) and 5 hours
(b).
Fig. 3 TEM micrographs of the cold-rolled sample and the samples annealed at 700℃
for 1hour. (a) dislocation cell-type microstructure; (b) lath-type microstructure; (c)
equiaxed grains; (d) lath typed grains. In Fig. 3c and 3d γ denotes austenite.
Fig. 4 (a) XRD spectra of samples after both various IA treatments (a) and the tensile
fracture (b). The fractions of retained austenite after annealing and fracture were both
calculated using Equation 1 and compared in (c).
Fig. 5 (a) Engineering stress-strain curves of 7Mn steel annealed at 680℃, 700℃ and
720℃ for 1 hour; (b) engineering stress-strain curves of 7Mn steel annealed at 720℃
for 0.5 hour, 1 hour and 5 hours respectively. The tensile tests were performed at room
temperature with a constant cross-head speed of 2mm/min.
Fig. 6 Mechanical properties of cold-rolled 7Mn steel annealed at different
temperatures and durations. (a) Low yield point (LYP) and ultimate tensile strength
(UTS); (b) total elongation(TE); (c) products of UTS and TE.
Fig. 7 The curves of work hardening rate vs. true strain and true stress vs. true strain
for the specimens after IA at 680℃ (a), 700℃(b) and 720℃ (c) for 1 hour.
Fig. 8 Dependence of the calculated equilibrium C, Mn and Al contents (a) and Ms
temperatures (b) of austenite phase on annealing temperatures for the 5Mn-0.2C,
7Mn-0.2C and 7Mn-0.3C-2Al steels.
(a) (b)
RD
(c) (d)
Fig. 1 SEM micrographs of cold-rolled sample (a) and samples intercritically annealed at 680℃
(b), 700℃ (c) and 720℃ (d) for 1hour.
(a) (b)
Fig. 2 SEM micrographs of samples annealed at 700℃ for 30minutes (a) and 5 hours (b).
(a) (b)
(c) (d)
Fig. 3 TEM micrographs of the cold-rolled sample and the samples annealed at 700℃ for 1hour. (a)
dislocation cell-type microstructure; (b) lath-type microstructure; (c) equiaxed grains; (d) lath
typed grains. In Fig. 3c and 3d γ denotes austenite.
Fig. 4 (a) XRD spectra of samples after both various IA treatments (a) and the tensile fracture (b).
The fractions of retained austenite after annealing and fracture were both calculated using
Equation 1 and compared in (c).
Fig. 5 (a) Engineering stress-strain curves of 7Mn steel annealed at 680℃, 700℃ and 720℃ for 1
hour; (b) engineering stress-strain curves of 7Mn steel annealed at 720℃ for 0.5 hour, 1 hour and 5
hours respectively. The tensile tests were performed at room temperature with a constant
cross-head speed of 2mm/min.
Fig. 6 Mechanical properties of cold-rolled 7Mn steel annealed at different temperatures and
durations. (a) Low yield point (LYP) and ultimate tensile strength (UTS); (b) total elongation(TE);
(c) products of UTS and TE.
Fig. 7 The curves of work hardening rate vs. true strain and true stress vs. true strain for the
specimens after IA at 680℃ (a), 700℃(b) and 720℃ (c) for 1 hour.
Fig. 8 Dependence of the calculated equilibrium C, Mn and Al contents (a) and Ms temperatures
(b) of austenite phase on annealing temperatures for the 5Mn-0.2C, 7Mn-0.2C and 7Mn-0.3C-2Al
steels.
Table 1 Chemical compositions of austenite and ferrite and Ms temperatures of austenite. The
content of Mn and Al were measured by STEM/EDS, the C content and Ms temperatures of
austenite were calculated by Equations 2 and 3 respectively. αMn and αAl denote Mn and Al content
of ferrite; γMn, γAl and γC represent the Mn, Al and C content of austenite, respectively. All the
compositions are in weight percentage.
680℃ 700℃ 720℃
0.5h 1h 5h 0.5h 1h 5h 0.5h 1h 5h
αMn(%) 3.83±0.05 3.54±0.03 3.74±0.07 3.88±0.10 4.19±0.37 3.88±0.14 4.34±0.12 4.47±0.38 4.26±0.07
αAl(%) 2.68±0.03 2.71±0.02 2.71±0.03 2.69±0.04 2.73±0.06 2.76±0.02 2.72±0.05 2.78±0.07 2.76±0.05
γMn(%) 10.85±0.60 10.70±0.88 10.74±0.57 9.34±0.86 10.01±0.54 9.21±0.30 9.52±0.85 9.18±0.34 9.05±0.48
γAl(%) 1.88±0.07 1.87±0.08 1.76±0.05 1.86±0.05 1.83±0.04 1.76±0.04 1.91±0.09 1.90±0.13 1.87±0.09
γC(%) 0.37 0.34 0.33 0.32 0.30 0.29 0.27 0.29 0.27
Ms(℃) 107.0 127.6 125.1 176.8 161.7 187.9 191.7 192.5 199.7